content
stringlengths
1
15.9M
\section{Introduction} The harmonic oscillator has a special significance in quantum mechanics due to its role as a bridge between the non-relativistic theory and quantum field theories. The equivalence of electromagnetic waves to a collection of non-relativistic harmonic oscillators provides the most direct route from quantum particles to quantum fields. There are many extra complications and difficulties to be encountered in quantum electrodynamics, but at the low energies relevant to quantum optics the direct connection to the harmonic oscillator is almost all that is required to begin the study of quantum light.\cite{loudon} Given that all quantum states of single-mode (one wave-vector, one polarization) light are states of the quantum harmonic oscillator, we can expect some important quantum states of light to have been discovered early in the twentieth century, without their significance being fully appreciated at the time. It is nevertheless impressive that what has become known as the coherent state was identified by Schr\"{o}dinger as early as 1926.\cite{sch26} It is also remarkable that a natural generalization of the coherent state to a hierarchy of similar states is rarely discussed in quantum mechanics or quantum optics. The coherent state has a Gaussian probability density whose peak follows the sinusoidal trajectory $\propto\sin(\omega t+\phi)$ of the classical particle. This leads to the description of the coherent state as a ``displaced" ground state since its probability density differs from that of the ground state only through its oscillation in time. Senitzky showed in 1954\cite{sen54} that the obvious generalization of the coherent state, in which any other energy eigenstate is displaced so that its probability density oscillates according to the classical trajectory, is also a solution for the harmonic oscillator. After brief attention in the 1950s, these displaced energy eigenstates were rediscovered in the 1970s.\cite{boi73,mar78} Roy and Singh\cite{roy82} introduced the designation ``generalized coherent states" (GCS) for Senitzky's solutions, and their properties in quantum optics were discussed by Oliveira {\it et al.}\cite{oli90} Nieto\cite{nie97} provides a helpful history of the literature on GCS, which should be supplemented by Marhic's rediscovery,\cite{mar78} often cited\cite{fuj80,mar90,eng90,won96,kry13,mah13a,mah13b} as the original reference for these states. There have also been a few subsequent rediscoveries of GCS.\cite{yan94,lop13,bohm} So-called ``squeezed" versions of GCS, in which the probability densities pulsate in time as well as following a sinusoidal trajectory, have also been studied.\cite{ple56,sat85,mol96,nie97,kry13} Interestingly, GCS have also been introduced independently in work on coupled field-matter systems, where they provide a useful set of basis states for some calculations. As our interest here is in the free oscillator, or single optical mode, we refer the reader to some relevant literature and the citations therein.\cite{cri92,iri05,iri07,mcc10} GCS are as basic and elementary as the coherent state, yet they receive scant attention in comparison. They provide a facinating set of quantum states of light that combines features of two stalwarts of quantum-optics courses: number states and coherent states. GCS are a natural and instructive accompaniment to the theory of coherent states, and as such they merit entry into the textbooks of quantum optics. Indeed, GCS are simpler than squeezed states, which are standard textbook material, and they are just as interesting from the theoretical point of view. Our aim here is to set out the basic quantum optics of GCS, in a manner accessible to any student with an elementary knowledge of single-mode quantum light. Currently the best single source for the quantum optics of GCS is Oliveira {\it et al.}'s analysis,\cite{oli90} but some of the formulae there can be simplified and additional visualisations can be given. Moreover, the behaviour of GCS at a beamsplitter, which we discuss below, provides perhaps the most striking demonstration of how GCS combine properties of coherent states and number states. \section{Generalized coherent states (GCS)} We consider the quantum harmonic oscillator with unit mass ($m=1$) and Hamiltonian \begin{equation} \label{Hho} \hat{H}=\frac{1}{2}{\hat{p}}^2+\frac{1}{2}\omega^2{\hat{x}}^2, \end{equation} and throughout we set $\hbar=1$. The following wave function is a solution of the time-dependent Schr\"{o}dinger equation for this oscillator:\cite{sen54,mar78} \begin{align} \psi_{n,\alpha}(x,t)=& \frac{(\omega/\pi)^{\frac{1}{4}}}{2^{\frac{n}{2}}\sqrt{n!}}e^{-\frac{\omega}{2}\left(x-\langle \hat{x}\rangle\right)^2} H_n[\sqrt{\omega}(x-\langle \hat{x}\rangle)] \nonumber \\ &\times e^{i\left[-\left(n+\frac{1}{2}\right)\omega t+x\langle \hat{p}\rangle-\frac{1}{2}\langle \hat{x}\rangle\langle \hat{p}\rangle\right]}, \label{GCS} \end{align} where $\langle \hat{x}\rangle$ and $\langle \hat{p}\rangle$ are the time-dependent expectation values of position and momentum in this state $\psi_{n,\alpha}(x,t)=\langle x|n,\alpha\rangle$: \begin{align} \langle \hat{x}\rangle=&\langle n,\alpha| \hat{x}|n,\alpha\rangle=\sqrt{\frac{2}{\omega}}|\alpha|\cos(\omega t-\theta), \label{xav} \\ \langle \hat{p}\rangle=&\langle n,\alpha| \hat{p}|n,\alpha\rangle=-\sqrt{2\omega}|\alpha|\sin(\omega t-\theta). \label{pav} \end{align} In (\ref{GCS}), $n$ is a non-negative integer and $H_n(z)$ are the Hermite polynomials; the solution (\ref{GCS}) also depends on an arbitrary constant complex number $\alpha=|\alpha|e^{i\theta}$ that appears in (\ref{xav}) and (\ref{pav}). Note that (\ref{GCS}) is separated into an $(x,t)$-dependent amplitude and an $(x,t)$-dependent phase factor. When $\alpha=0$, the solution (\ref{GCS}) reduces to the energy eigenstates $\psi_{n,0}(x,t)$ labelled by $n$ (including their time dependence). The set of solutions (\ref{GCS}) thus differs from the energy eigenstates in the following manner: the $x$-dependence of the amplitude is displaced by the time-dependent term (\ref{xav}) and the phase is displaced by an amount $x\langle \hat{p}\rangle-\frac{1}{2}\langle \hat{x}\rangle\langle \hat{p}\rangle$. The probability density of (\ref{GCS}) (the square of the amplitude) therefore maintains exactly the shape of the static density $|\psi_{n,0}(x,t)|^2$, but performs an oscillation in time that follows a classical trajectory. The dynamics of the states $\psi_{n,\alpha}(x,t)$ are depicted in Figs.~\ref{fig:gcs0}--\ref{fig:gcs2} for $n=0,1,\ \text{and}\ 2$. \begin{figure}[h!] \centering \includegraphics[width=8.6cm]{fig1a.pdf} \includegraphics[width=8.6cm]{fig1b.pdf} \caption{The coherent state $\psi_{0,\alpha}(x,t)$ with $\alpha=3$ and $\omega=1$. Top: the real and imaginary parts of the wave function plotted for one period $0\leq t\leq 2\pi/\omega$. Note that the wave function acquires an overall sign change after one period $2\pi/\omega$; two periods are required for the wave function to repeat because of the zero-point energy term $\omega t/2$ in the phase. Also shown is the classical trajectory with the same amplitude of oscillation. Bottom: the probability density. } \label{fig:gcs0} \end{figure} \begin{figure}[h!] \centering \includegraphics[width=8.6cm]{fig2a.pdf} \includegraphics[width=8.6cm]{fig2b.pdf} \caption{The first generalized coherent state $\psi_{1,\alpha}(x,t)$. Parameters are the same as in Fig.~\ref{fig:gcs0}. } \label{fig:gcs1} \end{figure} \begin{figure}[h!] \centering \includegraphics[width=8.6cm]{fig3a.pdf} \includegraphics[width=8.6cm]{fig3b.pdf} \caption{The second generalized coherent state $\psi_{2,\alpha}(x,t)$. Parameters are the same as in Fig.~\ref{fig:gcs0}. } \label{fig:gcs2} \end{figure} For $n=0$, the solution (\ref{GCS}) is the coherent state. Sch\"{o}dinger's discovery of $\psi_{0,\alpha}(x,t)$ was motivated\cite{sch26} by the desire to find a quantum state that would reproduce the classical motion at macroscopic scales.\cite{cp} For the coherent state, the Gaussian spread in $x$ of the probability density becomes smaller as a fraction of the amplitude $|\alpha|\sqrt{2/\omega}$ of its oscillation as $|\alpha|$ increases. The classical motion at macroscopic energies with trajectory $x(t)=|\alpha|\sqrt{2/\omega}\cos(\omega t-\theta)$ can then be viewed as an approximation to a coherent state with large $|\alpha|$. This picture is validated in quantum optics by results that show the coherent state to be the closest state to a classical plane wave allowed by quantum mechanics.\cite{loudon} Note that the same argument concerning the spread of the probability density as a fraction of the amplitude of its oscillation can be made for any GCS, since they all feature the same Gaussian factor $e^{-\omega\left(x-\langle \hat{x}\rangle\right)^2/2}$. But the states $\psi_{n,\alpha}(x,t)$ for $n>0$ are \emph{not} classical-like states. This will become clear in the quantum-optics setting of GCS, through their measurement properties and behaviour at a beamsplitter. In fact, the most interesting aspect of GCS is how they combine a ``quantum number" $\alpha$ that imparts classical features to the state with a quantum number $n$ that imparts quantum features. The position and momentum operators for the oscillator are related to the creation and annihilation operators by \begin{equation} \label{xaad} \hat{x}=\frac{1}{\sqrt{2\omega}}\left(\hat{a}+\hat{a}^\dagger\right), \quad \hat{p}=-i\sqrt{\frac{\omega}{2}}\left(\hat{a}-\hat{a}^\dagger\right). \end{equation} Here we are using the Schr\"{o}dinger picture and have included in (\ref{GCS}) the time dependence of the energy eigenstates $|n,0\rangle$, so that we have \begin{align} \hat{a}|n,0\rangle=&\sqrt{n}\,e^{-i\omega t}|n-1,0\rangle, \label{anum1} \\ \hat{a}^\dagger|n,0\rangle=&\sqrt{n+1}\,e^{i\omega t}|n+1,0\rangle. \label{anum2} \end{align} The coherent state $|0,\alpha\rangle$ is an eigenstate of $\hat{a}$ ($=\sqrt{\omega/2}(x+\omega^{-1}d/dx)$ in the coordinate representation (\ref{GCS})): \begin{equation} \label{a0al} \hat{a}|0,\alpha\rangle=\alpha e^{-i\omega t}|0,\alpha\rangle, \end{equation} an equation usually written for $t=0$. The transition to single-mode quantum optics\cite{loudon} is achieved through replacement of the position $\hat{x}$ by the electric-field operator \begin{equation} \label{Eaad} \hat{E}(x)=\frac{1}{\sqrt{2\omega}}\left(\hat{a}e^{ikx+\pi/2}+\hat{a}^\dagger e^{-ikx-\pi/2}\right), \quad k=\omega/c, \end{equation} which is time-independent in the Schr\"{o}dinger picture. The GCS $|n,\alpha\rangle$ then have electric-field probability distributions that are equal to the probability distributions for the position of the oscillator but with $\omega t$ replaced by $\omega t-kx-\pi/2$. In Fig.~\ref{fig:gcse} we plot the electric-field probability distributions for the first three GCS, which correspond to the position probability distributions in Figs.~\ref{fig:gcs0}--\ref{fig:gcs2}. The electric field of the coherent state $|0,\alpha\rangle$ has the familiar appearance of a plane wave with a noise band. For general GCS $|n,\alpha\rangle$, the electric field shows $n+1$ noise bands separated by $n$ nodes. In the photon number states $|n,0\rangle$ these noise bands do not oscillate, i.e.\ they show no dependence on the phase $kx-\omega t+\pi/2$. Many of the properties of the number states $|n,0\rangle$ are preserved by GCS $|n,\alpha\rangle$ with $\alpha>0$, as we shall see. The number of nodes in the electric-field probability distribution is thus the important signature of these properties, not the number of photons in the state; the latter is uncertain for GCS and can be of arbitrarily large average value for any $n$. \begin{figure}[h!] \centering \includegraphics[width=7.99cm]{fig4a.pdf} \includegraphics[width=7.99cm]{fig4b.pdf} \includegraphics[width=7.99cm]{fig4c.pdf} \caption{Electric-field probability distributions in $|n,\alpha\rangle$ for $n=0,\ 1$ and $2$, with $\alpha=3$ and $\omega=c=1$. Darker colours represent higher probabilities.} \label{fig:gcse} \end{figure} \section{GCS as a basis} Using the representation (\ref{GCS}) one can verify that the orthonormality relation $\langle n,0|m,0\rangle=\delta_{nm}$ for number states is preserved for GCS with the same complex amplitude $\alpha$: \begin{equation} \langle n,\alpha|m,\alpha\rangle=\delta_{nm}. \end{equation} For given $n$, however, GCS with different complex amplitudes are not orthogonal, as is familiar for the coherent state $|0,\alpha\rangle$; the general relation is \begin{equation} \langle n,\beta|n,\alpha\rangle=e^{-(|\alpha|^2+|\beta|^2-2\alpha\beta^*)/2}\,L_n\left(|\alpha-\beta|^2\right), \end{equation} where $L_n(z)$ are the Laguerre polynomials. The overlap between two GCS with equal $n$ thus becomes exponentially small as the difference $|\alpha-\beta|$ in their amplitudes increases, just as for coherent states. Ladder operators exist for the quantum number $n$ of $|n,\alpha\rangle$ and are in fact given by displaced versions of the number-state ladder operators $\hat{a}$ and $\hat{a}^\dagger$. Defining \begin{equation} \label{aal} \hat{a}_\alpha=\hat{a}-\alpha e^{-i\omega t} \quad \Longrightarrow \quad \left[\hat{a}_\alpha,\hat{a}_\alpha^\dagger\right]=1, \end{equation} one can show using the representation (\ref{GCS}) that \begin{align} \hat{a}_\alpha|n,\alpha\rangle=&\sqrt{n}\,e^{-i\omega t}|n-1,\alpha\rangle, \label{agcs1} \\ \hat{a}^\dagger_\alpha|n,\alpha\rangle=&\sqrt{n+1}\,e^{i\omega t}|n+1,\alpha\rangle, \label{agcs2} \end{align} which are the generalizations of (\ref{anum1}) and (\ref{anum2}). The displacement (\ref{aal}) of $\hat{a}$ is effected by the displacement operator\cite{loudon} \begin{gather} \hat{D}_\alpha=\exp\left(\alpha e^{-i\omega t}\hat{a}^\dagger-\alpha^* e^{i\omega t}\hat{a}\right), \label{Ddef} \\ \hat{D}_{-\alpha}^\dagger \hat{a} \hat{D}_{-\alpha} =\hat{a}-\alpha e^{-i\omega t} =\hat{a}_\alpha, \label{Dal} \end{gather} and GCS are displaced number states: \begin{equation} \label{Dgcs} |n,\alpha\rangle=\hat{D}_\alpha |n,0\rangle, \end{equation} as is verified by showing that (\ref{Dgcs}) and (\ref{Dal}) imply (\ref{agcs1}) and (\ref{agcs2}). It should not be surprising that $|n,\alpha\rangle$ for fixed $\alpha$ form a complete basis for single-mode states (just like number states), whereas $|n,\alpha\rangle$ for fixed $n$ form an over-complete basis (just like coherent states). The relevant relations are \begin{align} \sum_{n=0}^\infty |n,\alpha\rangle \langle n,\alpha| =I, \label{comn} \\ \frac{1}{\pi}\int d^2\alpha \, |n,\alpha\rangle \langle n,\alpha| =I . \label{comal} \end{align} The completeness relation (\ref{comn}) follows immediately from (\ref{Dgcs}), but the over-completeness relation (\ref{comal}) is perhaps most easily verified using the number state expansion of $|n,\alpha\rangle$ below (equation (\ref{gcsnum})). \section{Expectation values and uncertainty relations} The results of the previous section give the electric-field expectation value and uncertainty in GCS: \begin{gather} \langle n,\alpha| \hat{E}(x) |n,\alpha\rangle =\sqrt{\frac{2}{\omega}}\,|\alpha| \cos\left(kx-\omega t+\theta+\frac{\pi}{2}\right), \label{Eexp} \\ \left(\Delta E(x,t)\right)^2=\frac{2n+1}{2\omega}, \label{deltaE} \end{gather} where $\alpha=|\alpha|e^{i\theta}$ as before. Note from (\ref{deltaE}) that the uncertainty in the electric field is identical to that in the related number state (it is independent of $\alpha$). Note that the expectation value (\ref{Eexp}) and uncertainty (\ref{deltaE}) of the electric field does not convey the nodal structure of the electric-field probability distributions of GCS, as depicted in Fig.~\ref{fig:gcse}. The quadrature operators $\hat{X}$ and $\hat{Y}$, defined by \begin{equation} \label{quad} \hat{E}(x)=\frac{2}{\sqrt{\omega}}\left[\hat{X}\cos(kx+\pi/2)+\hat{Y}\sin(kx+\pi/2)\right], \end{equation} have uncertainties in GCS that are also identical to those in the related number state, as is easily verified. For the photon number operator $\hat{N}=\hat{a}^\dagger\hat{a}$ we obtain \begin{gather} \langle n,\alpha| \hat{N} |n,\alpha\rangle =n+|\alpha|^2, \label{Nexp} \\ \left(\Delta N\right)^2=|\alpha|^2=\langle n,\alpha| \hat{N} |n,\alpha\rangle -n, \label{deltaN} \\ \frac{\Delta N}{\langle n,\alpha| \hat{N} |n,\alpha\rangle}=\frac{|\alpha|}{n+|\alpha|^2}. \label{fracdeltaN} \end{gather} Contrary to the electric field, we see from (\ref{deltaN}) that the uncertainty in photon number is independent of $n$ and is thus identical to that in a coherent state with the same complex amplitude $\alpha$. The relation between $\Delta N$ and the expectation value of $\hat{N}$ is, however, different from the coherent state for $n>0$, as is shown by (\ref{fracdeltaN}). Equation (\ref{deltaN}) shows the important property of coherent states ($n=0$) that $\left(\Delta N\right)^2$ is equal to $\langle 0,\alpha| \hat{N} |0,\alpha\rangle$. This property that the variance is equal to average value is a characteristic of the Poisson distribution, and the coherent state has indeed a Poissonian photon-number distribution\cite{loudon} (the photon-number distributions for general GCS are given in the next section). Equation (\ref{deltaN}) also shows that $\left(\Delta N\right)^2$ is less than $\langle n,\alpha| \hat{N} |n,\alpha\rangle$ for $n>0$, which means that GCS for $n>0$ exhibit sub-Poissonian fluctuations,\cite{loudon} i.e the fluctuations in photon number in measurements of the states are less than those of a Poisson distribution and thus less than those of a coherent state. The photon-number probability distributions for GCS are given below and indeed differ greatly from the Poisson distribution when $n>0$. Sub-Poissonian fluctuations are a signature of nonclassical light.\cite{loudon} It is not surprising that we find by this criterion that GCS for $n>0$ are nonclassical; their nonclassical nature was already clear in Fig.~\ref{fig:gcse}. Note from (\ref{fracdeltaN}) that for all GCS the fractional uncertainty in photon number goes to zero as the amplitude $|\alpha|$ goes to infinity; the fluctuations approach Poissonian behaviour in this limit as we then have $n+|\alpha|^2 \approx |\alpha|^2$ . \section{Number-state expansion and photon statistics} GCS can be generated from the coherent state by using the creation operator $\hat{a}^\dagger_\alpha$ (see (\ref{agcs2})), or alternatively they can be generated from number states by using the displacement operator $\hat{D}_\alpha$ (see (\ref{Dgcs})). Either procedure allows the number-state expansion of $|n,\alpha\rangle$ to be calculated. It not a trivial matter to express the result in terms of special functions but one can verify that the number-state expansion is \begin{align} |n,\alpha\rangle= & \frac{1}{\sqrt{n!}} e^{-|\alpha|^2/2} \nonumber \\ & \times \sum_{k=0}^\infty (-1)^{n+k} \sqrt{k!}(\alpha^*)^{n-k} L_k^{n-k}\left(|\alpha|^2\right) |k,0\rangle, \label{gcsnum} \end{align} where $L_k^{m}(z)$ are generalized Laguerre polynomials. From (\ref{gcsnum}) we see that the probability $P_k(n,\alpha)$ of finding $k$ photons in the GCS $|n,\alpha\rangle$ is given by \begin{equation} \label{pnk} P_k(n,\alpha)= \frac{k!}{n!}e^{-|\alpha|^2} |\alpha|^{2(n-k)}\left[ L_k^{n-k}\left(|\alpha|^2\right) \right]^2. \end{equation} For the coherent state $n=0$, equation (\ref{pnk}) is of course a Poisson distribution and $P_k(1,\alpha)$ can also be written in a simple form: \begin{align} P_k(0,\alpha)=& e^{-|\alpha|^2} \frac{|\alpha|^{2k}}{k!}, \\ P_k(1,\alpha)=&e^{-|\alpha|^2} \frac{|\alpha|^{2(k-1)}(|\alpha|^2-k)^2}{k!}. \end{align} \begin{figure}[h!] \centering \includegraphics[width=8.6cm]{fig5a.pdf} \includegraphics[width=8.6cm]{fig5b.pdf} \includegraphics[width=8.6cm]{fig5c.pdf} \caption{Photon probability distributions in $|n,\alpha\rangle$ for $n=0,\ 1$ and $2$, with $|\alpha|=10$. } \label{fig:pkn} \end{figure} The distributions $P_k(n,\alpha)$ are plotted in Fig.~\ref{fig:pkn} for $n=0,\ 1$ and $2$. The most striking feature of these distributions is a similarity of their general shapes to those of the electric-field probability densities (and thus to the probability densities of the harmonic-oscillator energy eigenstates). But note that that there are in general no nodes in the distributions $P_k(n,\alpha)$, just local minima, and the relative sizes of the local maxima in $P_k(n,\alpha)$ are also not the same as in the electric-field probability densities. Oliveira {\it et al.}\cite{oli90} have given a phase-space interpretation of the oscillations in the photon-number distributions. \section{Second-order coherence} Optical coherence is an extremely complicated subject\cite{mandel} but basic measures of coherence are not difficult to define and calculate. First-order coherence is essentially a measure of correlations of the electric field in the beam, while second-order coherence is a measure of intensity correlations. The correlation function of the electric field, the average $\langle E((x_1,t_1) E((x_2,t_2)\rangle$ of the product of the electric field at two space-time points, becomes in quantum optics an expectation value quadratic in the annihilation and creation operators (see (\ref{Eaad})). Similarly, in quantum optics the intensity correlation function $\langle I((x_1,t_1) I((x_2,t_2)\rangle$ takes the form of an expectation value quartic in $\hat{a}$ and $\hat{a}^\dagger$. The ordering of $\hat{a}$ and $\hat{a}^\dagger$ in these expectation values is chosen to correspond to expressions for measured intensities.\cite{loudon} When the resulting expectation values are normalized, so that they take the value 1 when all space-time points coincide, they form the quantum measures of first- and second-order coherence, respectively. The resulting degree of first-order coherence, denoted by $g^{(1)}(x_1,t_1;x_2,t_2)$, simplifies greatly for a plane parallel single-mode beam such as a GCS. In fact, for such beams it is always the case that $|g^{(1)}(x_1,t_1;x_2,t_2)|=1$, and the beam is said to be first-order coherent.\cite{loudon} The second-order coherence function, denoted $g^{(2)}(x_1,t_1;x_2,t_2)$, takes the following simple form for a plane parallel single-mode beam\cite{loudon} \begin{equation} g^{(2)}(x_1,t_1;x_2,t_2)=\frac{\langle \hat{a}^\dagger\hat{a}^\dagger\hat{a}\hat{a}\rangle}{\langle \hat{a}^\dagger\hat{a}\rangle^2}. \end{equation} This expression is easily rewritten in terms of the photon number operator and can then be evaluated for GCS using (\ref{Nexp}) and (\ref{deltaN}), as follows: \begin{align} g^{(2)}(x_1,t_1;x_2,t_2)&=\frac{\langle \hat{N}^2 \rangle-\langle \hat{N} \rangle}{\langle \hat{N} \rangle^2}=1+\frac{(\Delta N)^2-\langle \hat{N} \rangle}{\langle \hat{N} \rangle^2} \nonumber \\ &=1-\frac{n}{(n+|\alpha|^2)^2} \label{g2} \end{align} A beam is said to be second-order coherent if $g^{(2)}=1$ in addition to $|g^{(1)}|=1$, which for GCS occurs only for the coherent state $n=0$. For $n>0$, GCS have $g^{(2)}<1$, which is another signature of nonclassical light.\cite{loudon} In the limit $n\to\infty$, or in the limit $|\alpha|\to\infty$, GCS approach second-order coherence $g^{(2)}\to1$. For given $n$, $g^{(2)}$ for GCS is minimized by the number states $\alpha=0$. Interestingly, for given $\alpha>0$ the minimum in $g^{(2)}$ is at the value of $n$ closest to $|\alpha|^2$; from (\ref{Nexp}) we see that this corresponds to an equal contribution from $n$ and $\alpha$ to the average photon number in the state. If $|\alpha|^2$ is an integer this minimum in $g^{(2)}$ for fixed $\alpha>0$ is $g^{(2)}=1-1/(4|\alpha|^2)$. \section{Behaviour at a beamsplitter} The results so far have shown that GCS combine features of coherent states and number states. This is perhaps most vividly illustrated by the behaviour of the GCS $|n,\alpha\rangle$ at a beamsplitter, where we will find that the ``$\alpha$" part of the state behaves like a coherent state and the ``$n$" part behaves like a number state. We consider a symmetric beam splitter with input arms 1 and 2, and output arms 3 and 4 (see Fig.~\ref{fig:bs}). The beamsplitter input-output relations are given by\cite{loudon} \begin{gather} \hat{a}_3=R\hat{a}_1+T\hat{a}_2, \quad \hat{a}_4=T\hat{a}_1+R\hat{a}_2, \\ \hat{a}_1=R^*\hat{a}_3+T^*\hat{a}_4, \quad \hat{a}_2=T^*\hat{a}_3+R^*\hat{a}_4, \label{a134} \\ |R|^2+|T|^2=1, \quad RT^*+TR^*=0, \end{gather} where $R$ and $T$ are the complex reflection and transmission coefficients. These are the relations that would be satisfied classically by the complex amplitudes of the electric field in the arms of the beamsplitter.\cite{loudon} \begin{figure}[h!] \centering \includegraphics[width=4cm]{fig6.pdf} \caption{Beamsplitter with input and output arms showing the annihilation operators for the corresponding modes.} \label{fig:bs} \end{figure} Consider a GCS entering the beam splitter in arm 1, with arm 2 in the vacuum state. The input state is thus $|n,\alpha\rangle_1\, |0,0\rangle_2$, which can be written \begin{gather} |n,\alpha\rangle_1\, |0,0\rangle_2=\hat{D}_{1\alpha}\frac{1}{\sqrt{n!}}\left(e^{-i\omega t}\hat{a}_1^\dagger\right)^n|0,0\rangle_1\, |0,0\rangle_2, \label{12input} \\ \hat{D}_{1\alpha}=\exp\left(\alpha e^{-i\omega t}\hat{a}^\dagger_1-\alpha^* e^{i\omega t}\hat{a}_1\right). \label{D1al} \end{gather} In (\ref{12input}) the state $|n,\alpha\rangle_1$ is built up from the number state $|n,0\rangle_1$, created using $n$ factors of $\hat{a}_1^\dagger$, followed by the displacement operator $\hat{D}_{1\alpha}$ for arm 1, which creates the GCS from the number state as in (\ref{Dgcs}). Recall that we include the time dependence in the number states and in GCS. By substitution for $\hat{a}_1^\dagger$ using (\ref{a134}) the number state $|n,0\rangle_1$ is expressed in terms of the output modes in the standard manner:\cite{loudon} \begin{gather} |n,0\rangle_1=\frac{1}{\sqrt{n!}}\left(e^{-i\omega t}\hat{a}_1^\dagger\right)^n|0,0\rangle_1 \nonumber \\ =\sum_{m=0}^n\left(\frac{n!}{m!(n-m)!}\right)^{1/2}R^mT^{n-m}|m,0\rangle_3\,|n-m,0\rangle_4. \label{numinput} \end{gather} It is also a standard result\cite{loudon} that the displacement operator (\ref{D1al}) for arm 1 becomes a product of displacement operators in the output arms when (\ref{a134}) is employed: $\hat{D}_{1\alpha}=\hat{D}_{3R\alpha}\hat{D}_{4T\alpha}$. Using this last result and (\ref{numinput}) in (\ref{12input}) gives \begin{align} |n,\alpha\rangle_1\, |0,0\rangle_2 =& \sum_{m=0}^n\left(\frac{n!}{m!(n-m)!}\right)^{1/2}R^mT^{n-m} \nonumber \\ &\times |m,R\alpha\rangle_3\, |n-m,T\alpha\rangle_4. \label{gcsout} \end{align} For $\alpha=0$, the right-hand side of (\ref{gcsout}) gives the familiar superposition of number states in the output arms of the beamsplitter, produced when a number-state is sent into one arm. For $n=0$, the result (\ref{gcsout}) is the familiar product state of independent coherent states in the output arms, produced by a coherent-state input. For general GCS with nonzero $\alpha$ and $n$, we have in (\ref{gcsout}) the remarkable result that the quantum number $n$ of GCS behaves at a beamsplitter exactly like a number state with $n$ photons, while the ``quantum number" $\alpha$ behaves exactly like a coherent state with complex amplitude $\alpha$. The quantum number $n$ of GCS in no way measures the number of photons in the state; instead it measures the number of nodes in the electric-field probability distribution, and the number of local minima in the photon distribution. Yet the $n$ of GCS produces the same entanglement at a beamsplitter as that produced by $n$ photons. Each electric-field node in arm 1 has a probability amplitude $R$ to be reflected into arm 3 and a probability amplitude $T$ to be transmitted into arm 4, just like a photon. For the case of $n=1$ we have from (\ref{gcsout}) \begin{align} |1,\alpha\rangle_1\, |0,0\rangle_2 = R|1,R\alpha\rangle_3\, |0,T\alpha\rangle_4+T|0,R\alpha\rangle_3\, |1,T\alpha\rangle_4. \label{0alout} \end{align} Thus, if the GCS in the middle plot of Fig.~\ref{fig:gcse} is sent through a beam splitter and both output arms are measured, then one output arm will contain the same state as the input, but with reduced amplitude, while the other output arm will contain a coherent state. Results like (\ref{0alout}) are particularly interesting because the input GCS in this relation can have arbitrarily large (average) energy, yet it produces the same entanglement at the beamsplitter as a single photon. \section{Generation of GCS} Returning to the case of the quantum harmonic oscillator, let us recall the exact solution for dynamically generating Schr\"{o}dinger's coherent state from an initial ground state.\cite{merzbacher} The coherent state is generated from the ground state by acting on it with the displacement operator (equation (\ref{Dgcs}) with $n=0$). The time-development operator $\hat{T}(t,t_0)$ takes the state $|\psi(t_0)\rangle$ at time $t_0$ to the state $|\psi(t)\rangle$ at time $t$: $|\psi(t)\rangle=\hat{T}(t,t_0)|\psi(t_0)\rangle$. It follows that to generate the coherent state from the ground state, the time-development operator must be equal to the displacement operator, up to phase factors. If the oscillator is subjected to a classical external force $f(t)$, which corresponds to adding a term $f(t)\hat{x}$ to the Hamiltonian (\ref{Hho}), then the dynamics can be solved exactly and the resulting time-development operator is\cite{merzbacher} \begin{align} \hat{T}(t,t_0)=&e^{i\beta(t,t_0)}\exp\left[\zeta(t,t_0)\hat{a}^\dagger e^{-i\omega t}-\zeta^*(t,t_0)\hat{a}e^{i\omega t} \right] \nonumber \\ &\times e^{-i \hat{H}_0(t-t_0)/\hbar}, \label{Top} \end{align} where $\hat{H}_0$ is the Hamiltonian without the external driving term $f(t)\hat{x}$ (i.e. the Hamiltonian (\ref{Hho})), $\zeta(t,t_0)$ is given by \begin{equation} \zeta(t,t_0)=-\frac{i}{\sqrt{2\omega}}\int_{t_0}^tdt'\,f(t')e^{i\omega t'}, \end{equation} and the c-number real phase $\beta(t,t_0)$ is given by \begin{equation} \beta(t,t_0)=\frac{1}{2\omega}\int_{t_0}^tdt'\int_{t_0}^{t'} dt''\,f(t')f(t'')\sin[\omega(t'-t'')]. \end{equation} If the initial state $|\psi(t_0)\rangle$ is an energy eigenstate, then (\ref{Top}) shows that the driving force gives a time-development operator that is a displacement operator (\ref{Ddef}), up to a c-number phase factor. Thus the oscillator is driven into a coherent state by a classical external force if $|\psi(t_0)\rangle$ is the ground state, an important and well-known result in quantum mechanics.\cite{merzbacher} From (\ref{Dgcs}) we also find that a classical external force will drive the oscillator into the GCS $|n,\alpha\rangle$, with $\alpha$ determined by the external force and the interaction time, if the initial state is the energy eigenstate $|n,0\rangle$. The foregoing results apply also to single-mode quantum light interacting with a classical external current $\bm{j}$, since the coupling term in the Hamiltonian is $\bm{j\cdot\hat{A}}$, where $\bm{\hat{A}}$ is the vector potential operator.\cite{merzbacher} There are differences in the formulae due to the spatial dependence of the quantum fields and the fact that the current is coupled to $\bm{\hat{A}}$ rather than $\bm{\hat{E}}=-\partial_t\bm{\hat{A}}$, but the calculation goes through as before\cite{merzbacher} and thus a GCS can be generated by acting on a number state with a classical external current. This fact was noted by Oliveira {\it et al.},\cite{oli90} who point out that the preparation of number states can be carried out in a cavity, which can then be driven by a current that is classical up to negligible quantum fluctuations. It appears however that quantum-optical GCS have yet to be generated experimentally, though GCS of a quantum oscillator consisting of a trapped ion have very recently been reported.\cite{zie13} \section{Conclusions} We have given a basic description of generalized coherent states (GCS) in quantum optics. These states are an interesting and natural accompaniment to presentations of number states and coherent states. GCS combine features of number states and coherent states in a nontrivial manner. The electric-field probability distributions of GCS show the phase oscillation of coherent states but also contain nodes that behave like single photons. This allows GCS to have macroscopic energies while still having some of the properties of few-photon number states. \begin{acknowledgments} I thank S. Horsley for helpful discussions. I am also indebted to E. K. Irish for information on literature. \end{acknowledgments}
\section{Introduction} {The formalization of mathematical reasoning has long been a venerated activity, even before the invention of modern-day computers. As is well known, the use of computers to assist with formalized mathematical reasoning is quite old~\cite{davis2001history}. Early on in this history, proof search (the problem of determining whether a conclusion follows from a set of premises) was separated from proof checking (the problem of determining whether a formal derivation is in fact a derivation). In the 1960s, for instance, H.~Wang emphasized proof checking as an area where computers could potentially be of great use:} \begin{quote} While formalizing known or conjectured proofs and proving new theorems are intimately related, it is reasonable to suppose that the first type of problem is much easier for the machine. That is why the writer believes that machines may become of practical use more quickly for mathematical research, not by proving new theorems, but by formalizing and checking outlines of proofs. This proof formalization could be developed, say, from textbooks to detailed formulations more rigorous than \emph{Principia} [\emph{Mathematica}], from technical papers to textbooks, or from abstracts to technical papers.~\cite{wang1960toward} \end{quote} {Then, in the 1970s, the {\systemname{Mizar}} system was born.\footnote{It is not necessary to go into more detail about the early history of {\systemname{Mizar}} and its place in the history of automated reasoning. The present article is part of a celebration of {\systemname{Mizar}}'s first $40$ years. For an authoritative discussion of the first $30$ years, see~\cite{mizar-the-first-30-years}. For the present author, a major feature of {\systemname{Mizar}} added in the last decade is the ``XMLization'' of {\systemname{Mizar}} begun by Urban~\cite{DBLP:conf/mkm/Urban05}, and which continues to the current day (e.g.,~\cite{DBLP:conf/aisc/2012}). The XMLization of various aspects of {\systemname{Mizar}} has opened up the system to experimenters who, like the author, wish to get their hands dirty with {\systemname{Mizar}} but who do not have enough familiarity with the {\systemname{Mizar}} codebase to confidently modify the internal workings of the system.}} {{\systemname{Mizar}} is a language and a suite of tools for assisting with the formalization of mathematics quite in line with what Wang envisioned. Going even further, one can see {\systemname{Mizar}} as a platform for Wang's inferential analysis with an eye towards traditional foundations of mathematics (FOM). The present paper is a contribution along these lines.} {In more detail: {\systemname{Mizar}} is based on Tarski-Grothendieck set theory ($\TG$)~\cite{TARSKI.ABS,tarski1939wellordered}. $\TG$ is stronger than the standard Zermelo-Fraenkel set theory $\ZF$ (even $\ZFC$, in which one includes the Axiom of Choice) because in $\TG$ one can prove the existence, for any set $S$, of the so-called universe for $S$, which is a set $U$ containing $S$ closed under taking subsets and power sets and which also satisfies the condition:} \[ \text{If $X \subseteq U$ and $|X| < |U|$, then $X \in U$}. \] {The universe axiom of $\TG$ yields the existence of strongly inaccessible cardinal numbers, a principle that is independent of $\ZF$ (as well as $\ZFC$)~\cite{kunen2011set,jech2008axiom}.} {Thanks to the rather long history of {\systemname{Mizar}} and the steady guidance of its founder, Andrzej Trybulec, a rather large array of matematical theorems can be found in the {\systemname{Mizar}} Mathematical Library {\libname{MML}}~\cite{Trybulec:2013:FMM}. One could say that more or less an entire undergraduate mathematics curriculum, as well as a number of more advanced results, have been formalized in {\systemname{Mizar}}.} {From the perspective of foundations of mathematics, the {\libname{MML}} can be viewed as a large collection of theorems proved in set theory. The {\libname{MML}} can serve as the source for many potential experiments of interest for research in foundations of mathematics. Some experiements along these lines have already been carried out~\cite{Pak:2013:MLE,GrabowskiTAC:2012,DBLP:conf/cade/Urban07,DBLP:conf/mkm/Urban03,DBLP:journals/jar/Alama13,alama-mamane-urban}.} {Concerning the complexity of theorems of mathematics, one notices that usually mathematicians avoid many alternating quantifiers, that is, universals within existentials within universals\dots . Some theorems, or even definitions, seem to be somewhat inherently complex in terms of quantifier alternation, such as the definition of a differentiable function or Borel set. Commenting on the possibility of reformulating apparently difficult (many quantifier alternations) mathematical statements in more compact ways, H.~Friedman conjectured:} \begin{quote} Call a sentence of predicate calculus logically tight if and only if it is not logically equivalent to any sentence of lower complexity. The conjecture is that sentences occurring in actual mathematics are (generally) logically tight. This conjecture can be interpreted in two different ways. Firstly, without unraveling abbreviations. This seems like a safe bet since mathematical statements seem to always be quite short. Once they threaten to get long, one brings in abbreviations. Secondly, with unraveling abbreviations. Then the conjecture seems much more problematic and needs systematic investigation.~\cite{friedman1998complexity} \end{quote} {One step toward addressing this conjecture is to first of all get a sense of the actual sentence complexity of a wide range of formalized mathematical theorems. In this note we report on a experiment conducted with the {\libname{MML}} to measure the sentence complexity (defined in Section~2) of the theorems in {\systemname{Mizar}}. Section~3 reports on the data, which is discussed further in Section~4. Section~5 proposes further problems and concludes.} \section{Sentence complexity of {\systemname{Mizar}} theorems} {To measure the complexity of mathematical theorems, we need a metric that allows us to say that one mathematical statement is more complex than another. We are interested, in the first place, in measuring quantifier alternation. To accomplish this, we use the familiar notion of prenex normal form, with the help of which we can define a hierarchy of formulas.} \begin{definition}[Sentence complexity hierarchy] {Given a first-order signature $\sigma$, define $\Sigma_{0} = \Pi_{0}$ as the class of $\sigma$-atomic formulas. For a positive natural number $n$, define} \[ \Sigma_{n} \coloneqq \{ \exists x_{1}, \dots, x_{k} \phi \colon \phi \in \Pi_{n - 1} \} {\quad} \text{and} {\quad} \Pi_{n} = \coloneqq \{ \forall x_{1}, \dots, x_{k} \phi \colon \phi \in \Pi_{n - 1} \} \] A formula $\phi$ is in prenex normal form if there exists a natural number $k$ such that $\phi \in \Pi_{k}$ or $\phi \in \Sigma_{k}$. \end{definition} {Thus, from an atomic formula $\phi$ one obtains a $\Pi_{1}$ formula by prefixing $\phi$ by any (positive) number of universal quantifiers, and similarly by prefixing $\phi$ by a some existential quantifiers, one obtains a $\Sigma_{1}$ formula.} \begin{fact} {For every formula $\phi$ there exists a formula $\phi^{*}$ in prenex normal form such that $\vdash{\phi} \leftrightarrow {\phi^{*}}$.} \end{fact} {Our measure of sentence complexity is entirely syntactic. To compute the complexity of a formula, one does not need to refer to provability relative to a theory such as Peano Arithmetic or $\ZFC$, nor does one need to use structures (such as truth of a sentence in the standard structure~$\mathbf{N}$ of the natural numbers, as in the notion of arithmetical hierarchy~\cite{enderton2001mathematical}). In practice, the construction of a prenex normal form, or even the construction of all possible prenex normal forms of a given formula, is clearly bounded by the complexity of $\phi$ (can count the number of subformulas of $\phi$ for which one of the prenex normal form rewrite operations applies).} {Prenex normal form is not without its disadvantages. First of all, sentences can have multiple prenex normal forms, with distinct prefixes. Thus, a sentence could have complexity $\Pi_{1}$ or $\Sigma_{2}$. This possibility is somewhat troubling for our purposes, but we can resolve the issue by choosing, whenever there are multiple prenex normal forms, the one lowest in the prenex complexity hierarchy. (If there are multiple minimal prenex normal forms because we choose one arbitrarily.) By resolving ambiguity this way, we express a preference for simplicity, however defined~\cite{pambuccian1988simplicity}. Dealing with quantifier alternations becomes psycholigically taxing rather quickly, even with advanced training in logic and mathematics; if there is therefore a prenex normal form that has fewer quantifier alternations than another prenex normal form, is is thus not unreasonable to opt for the simpler one on the theory that, since a human formalizer wrote down the formula, he probably did not have in mind something more complicated.} {Different prenex normal forms of sentences can influence proof search done with these sentences, but such differences are of no concern here because, upon computing a minimal prenex normal form, no further computation is done with it. For a discussion, see~\cite{baaz2001normal}.} {Another disadvantage of prenex normal form is that it can render straightforwardly compehensible sentences into hard-to-understand logical gibberish. Although this is undoubtedly correct, we needn't be bothered by such rewritings because we don't actually care about the comprehensibility of prenex normal forms, but just their prefixes.} {Although the language of {\systemname{Mizar}} permits one to write statements using the standard battery of first-order connectives ($\wedge$, $\vee$, $\rightarrow$, $\leftrightarrow$) and quantifiers ($\forall$, $\exists$)~\cite{mizar-in-a-nutshell}, {\systemname{Mizar}} works internally only with $\battery$ (see~\cite{wiedijk-checker} for more details). The surface forms are rewritten into the more restricted internal form by making use of standard results on interdefinability of connectives and quantifiers for classical first-order logic~\cite{wiedijk-checker}.} {To more accurately assess the complexity of {\systemname{Mizar}} sentences, we work with the internal (semantic) representations rather than with the surface level layer of sentences that formalizers actually use. Working with the internal representations is justified because these formulas are what {\systemname{Mizar}} is actually computing with. Since existential quantifers do not appear in the internal representation, we work with them indirectly as negated universals. The prenex normal forms of a surface formula $\phi$ are therefore generally different from the prenex normal forms of $\phi$'s internal representation. {Due to the austere battery $\battery$ in which all {\systemname{Mizar}} sentences are internally formulated, some theorems have a prenex complexity somewhat greater than they would be if a different battery of connectives were chosen. For example, in the {\systemname{Mizar}} language one can write formulas such as} \begin{equation}\label{star} \forall x (\phi \rightarrow \psi), \end{equation} {is represented internally (ignoring differences between $\phi$ and $\psi$ their internal representations) as} \begin{equation}\label{star-star} \forall x \neg(\phi \wedge \neg\psi) \end{equation} {If $\phi$ is itself a universal formula, say, $\forall y \chi$, then the minimal prenex complexity of (\ref{star}) is no greater than the minimal prenex complexity of} \[ \forall x \forall y (\chi \rightarrow \psi). \] {If one postulates further that $\chi$ and $\psi$ are atomic formulas, then the complexity of (\ref{star}) is $\forall$. Turning our attention to (\ref{star-star}), we find that the complexity is $\forall\exists$.} {Going the other way around, prenex normal form can be ``abused'' into being misleadingly low:} \begin{fact} {For every theorem $\phi$ of $\ZFC$, there exists an extension by definitions $\ZFC^{\prime}$ of $\ZFC$ and an $0$-ary atomic formula $p$ such that $\proves{\ZFC^{\prime}}{p}$.} \end{fact} {There is nothing mysterious going on here: Simply introduce the new $0$-ary predicate symbol $p$ and postulate the definition} \[ \equivalence{p}{\phi}. \] {This amounts essentially to naming a theorem.} {Although it may be true that we introduce definitions whenever our statements become too complicated, somehow conventional mathematical practice tolerates, or even encourages a mild amount of complexity (quantifier alternation). Indeed, one could even claim (contra the possibility of trivialization of any theorem via ``atomization''), that some mathematical statements inherently involve some amount of quantifier alternation, such as the notion of continuity of a real function.} {The answer is that we are interested in taking {\systemname{Mizar}} theorems as-is. The fact that a human formalizer, with knowledge of logic and mathematics, formalized a theorem in a certain way, suggests that the as-formalized sentence has a mathematical naturalness to it. That is, after all, one of the central goals of the {\systemname{Mizar}} project.} \section{Results} {We computed the minimal prenex normal form complexity for all theorems across the {\libname{MML}}\footnote{(We worked with {\systemname{Mizar}} system version 8.1.02 and {\libname{MML}} version 5.20.1189.)}. For us, a ``theorem'' of the {\libname{MML}} includes rather more than just those items that are explicitly called \verb+theorem+ in the official {\systemname{Mizar}} syntax. In addition to official theorems, by ``theorem'' we mean:} \begin{itemize} \item {Toplevel propositions (``lemmas''), of various kinds:} \begin{itemize} \item {Straightforward statements (usually these are ``article-local'' lemmas that a formalizer prefers not to ``export'', that is, to make it available to any other formalizer} \item {So-called diffuse reasoning blocks} \item {Statements introduced by (in effect) existential introduction, (``Consider an $x$ such that\dots'')} \item {Type-changing statements (statements to the effect that an object ought to be seen as belonging to a type other than what it directly has~\cite{Wiedijk:2007:MST:1792233.1792261}, similar to ``casting'' in programming language terminology)} \end{itemize} \item {Properties~\cite{Naumowicz:2004:IMT}, such as the property of $\leq$ being a connected relation on the set of real numbers or the property that inversion on complex numbers is an involution} \item {Rewrite rules~\cite{Kornilowicz:2013:RRM}} \item {Identifications} \item {Compatibility of redefinitions. (An already-defined notion can be defined in a new way, in which case {\systemname{Mizar}} requires one to prove that the new definiens is equivalent to the old definiens. One can also define a notion to have a relatively general type, and then redefine it to have a new type, which also requires some justification.)} \item {Scheme instances (Schemes themselves, such as the axiom scheme of replacement or the scheme of mathematical induction, are essentially parameterized sets of theorems, and are thus not, strictly speaking, theorems.)} \item {Existence and uniqueness conditions for new function symbols (these ensure conservativity when extending the language)} \item {Existence conditions for types (another conservativity condition: in the {\systemname{Mizar}} logic, one can introduce new types only on the condition that they are provably non-empty)} \item {Definitional theorems.} \end{itemize} {Tables~\ref{tab:theorem-data-table} presents our findings, restricted to theorems; Table~\ref{tab:definition-data-table} is restricted just to definitional theorems.} \begin{table} \begin{tabular}{l|l|l|} $n$ & Number of $\Pi_{n}$ sentences & Number of $\Sigma_{n}$ sentences\\ \hline $0$ & $3254$ & $3254$\\ $1$ & $67252$ & $1374$\\ $2$ & $12596$ & $146$\\ $3$ & $5456$ & $9$\\ $4$ & $567$ & $10$\\ $5$ & $355$ & $1$\\ $6$ & $12$ & $4$\\ $7$ & $14$ & $0$\\ \hline Total & $89506$ & $4798$ \end{tabular} \caption{\label{tab:theorem-data-table}Numbers of theorems in the {\libname{MML}} by their prenex complexities.} \end{table} \begin{table} \begin{tabular}{l|l|l|} $n$ & Number of $\Pi_{n}$ sentences & Number of $\Sigma_{n}$ sentences\\ \hline $0$ & $270$ & $270$\\ $1$ & $4886$ & $0$\\ $2$ & $3541$ & $0$\\ $3$ & $1854$ & $0$\\ $4$ & $71$ & $0$\\ $5$ & $34$ & $0$\\ $6$ & $1$ & $0$\\ \hline Total & $10657$ & $270$ \end{tabular} \caption{\label{tab:definition-data-table}Numbers of definitional theorems in the {\libname{MML}} by their prenex complexities.} \end{table} {$\Pi_{m}$ and $\Sigma_{m}$, restricted to theorems, are empty for all $m > 7$. $\Pi_{m}$ and $\Sigma_{m}$, restricted to definitions, are empty for all $m > 6$ The counts for $\Pi_{0}$ and $\Sigma_{0}$ are identical in both tables because they are the same set of sentences (that is, they are ``singular propositions'', that is, they have no quantifiers.)} \section{Discussion} {Table~\ref{tab:theorem-data-table} shows clearly that the lion's share of {\systemname{Mizar}} theorems belongs to a $\Pi$ hierarchy rather than the $\Sigma$ hierarchy. This likely accords with standard mathematical practice; it does seem that genuinely existential theorems are uncommon. (Existential statements that involve parameters are, according to our analysis, not purely existential but rather belong to some level of the $\Pi$ hierarchy because the parameters are implicitly universally quantified at the outermost level of the statement.)} {In the table of definitions (Table~\ref{tab:definition-data-table}), notice that there are no $\Sigma_{k}$ definitions for $k > 0$. This is as expected, because when one follows the standard format for extension by definitions~\cite[chapter 8]{suppes1999introduction} a definitional theorem can never be a properly existential formula, that is, a $\Sigma_{k}$ for some $k > 0$. In other words, the only $k$ for which a definitional theorem belongs to $\Sigma_{k}$ is $k = 0$. Such definitions are equations that don't involve quantifiers, not even implicitly present ones. For example,} \[ 0 = \emptyset \] {(definition of the natural number $0$ as the empty set) or} \[ {\arcsin} = ({\sin} \mid \left [ -\pi/2, \pi/2 \right ])^{\smile} \] {(definition of $\arcsin$, as a relation, as the converse ($^{\smile}$) of $\sin$ (considered as a relation) composed with the closed interval $[-\pi/2,\pi/2]$), or} \[ e = \sum_{k \geq 0} \frac{1}{k!} \] {(the definition of Euler's constant $e$).} {The most complex theorems can be divided into a few classes:} \begin{itemize} \item {Theorems characterizing of a class of mathematical objects as satisfying a certain list of properties, each of which is a mildly complicated formula. When such a theorem is put into prenex normal form, it understandably can becomes rather complex (even when we take the least complicated prenex normal form);} \item {Reprepresentation-style theorems saying that every object in a class has, or is isomorphic to, a certain standard form;} \item {Existence or uniqueness conditions for functions defined with complicated definientia} \item {``Inherently'' complicated theorems.} \end{itemize} {As an example of the second type, consider the following characterization theorem for finite commutative groups in terms of direct products~\cite[Theorem 34]{GROUP_17.ABS}:} \begin{quote} {Let $G$ be a finite commutative group. Suppose $|G| > 1$. Then there exists:} \begin{itemize} \item {a non empty finite set $I$} \item {an associative group-like commutative multiplicative magma family $F$ of $I$, and} \item {a homomorphism $H_{0}$ from $\Pi F$ to $G$} \end{itemize} {such that} \begin{itemize} \item {$I = PrimeFactorization(G)$,} \item {for every element $p$ of $I$, $F(p)$ is a subgroup of $G$,} \item {$F(p) = (PrimeFactorization(G))(p)$, and} \item {for every element $p$ and $q$ of $I$ such that $p \neq q$ we have} \begin{itemize} \item {$F(p) \cap F(q) = \mathbf{1}_{G}$,} \item {$H_{0}$ is bijective, and} \item {for every $G$-valued total $I$-defined function $x$ satisfying} \begin{quote} {for every element $p$ of $I$, $x(p) \in F(p)$} \end{quote} {we have $x \in \Pi F$ and $H_{0}(x) = \Pi x$.} \end{itemize} \end{itemize} \end{quote} {Without looking up relevant definitions, it is evident that this theorems is quite complex (its complexity is $\Pi_{6}$).} {As an example of the third kind, consider the definition of the set of free variables of a formula in the language of set theory~\cite{ZF_MODEL.ABS} in which one explicitly enumerates the various kinds of formulas---equations, membership statements (``$x in y$''), negations, conjunctions, quantified statements. The definition is rather long and explicit; it is perhaps no surprise that the existence condition for function defined by a long definiens, when put into prenex normal form, is rather complicated.} {The fourth category is undefined; it remains unclear what it means to say that a theorem is inherently complex. It seems that some theorems in the {\libname{MML}} could be simplified (in terms of quantifier complexity) by postulating new definitions.} {As expected, the complexity of theorems and definitions drops off very quickly. Most theorems are straightforwardly universal ($\Pi_{1}$), though also as expected there are a fair number of universal-existential ($\Pi_{2}$) statements and even universal-existential-universal ($\Pi_{3}$). These counts can perhaps be seen as a confirmation of the abstract style of mathematics that developed mainly in the 19th century, and which seems to be especially dominant when formalizing mathematics in set theory.} \section{Conclusion and future work} {By viewing the {\systemname{Mizar}} Mathematical Library as a large collection of mathematical theorems formalized in classical first-order set theory, one can adopt an experimental approach to foundations of mathematics and concretely investigate problems. Such an approach compliments the research methods traditionally used in foundations of mathematics.} {Our starting point was the question of the sentence complexity of mathematical statements. The question is, at first blush, somewhat vague. It becomes more precise when we specify a formal theory which we take as a proxy for ``mathematics''. In foundations of mathematics, set theory is often (rightly or wrongly) regarded as such a proxy. Even more precisely, one can single out a specific set theory (e.g., $\ZFC$) and, as far as possible, work only within it. In this paper, the focus was on $\TG$, in the guise of {\systemname{Mizar}} and its associated library of formalized mathematics, the {\libname{MML}}. We looked at the definitions and theorems across the {\libname{MML}} and used prenex normal form as a measure of the complexity of these items.} {Not unexpectedly, although we gained some insight into our original problem, we learn at the same time something about our methods themselves. Thus, despite its flaws, we take prenex normal form is a measure of sentence complexity. The simple notion of quantifier alternation becomes the standard by which to compare formal sentences. Using this metric, we measued the complexity of the contents of the {\libname{MML}} (ignoring proofs) and found that, in accordance with one's informal intuition about mathematics, most mathematical statements involve, at most, relatively few quantifier alternations. We found a handful of extreme cases. Do these cases show that there are highly complex corners of mathematics, as judged simply by the complexity of the statements (again, ignoring proofs)? By investigating extreme cases, we found that high quantifier alternation can usually be accounted for by appealing to the original statement without going through the prenex normal form computation. Thus, we find that prenex normal form can perhaps overestimate (and thereby mislead us) the complexity of statements by assigning a nearly incomprehensible prefix to certain kinds of relatively benign mathematical statements.} {The {\systemname{Mizar}} language includes Fraenkel terms such as} \[ \{ f(x) : \phi(x) \} \] {and they occur often among the theorems in the {\libname{MML}}. As with definitions introduced whenever complexity threatens to become too high, Fraenkel terms have the effect of bringing down complexity. In our experiments, Fraenkel terms were treated like any other term, so they were effectively ignored. One could redo the experiement by eliminating Fraenkel terms. Presumably, the complexity of the {\libname{MML}} would increase because intuitively one uses Fraenkel terms when one wants to refer to somewhat complicated-to-describe sets in a compact way.} {One can imagine even further experiments in which not just Fraenkel terms, but in which all definitions are expanded. This would certaily increase complexity (and one would have to address the ambiguity and nondeterminism inherent in ``expanding'' defined notions), but such an experiment would make progress toward the vision of precisely measuring the ``rock bottom'' complexity of real mathematics formalized in set theory.} \begin{acknowledgements} \par{To Andrzej Trybulec.} \end{acknowledgements} \bibliographystyle{spmpsci}
\section{Introduction} Parity nonconservation (PNC) in atoms can serve as a very precise low-energy test of the standard model that is a relatively inexpensive alternative to tests performed at high energy (e.g.~at CERN). For more information regarding the history and future prospects of PNC in atoms see, e.g.~\cite{Khr91,GFrev04,DzubaReview2012}. Currently, the combination of measurements~\cite{meas1,meas2} and calculations~\cite{CsPNC89,CsPNC90,CsPNC92,CsPNC01,CsPNC02,CsPorsev,CsOur} for the $6s$-$7s$ parity-forbidden E1 transition in cesium provides the most precise atomic PNC result, leading to the best atomic test of the electroweak theory so far. It is the direct aim of such investigations to determine an experimental value for the nuclear weak charge $Q_W$, a dimensionless coupling constant quantifying the strength of the $Z^0$ exchange between the nucleus and electrons. The result of this investigation leads to an observed value of the Cs weak charge that gives a strong indication that improvements and new avenues for investigation in this field could lead to important results~\cite{CsOur,CsQw}. One way to proceed would be to try to improve the accuracy in both the measurements and calculations in cesium, though it is not expected that significant improvement could be made here in the near future. Another possibility is to look to other systems. Several proposals have been put forward to search for PNC in heavier atoms, where the PNC signal is expected to be larger (e.g.~\cite{Dzuba2011Yb,Ra,Ba+,KVI,FrPNC}), and in systems such as Rb~\cite{OurRb2012}, where the accuracy could be higher. A promising alternative is to perform measurements of PNC in a chain of isotopes~\cite{IsoChain}, where the accuracy is limited only by the knowledge of the (poorly understood) neutron distribution. In this work however, we focus our attention on another area, the measurement the $P$-odd $T$-even nuclear moment that arises due to parity violation in the nucleus, the so-called nuclear anapole moment~\cite{AnM,Flambaum1985}. The experiment~\cite{meas1} of Weiman {\em et al.} provides the only measurement of a nuclear anapole moment. Measurements of the anapole moment (ANM) could prove to be invaluable tools in the study of parity violation in the hadron sector. There is interest in measuring PNC in the $6s$-$5d_{5/2}$ transition in cesium~\cite{propose}, and the possibility of measuring PNC in this transition in Ba$^+$ and in the $7s$-$6d_{5/2}$ transition of Ra$^+$ has been discussed~\cite{Geetha1998,Sahoo2011}. In this work we perform calculations of this and similar amplitudes for several isotopes of Rb, Cs, Ba$^+$, Yb$^+$, Fr, Ra$^+$ and Ac$^{2+}$ with the hope of motivating experiment in this important area. The $s$-$d_{5/2}$ transitions have practically no contribution from the nuclear weak charge, and thus provide good systems for the extraction of the anapole moment. PNC in $s$-$d$ transitions of moderately charged ions could potentially be measured using techniques put forward by N.~Fortson~\cite{Fortson}. The prospect of using these elements in measurement of nuclear-spin-independent PNC has been discussed in our recent work~\cite{Fr-like}. \section{Theory} The effective Hamiltonian describing the parity violating electron-nucleus interaction can be expressed as the sum of the nuclear-spin-independent (SI) and nuclear-spin-dependent (SD) parts (unless otherwise stated we use atomic units, $\hbar=|{e}|=m_e=1$, $c=1/\alpha\approx137$ throughout): \begin{equation} \hat h_{\rm PNC} = \hat h_{\rm SI} + \hat h_{\rm SD} = \frac{G_F}{\sqrt{2}} \left(-\frac{Q_W}{2}\gamma_5 + \frac{\varkappa}{I} \boldsymbol{\alpha} \boldsymbol{I}\right)\rho(r), \label{eq:hpnc} \end{equation} where $G_F\approx 2.2225\times10^{-14}$ a.u. is the Fermi weak constant, $Q_W$ is the nuclear weak charge, $\boldsymbol{\alpha}=\gamma_0\boldsymbol{\gamma}$ and $\gamma_5= i \gamma_0\gamma_1\gamma_2\gamma_3$ are Dirac matrices, $I$ is the nuclear spin, $\rho(r)$ is the normalised nuclear density, $\int \rho\, {\rm d}^3 r=1$, and $\varkappa$ is a dimensionless constant that quantifies the strength of the SD interaction~\cite{NoteDefinition}. There are three main sources that contribute to $\varkappa$: (i) the interaction with the so called anapole moment of the nucleus~\cite{AnM}, this is by far the dominating effect in heavy elements; (ii) the contribution from the spin-dependent electron-nucleus weak interaction ($Z^0$ exchange), see e.g.~Ref.~\cite{Flambaum1977}; (iii) the combination of the SI-PNC contribution (i.e.~$Q_W$) with the hyperfine interaction~\cite{Flambaum1985} (see also \cite{Bouchiat1991,Johnson2003}). The contribution of the combined $Q_W$ and hyperfine effects is discussed in Section~\ref{sec:discuss}. For greater detail we direct the reader to the review~\cite{GFrev04} and the book~\cite{Khr91}. The parity-violating ``E1'' transition between two states of the same parity ($a\to b$) is given by the sum \begin{equation} \small E_{\rm PNC} = \sum_n \Big[ \frac{ \bra{b} \hat{d}_{\rm E1} \ket{n} \bra{n} \hat{h}_{\rm PNC} \ket{a}} {E_{a}-E_{n}} + \frac{ \bra{b} \hat{h}_{\rm PNC} \ket{n} \bra{n} \hat{d}_{\rm E1} \ket{a}} {E_{b}-E_{n}} \Big], \label{eq:pnc} \end{equation} where $\hat{d}_{\rm E1}$ is the electric dipole (E1) operator, and $\ket{a}\equiv\ket{J_aF_aM_a}$ with $F=I+J$ the total atomic angular momentum. With use of the Wigner-Eckart theorem the amplitude can be expressed via the reduced matrix elements: \begin{equation} E_{\rm PNC} = (-1)^{F_b - M_b} \threej{F_b}{1}{F_a}{-M_b}{q}{M_a}\bra{J_bF_b}|d_{\rm PNC}| \ket{J_aF_a}, \label{eq:pnc-rme} \end{equation} where for the SI amplitude \begin{align} &\bra{J_bF_b}| d_{\rm SI} | \ket{J_aF_a} = \frac{G_F}{2\sqrt{2}}(-Q_W)(-1)^{I+F_a + J_b+1} \notag\\ &\times\sqrt{(2F_b+1)(2F_a+1)}\sixj{J_a}{J_b}{1}{F_b}{F_a}{I} \notag\\ & \times \sum_n \Bigg[ \frac{ \bra{J_b} |\hat{d}_{\rm E1}| \ket{J_n} \bra{J_n} |\gamma_5\rho| \ket{J_a}} {E_{a}-E_{n}} \notag\\ & \phantom{\times\sum_n \Big[ \;} + \frac{ \bra{J_b} | \gamma_5\rho| \ket{J_n} \bra{J_n} |\hat{d}_{\rm E1}| \ket{J_a}} {E_{b}-E_{n}} \Bigg], \label{eq:si-pnc-rme} \end{align} and for the SD amplitude \begin{align} &\bra{J_bF_b}| d_{\rm SD} | \ket{J_aF_a} \notag\\ &= \frac{G_F}{\sqrt{2}}\varkappa \sqrt{(I+1)(2I+1)(2F_b+1)(2F_a+1)/I} \notag\\ & \times \sum_n \Bigg[ (-1)^{J_b-J_a} \sixj{J_n}{J_a}{1}{I}{I}{F_a} \sixj{J_n}{J_b}{1}{F_b}{F_a}{I} \notag\\ & \phantom{\times\sum_n \Bigg[ \;} \times \frac{\bra{J_b} |\hat{d}_{\rm E1}| \ket{J_n} \bra{J_n} |\boldsymbol{\alpha}\rho| \ket{J_a}} {E_{a}-E_{n}} \notag\\ & \phantom{\times\sum_n \Bigg[ \;} + (-1)^{F_b-F_a} \sixj{J_n}{J_b}{1}{I}{I}{F_b} \sixj{J_n}{J_a}{1}{F_a}{F_b}{I} \notag\\ & \phantom{\times\sum_n \Bigg[ \;} \times\frac{\bra{J_b} | \boldsymbol{\alpha}\rho| \ket{J_n} \bra{J_n} |\hat{d}_{\rm E1}| \ket{J_a}} {E_{b}-E_{n}}\Bigg]. \label{eq:sd-pnc-rme} \end{align} In tables we present the $z$-components: \begin{equation} E_{\rm PNC}(z) = (-1)^{F_b - F_z} \threej{F_b}{1}{F_a}{-F_z}{0}{F_z}\bra{J_bF_b}|d_{\rm PNC}| \ket{J_aF_a}, \label{eq:pnc-z} \end{equation} where we take $F_z={\rm min}(F_a,F_b).$ \section{Calculations} \label{sec:calcs} If the states $a$, $b$ and $n$ in (\ref{eq:pnc}) are the physical many-electron wavefunctions of the atom then these equations are exact and the summation is over all excited states. In calculations obviously this is not the case, we use single-electron orbitals as the wavefunctions and extend the sum over all states (the summation over core states corresponds to including the highly excited autoionization states). We begin with the relativistic Hartree-Fock (RHF) approximation, generating the single-particle orbitals in a $V^{N-1}$ potential. Core-valence correlation effects are then included using the correlation potential (CP) method~\cite{CPM}, and the polarization of the core electrons and interactions with external fields are taken into account using the time-dependent Hartree-Fock (TDHF) approximation~\cite{CPM,CPM2}. The correlation potential, an {\em ab initio}, non-local (integration), energy-dependent operator, $\hat\Sigma=\hat\Sigma(E,l,j)$, is calculated using a summation of dominating diagrams of many-body perturbation-theory (including screening of the electron-electron interaction and the particle-hole interaction) to all orders using the Feynman diagram technique and relativistic Hartree-Fock Green's functions~\cite{CPM2}. Then by solving the relativistic Hartree-Fock-like equations with the extra operator $\hat \Sigma$, \begin{equation} (\hat H_0 +\hat \Sigma - \epsilon^{(\rm BO)}_n)\psi_n^{(\rm BO)}=0, \label{eq:BO} \end{equation} we construct the Brueckner orbitals (BOs) for the valence electron. Here, $H_0$ is the RHF Hamiltonian and the index $n$ denotes valence states. Part of the missing diagrams can be expressed in terms of the energy derivatives of $\Sigma$, or can also be calculated separately. These contributions are very small for alkaline atoms but may be significant in atoms where radius of valence electron is close to the core electron radius (e.g.~in Yb$^+$). The correlation potential method is especially accurate in atoms with one electron above closed subshells (which are the topic of the present work), where it gives an accuracy of about $\sim0.1\%$ for the ionization energies of valence electron orbitals. Note that the correlation potential is calculated independently for orbitals with different $l$, $j$. Therefore, we may estimate the missing (and very small) contributions of higher order diagrams by using a simple semi-empirical procedure of rescaling the CP operator, i.e.~$\hat\Sigma\to\lambda\hat\Sigma$ in Eq.~(\ref{eq:BO}). A different parameter is chosen for each partial wave (i.e.~$ns$, $np_{1/2}$, $np_{3/2}$, $nd_{3/2}$, and $nd_{5/2}$) to reproduce exactly the experimental energies corresponding to the lowest (valence) principal quantum number for each partial wave. It should also be noted that these parameters typically differ from 1 by only a small fraction, e.g.~for cesium they are $\lambda_{s}$=$0.99$, $\lambda_{p_{1/2}}$=0.96, $\lambda_{p_{3/2}}$=0.97, $\lambda_{d_{3/2}}$=0.94, and $\lambda_{d_{5/2}}$=0.94, indicative of the already very good accuracy of the {\em ab initio} all-order CP method. This fitting makes only a small difference to most PNC amplitudes, and the difference between amplitudes calculated with and without the fitting provides a good indication of the relative size of any missed correlations and thus serves as a good estimate of the uncertainty. In the transitions here however, the uncertainty is dominated by core-polarisation effects, not the correlation potential. It is also important to note that even in cases where this fitting does make a difference its effect on the ratio of the SI to SD parts is negligible. For Yb$^+$ we use only the second-order CP method due to the more complicated electron structure. The presence of the $4f^{14}$ shell means there are other correlation effects that are larger than the all-order corrections, see e.g.~Ref.~\cite{Dzuba2011}. The second order CP operator provides reasonable accuracy as is, and the process of rescaling means the accuracy is good here also, as discussed in the next section. In the evaluation of the amplitude, the operators $\hat{d}_{\rm E1}$ and $\hat{h}_{\rm PNC}$ in Eq.~(\ref{eq:pnc}) are modified to include the effect of the polarization of the core electrons due to the interaction with the external E1 and weak fields: $\hat{d}_{\rm E1} \rightarrow \hat{d}_{\rm E1} + \delta V_{\rm E1}$ and $\hat{h}_{\rm PNC} \rightarrow \hat{h}_{\rm PNC} + \delta V_{\rm PNC}$. Here $\delta V_{\rm E1}$ ($\delta V_{\rm PNC}$) is the modification to the RHF potential due to the effect of the external field $\hat{d}_{E1}$ ($\hat h_{\rm PNC}$). In the TDHF method, the single-electron orbitals are perturbed in the form $\psi=\psi_0+\delta\psi$ where $\psi_0$ is an eigenstate of the RHF Hamiltonian, and $\delta\psi$ is the correction due to the external field. The corrections to the potential are then found by solving the set of self-consistent TDHF equations for the core states: \begin{equation} (\hat H_0 -\varepsilon_c)\delta \psi_{c} = -(\hat f +\delta V_f)\psi_{0c}, \label{eq:RPA} \end{equation} where the index $c$ denotes core states and $\hat f$ is the operator of external field (be that $\hat{d}_{E1}$ or $\hat h_{\rm PNC}$). Note that the approach described above does not take into account the effect of core polarization due to simultaneous action of the weak and E1 fields. This `double-core-polarization' (DCP) effect was the study of our recent work, Ref.~\cite{DCP}. Accurate calculations would require the use of the `solving equations' approach (see e.g.~\cite{CsPNC02}), a more numerically stable method based on solving differential equations, which includes the DCP contribution. However, since high accuracy is not needed for the SD-PNC, we use simpler approach which is based on a direct summation over states. We use Ref.~\cite{DCP} to include the DCP correction into the SI amplitudes, but do not include this term into the SD amplitudes since the accuracy of analysis is less important here. To use the direct-summation method, we employ the B-spline technique~\cite{B-spline} to construct the set of single-electron orbitals used for the summation in Eq.~(\ref{eq:pnc}), as well as for the calculation of $\hat \Sigma$. The states used in the calculation of $\hat \Sigma$ are linear combinations of the B-splines which are eigenstates of the RHF Hamiltonian, whereas those used for the evaluation of (\ref{eq:pnc}) are the Brueckner orbitals (eigenstates of the $\hat H_0 + \hat \Sigma$ Hamiltonian). For the summation we use 90 B-splines of order 9 for each partial wave in a cavity of radius 75~$a_0$. \subsection{Accuracy of the calculations} \begin{table \centerin \caption{Calculated ionization energies for cesium in various approximations and comparison with experiment (Ref.~\cite{NIST}). Blank means calculated value matches exactly with experiment by construction. Units: cm${}^{-1}$. } \begin{ruledtabular \begin{tabular}{rrrrrr} \multicolumn{1}{c}{Level } & \multicolumn{1}{c}{ $\Sigma^{(2)}$ } & \multicolumn{1}{c}{ $\lambda\Sigma^{(2)}$ } & \multicolumn{1}{c}{ $\Sigma^{(\infty)}$ } & \multicolumn{1}{c}{ $\lambda\Sigma^{(\infty)}$ } & Exp. \\ \hline $6s_{1/2}$ & -32416 & & -31457 & & -31406 \\ $6p_{1/2}$ & -20539 & & -20290 & & -20228 \\ $6p_{3/2}$ & -19940 & & -19722 & & -19674 \\ $5d_{3/2}$ & -17567 & & -17146 & & -16907 \\ $5d_{5/2}$ & -17407 & & -17030 & & -16810 \\ $7s_{1/2}$ & -13024 & -12832 & -12827 & -12817 & -12871 \\ $7p_{1/2}$ & -9710 & -9628 & -9640 & -9624 & -9641 \\ $7p_{3/2}$ & -9521 & -9448 & -9458 & -9445 & -9460 \\ $8p_{1/2}$ & -5724 & -5689 & -5694 & -5687 & -5698 \\ $8p_{3/2}$ & -5639 & -5607 & -5611 & -5606 & -5615 \\ \end{tabular \end{ruledtabular \label{tab:EnvsExp \end{table Without any rescaling of the correlation potential (see Section~\ref{sec:calcs}) our energies agree with experiment to around 0.1\%-0.5\% for most levels, and the important $s$-$p$ intervals are reproduced to about 0.3\%. A detailed analysis of the accuracy in these systems has also been performed in our recent papers, Ref.~\cite{Fr-like,Dzuba2011}, where we present calculations for the same atoms and ions investigated here. In Table~\ref{tab:EnvsExp} we present calculated energy levels for cesium using the second-order ($\Sigma^{(2)}$) and the all-order ($\Sigma^{(\infty)}$) CP method, both with and without scaling. Table~\ref{tab:intervalvsExp} presents the percentage discrepancies for the relevant energy intervals in cesium. This shows the small effect that scaling has directly on the energies, but the relatively large improvements it makes on the intervals. The rescaling of the correlation potential helps to numerically stabilize the results. The rescaling means there is no significant loss in the accuracy for the energy-levels when using $\Sigma^{(2)}$ instead of $\Sigma^{(\infty)}$. This is important for the case of Yb$^+$, where only the second-order correlation potential was used. \begin{table \centerin \caption{Percentage variation between the experimental (from~\cite{NIST}) energy intervals of relevance to parity nonconservation in cesium and calculations in various approximations. Blank means calculated value matches exactly with experiment by construction.} \begin{ruledtabular \begin{tabular}{ldddd} \multicolumn{1}{c}{Interval} & \multicolumn{1}{r}{$\Sigma^{(2)}$} & \multicolumn{1}{r}{$\lambda\Sigma^{(2)}$} & \multicolumn{1}{r}{$\Sigma^{(\infty)}$} & \multicolumn{1}{r}{$\lambda\Sigma^{(\infty)}$} \\ \hline $6s_{1/2}-6p_{1/2}$ & 6.35 & & -0.10 & \\ $6s_{1/2}-6p_{3/2}$ & 6.31 & & 0.02 & \\ $6s_{1/2}-7p_{1/2}$ & 4.08 & -0.02 & 0.24 & 0.08 \\ $6s_{1/2}-7p_{3/2}$ & 4.08 & -0.01 & 0.24 & 0.07 \\ $5d_{3/2}-6p_{1/2}$ & -5.17 & & -5.35 & \\ $5d_{3/2}-6p_{3/2}$ & -7.32 & & -6.92 & \\ $5d_{3/2}-7p_{1/2}$ & 4.83 & -0.05 & 3.30 & 0.24 \\ $5d_{3/2}-7p_{3/2}$ & 4.81 & -0.04 & 3.24 & 0.20 \\ $5d_{5/2}-6p_{3/2}$ & -5.54 & & -6.02 & \\ $5d_{5/2}-7p_{3/2}$ & 4.28 & -0.04 & 3.02 & 0.20 \\ \end{tabular \end{ruledtabular \label{tab:intervalvsExp \end{table \begin{table* \centerin \caption{Calculations of reduced matrix elements (a.u.) of electric dipole transitions of interest to PNC studies in cesium and comparison with experiment. The last column shows the percentage difference between final calculations (using the rescaled all-order correlation potential, $\lambda\Sigma^{(\infty)}$) and experiment. } \begin{ruledtabular \begin{tabular}{ldddddld} & \multicolumn{4}{c}{This work } & \multicolumn{3}{c}{ Experiment} \\ \cline{2-5}\cline{6-8} \multicolumn{1}{c}{Transition} & \multicolumn{1}{r}{ $\Sigma^{(2)}$ } & \multicolumn{1}{r}{ $\lambda\Sigma^{(2)}$ } & \multicolumn{1}{r}{ $\Sigma^{(\infty)}$ } & \multicolumn{1}{r}{ $\lambda\Sigma^{(\infty)}$ } & \multicolumn{1}{c}{ Value } & \multicolumn{1}{c}{ Ref. } & \multicolumn{1}{c}{ \% Diff. } \\ \hline $6s_{1/2}-6p_{1/2}$ & 4.387 & 4.503 & 4.506 & 4.512 & 4.4890(65) & \cite{rafac} & 0.51 \\ & & & & & 4.5097(74) & \cite{young} & 0.05 \\ $6s_{1/2}-6p_{3/2}$ & 6.170 & 6.337 & 6.343 & 6.351 & 6.3238(73) & \cite{rafac}& 0.42 \\ & & & & & 6.3403(64) & \cite{young} & 0.16 \\ $6s_{1/2}-7p_{1/2}$ & 0.2995 & 0.2744 & 0.2645 & 0.2724 & 0.2757(20) & \cite{vasilyev} & 1.19 \\ & & & & & 0.2825(20) & \cite{shabanova,vasilyev}& 3.56 \\ $6s_{1/2}-7p_{3/2}$ & 0.6050 & 0.5686 & 0.5581 & 0.5659 & 0.5795(100) & \cite{vasilyev} & 2.34 \\ & & & & & 0.5856(50) & \cite{vasilyev} & 3.36 \\ $5d_{3/2}-6p_{1/2}$ & 6.744 & 7.039 & 6.927 & 7.032 & 7.33(6) & \cite{diBerardino} & 4.07 \\ $5d_{3/2}-6p_{3/2}$ & 3.037 & 3.173 & 3.121 & 3.170 & 3.28(3) & \cite{diBerardino} & 3.37 \\ $5d_{5/2}-6p_{3/2}$ & 9.254 & 9.629 & 9.481 & 9.616 & 9.91(3) & \cite{diBerardino} & 2.97 \\ \end{tabular \end{ruledtabular \label{tab:MEvsExpCs \end{table* In Table~\ref{tab:MEvsExpCs} we compare calculations of several of the relevant $E1$ reduced matrix elements for cesium with their corresponding experimental values. This shows very good agreement with experiment, to better than 0.5\% for the lowest $s$-$p$ transitions, and better than 5\% for the transitions involving $d$ and higher $p$ states. Again, we present calculations using the second-order ($\Sigma^{(2)}$) and the all-order ($\Sigma^{(\infty)}$) CP method, both with and without scaling. We demonstrate that by including the rescaling of the correlation potential we can correct for the discrepancies that arise from using the second-order correlation potential, effectively meaning that the rescaled second order CP is practically as good as using the all-order method. \begin{table \centerin \caption{Calculated reduced matrix elements (a.u.) for electric dipole transitions of interest in Ba$^+$ and Yb$^+$ and comparison with experiment where available.} \begin{ruledtabular \begin{tabular}{lllllll} & \multicolumn{3}{c}{Ba$^+$} & \multicolumn{2}{c}{Yb$^+$} \\ \cline{2-4}\cline{5-7} \multicolumn{1}{c}{Transition} & \multicolumn{1}{c}{Calc.} & \multicolumn{2}{c}{ Exp. } & \multicolumn{1}{c}{Calc.} & \multicolumn{2}{c}{ Exp. } \\ \hline \smallspace $6s_{1/2}-6p_{1/2}$ & 3.322 & 3.36(4) & \cite{Davidson} & 2.705 & 2.471(3) & \cite{Olmschenk} \\ $6s_{1/2}-6p_{3/2}$ & 4.690 & 4.55(10) & \cite{Davidson} & 3.817 & 3.36(2) & \cite{Pinnington} \\ $5d_{3/2}-6p_{1/2}$ & 3.063 & 3.03(9) & \cite{Kastberg} & 3.094 & 2.97(4) & \cite{Olmschenk} \\ & & 3.14(8) & \cite{Sherman} & & & \\ & & 2.90(9) & \cite{Davidson} & & & \\ $5d_{3/2}-6p_{3/2}$ & 1.338 & 1.36(4) & \cite{Kastberg} & 1.366 & & \\ & & 1.54(19) & \cite{Davidson}& & & \\ $5d_{5/2}-6p_{3/2}$ & 4.127 & 4.15(20) & \cite{Kastberg} & 4.271 & & \\ \end{tabular \end{ruledtabular \label{tab:MEvsExpBaYb \end{table We present $E1$ reduced matrix elements for Ba$^+$ and Yb$^+$ in Table~\ref{tab:MEvsExpBaYb}, along with experimental values for comparison where available. This demonstrates very good agreement between our calculations and experiment for Ba$^+$, and reasonably good agreement for Yb$^+$. The discrepancies for the Yb$^+$ values, on the order of 5\% -- 10\%, are due mainly to the more complicated electron structure due to the closeness of the $4f^{14}$ core shell to the valence $6s$ state. The most important $E1$ transition for the $6s$-$5d_{3/2}$ PNC amplitude in Yb$^+$ is the $p_{1/2}$-$d_{3/2}$ transition. This transition corresponds to the weak $s$-$p_{1/2}$ mixing, which dominates the amplitude. This $p_{1/2}$-$d_{3/2}$ $E1$ matrix element agrees with experiment to about 4\%. However, for the $6s$-$5d_{5/2}$ PNC amplitude considered here, the most important $E1$ amplitudes are the $s$-$p_{3/2}$ and $p_{3/2}$-$d_{5/2}$ transitions. The $s$-$p_{3/2}$ amplitude agrees to only 13\% with experiment, and an experimental value for the $p_{3/2}$-$d_{5/2}$ transition is, to the best of our knowledge, not known. \begin{table* \centerin \caption{Calculated magnetic dipole hyperfine constants $A$ (MHz) for the lowest valence states of Cs, Ba$^+$ and Yb$^+$, and a comparison with experiment.} \begin{ruledtabular \begin{tabular}{lrrlrrlrrl} & \multicolumn{3}{c}{${}^{133}$Cs} & \multicolumn{3}{c}{${}^{135}$Ba$^+$} & \multicolumn{3}{c}{${}^{171}$Yb$^+$} \\ \cline{2-4}\cline{5-7}\cline{8-10} \multicolumn{1}{c}{Level} & \multicolumn{1}{c}{Calc.} & \multicolumn{2}{c}{ Exp. } & \multicolumn{1}{c}{Calc.} & \multicolumn{2}{c}{ Exp. } & \multicolumn{1}{c}{Calc.} & \multicolumn{2}{c}{ Exp. } \\ \hline \smallspace $s_{1/2}$ & 2315 & 2298.2 & \cite{Arimondo} & 3674 & 3593.3(22) & \cite{Wendt} & 13202 & 12645(2) & \cite{Martensson-Pendrill}\\ $p_{1/2}$ & 290 & 291.89(8) & \cite{RafacTanner} & 668 & 664.6(3) & \cite{Villemoes} & 2515 & 2104.9(13) & \cite{Martensson-Pendrill}\\ \end{tabular \end{ruledtabular \label{tab:hfs \end{table* The accuracy of the weak-charge and anapole-moment induced PNC interaction matrix elements relies on the accuracy of the wavefunctions at short distances (near the nucleus). One way to test the accuracy of the wavefunctions at this distance scale is to calculate magnetic dipole hyperfine structure constants, which also depend on the wavefunctions close to the nucleus. The hyperfine structure constants are typically reproduced very well for $s$ and $p$ states, but not so well for $d$ states (see, e.g.~Ref.~\cite{Dzuba2011}). The direct applicability of using hyperfine structure calculations as a test for $p$-$d$ $ h_{\rm PNC}$ matrix elements has not been fully investigated, and will be the focus of future work. The uncertainty in the calculations of the hyperfine structure constants is dominated by core polarization, which is much larger for the hyperfine constants than for the weak matrix elements. The implication of this is that the accuracy of the $s$-$p$ PNC interaction matrix elements can be high, and importantly can be controlled by computing hyperfine constants. For the $p$-$d$ weak matrix elements, however, there is no guarantee of high accuracy, and it is not clear how the accuracy can be reliably judged. In Table~\ref{tab:hfs} we present calculations of magnetic-dipole hyperfine structure constants $A$, for the $6s$ and $6p_{1/2}$ states of Cs, Ba$^+$ and Yb$^+$, along with experimental values for comparison. The $h_{\rm PNC}$ interaction, to lowest order, is effectively a contact interaction and as such only significantly mixes $s$ and $p_{1/2}$ states. Due to core polarization, however, mixing between $s$ and $p_{3/2}$ states, as well as between $p_{3/2}$ and $d_{3/2,5/2}$ states, is not so small. For $s$-$s$ PNC amplitudes there is nothing to worry about, since these contain only terms involving $s$-$p_{1/2}$ mixing. The $s$-$d_{3/2}$ amplitudes contain also terms involving $p$-$d$ mixing, however the $s$-$p_{1/2}$ mixing is many times larger, meaning that these amplitudes are dominated by the $s$-$p_{1/2}$ mixing terms, which contribute between 70\% and 90\% to the total amplitude. For the spin-independent amplitudes ($s$-$d_{3/2}$), the accuracy should be about 1-2\% (see Ref.~\cite{Fr-like}). This is due to the very good agreement with energy levels, hyperfine structure constants and matrix elements. The spin-dependent parts of the $s$-$d_{3/2}$ amplitudes are likely to be somewhat less accurate, due mainly to core-polarization effects and the larger number of contributing states (since the spin-dependent PNC interaction can mix states with $\Delta J=1$). Because of this, without the double-core-polarization contribution, the accuracy for these amplitudes is likely to be between 5\% and 10\%. For the $s$-$d_{5/2}$ amplitudes, there are no $s$-$p_{1/2}$ mixing terms, instead there are terms involving $s$-$p_{3/2}$ and $p_{3/2}$-$d_{5/2}$ mixing. Due to core-polarization, there is no significant difference between the extent of the PNC mixing between these two contributions, and the size of the respective matrix elements is roughly the same. For the part of these PNC amplitudes coming from the $s$-$p_{3/2}$ mixing, i.e.~the first term in Eq.~(\ref{eq:pnc}), the accuracy is likely to be good. However, for the contribution from the $6p_{3/2}$-$5d_{5/2}$ $\hat h_{\rm SD}$ matrix elements the accuracy is likely to be significantly worse. There is not enough information to determine reliably how accurate the $p_{3/2}$-$d_{5/2}$ $\hat h_{\rm SD}$ matrix elements are, and as such the $s$-$d_{5/2}$ SD-PNC amplitudes should be considered order-of-magnitude estimates. This low level of accuracy is sufficient for the purpose of the current work, which is to demonstrate the magnitude and relative sizes of these transitions in different elements. Note also that the very high accuracy that is required of the SI-PNC calculations for the extraction of the nuclear weak charge is not required in the search for anapole moments. In Table~\ref{tab:compare} we compare our calculations of the SD-PNC amplitudes in Ba$^+$ and Ra$^+$ with several of those available in the literature. The agreement between results for the $s$-$d_{3/2}$ transitions is reasonable. For the $s$-$d_{5/2}$ we agree with calculations of Ref.~\cite{Geetha1998} but not of Ref.~\cite{Sahoo2011}. For atoms and ions similar to Yb$^+$, in which an external electron is close to the core and strongly interacts with its electrons, a different higher-order effect described by the so-called `ladder diagrams'~\cite{ladder} becomes important. The inclusion of ladder diagrams also significantly improves the accuracy of calculations in ions, for which the valence electrons lie closer to the core, and improves the accuracy of the $d$-states for atoms and ions, see e.g.~\cite{Dzuba13,Fr-like}. With the inclusion of ladder diagrams, as well as the double-core-polarization effect, the accuracy for these calculations can potentially approach the level of several percent, though this would need further investigation. The accuracy could then be further improved by including the Breit~\cite{Breit} and QED~\cite{QED} corrections, as well as higher order non-Brueckner electron correlations, such as structure radiation, the weak correlation potential and renormalization of states (see e.g. Ref.~\cite{CPM2,CsPNC02}). \begin{table \centerin \caption{ Reduced matrix elements $\bra{J_b,F_b}|d_{\rm SD}|\ket{J_a,F_a}$ of the spin-dependent PNC amplitudes of Ba$^+$ and Ra$^+$ and comparison with other works~\cite{NoteDefinition}. Units: $10^{-13}ea_0\varkappa$.} \begin{ruledtabular \begin{tabular}{ldcddl} & & \multicolumn{4}{c}{$E_{\rm PNC}$} \\ \cline{3-6} \smallspace & I & \multicolumn{1}{c}{Transition} & \multicolumn{1}{c}{This work} & \multicolumn{2}{c}{Others} \\ \hline \smallspace $^{135}$Ba$^+$ & 1.5 & $\bra{5d_{5/2},3}|d_{\rm SD}|\ket{6s,2}$ & 0.85 & 0.82& \cite{Geetha1998}\\ & & & & 0.274 &\cite{Sahoo2011}\\ \smallspace & & $\bra{5d_{3/2},3}|d_{\rm SD}|\ket{6s,2}$ & 17.15 & 19.44 & \cite{Sahoo2011}\\ \smallspace $^{223}$Ra$^+$ & 1.5 & $\bra{6d_{5/2},3}|d_{\rm SD}|\ket{7s,2}$ & 11.4 & 12.7& \cite{Geetha1998} \\ & & & & 3.504 &\cite{Sahoo2011}\\ \smallspace & & $\bra{5d_{3/2},3}|d_{\rm SD}|\ket{6s,2}$ & 210.9 & 234.690 & \cite{Sahoo2011}\\ \end{tabular \end{ruledtabular \label{tab:compare \end{table \section{Results and discussion}\label{sec:discuss} \begin{table \centerin \caption{ SD-PNC amplitudes of the $\ket{5sF_a}\to\ket{4d_{5/2}F_b}$ transition in Rb, and the $\ket{6sF_a}\to\ket{5d_{5/2}F_b}$ transitions in Cs, Ba$^+$ and Yb$^+$. Both the reduced matrix elements (RME) and the $z$ components are shown. Units: $10^{-13}ea_0\varkappa$. } \begin{ruledtabular \begin{tabular}{lllldd} & & & & \multicolumn{2}{c}{$E_{{\rm PNC}}$} \\ \cline{5-6} \smallspace & $I$ & $F_a$ & $F_b$ & \multicolumn{1}{r}{RME} & \multicolumn{1}{r}{$z$-component} \\ \hline \smallspace $^{85}$Rb & 2.5 & 2 & 1 & 0.224 & 0.0708 \\ & & 2 & 2 & 0.409 & 0.149 \\ & & 2 & 3 & 0.448 & -0.0977 \\ & & 3 & 2 & 0.219 & 0.0477 \\ & & 3 & 3 & 0.501 & 0.164 \\ & & 3 & 4 & 0.733 & -0.122 \\ \smallspace $^{87}$Rb & 1.5 & 1 & 1 & 0.273 & 0.112 \\ & & 1 & 2 & 0.417 & -0.132 \\ & & 2 & 1 & 0.122 & 0.0386 \\ & & 2 & 2 & 0.417 & 0.152 \\ & & 2 & 3 & 0.746 & -0.163 \\ \smallspace $^{133}$Cs & 3.5 & 3 & 2 & 3.40 & 0.743 \\ & & 3 & 3 & 5.03 & 1.65 \\ & & 3 & 4 & 4.89 & -0.815 \\ & & 4 & 3 & 2.91 & 0.484 \\ & & 4 & 4 & 5.78 & 1.72 \\ & & 4 & 5 & 7.71 & -1.04 \\ \smallspace $^{135}$Ba$^+$ & 1.5 & 1 & 1 & -0.311 & -0.127 \\ & & 1 & 2 & -0.475 & 0.150 \\ & & 2 & 1 & -0.139 & -0.0440 \\ & & 2 & 2 & -0.475 & -0.174 \\ & & 2 & 3 & -0.850 & 0.186 \\ \smallspace $^{171}$Yb$^+$ & 0.5 & 1 & 2 & -11.3 & 3.57 \\ \smallspace $^{173}$Yb$^+$ & 2.5 & 2 & 1 & -2.67 & -0.845 \\ & & 2 & 2 & -4.88 & -1.78 \\ & & 2 & 3 & -5.34 & 1.17 \\ & & 3 & 2 & -2.61 & -0.569 \\ & & 3 & 3 & -5.98 & -1.96 \\ & & 3 & 4 & -8.75 & 1.46 \\ \end{tabular \end{ruledtabular \label{tab:sd5on2-Cs \end{table \begin{table \centerin \caption{ SD-PNC amplitudes of the $\ket{7sF_a}\to\ket{6d_{5/2}F_b}$ transitions in Fr, Ra$^+$ and Ac$^{2+}$. Units: $10^{-13}ea_0\varkappa$.} \begin{ruledtabular \begin{tabular}{lllldd} & & & & \multicolumn{2}{c}{$E_{{\rm PNC}}$} \\ \cline{5-6} \smallspace & $I$ & $F_a$ & $F_b$ & \multicolumn{1}{r}{RME} & \multicolumn{1}{r}{$z$-component} \\ \hline \smallspace $^{211}$Fr & 4.5 & 4 & 3 & 24.3 & 4.05 \\ & & 4 & 4 & 32.6 & 9.72 \\ & & 4 & 5 & 29.6 & -3.99 \\ & & 5 & 4 & 19.7 & 2.65 \\ & & 5 & 5 & 36.3 & 10.0 \\ & & 5 & 6 & 45.8 & -5.18 \\ \smallspace $^{221}$Fr & 2.5 & 2 & 1 & 13.2 & 4.17 \\ & & 2 & 2 & 24.1 & 8.79 \\ & & 2 & 3 & 26.4 & -5.76 \\ & & 3 & 2 & 12.9 & 2.81 \\ & & 3 & 3 & 29.5 & 9.65 \\ & & 3 & 4 & 43.2 & -7.20 \\ \smallspace $^{223}$Fr & 1.5 & 1 & 1 & 16.1 & 6.57 \\ & & 1 & 2 & 24.6 & -7.77 \\ & & 2 & 1 & 7.20 & 2.28 \\ & & 2 & 2 & 24.6 & 8.97 \\ & & 2 & 3 & 44.0 & -9.59 \\ \smallspace $^{223}$Ra$^+$ & 1.5 & 1 & 1 & 4.16 & 1.70 \\ & & 1 & 2 & 6.35 & -2.01 \\ & & 2 & 1 & 1.86 & 0.588 \\ & & 2 & 2 & 6.35 & 2.32 \\ & & 2 & 3 & 11.4 & -2.48 \\ \smallspace $^{225}$Ra$^+$ & 0.5 & 1 & 2 & 14.4 & -4.55 \\ \smallspace $^{229}$Ra$^+$ & 2.5 & 2 & 1 & 3.41 & 1.08 \\ & & 2 & 2 & 6.22 & 2.27 \\ & & 2 & 3 & 6.82 & -1.49 \\ & & 3 & 2 & 3.33 & 0.726 \\ & & 3 & 3 & 7.62 & 2.49 \\ & & 3 & 4 & 11.2 & -1.86 \\ \smallspace $^{227}$Ac$^{2+}$ & 1.5 & 1 & 1 & 4.59 & 1.88 \\ & & 1 & 2 & 7.02 & -2.22 \\ & & 2 & 1 & 2.05 & 0.650 \\ & & 2 & 2 & 7.02 & 2.56 \\ & & 2 & 3 & 12.6 & -2.74 \\ \end{tabular \end{ruledtabular \label{tab:sd5on2-Fr \end{table Our calculations of the $s$-$d_{5/2}$ SD-PNC amplitudes of several isotopes of Rb, Cs, Ba$^+$ and Yb$^+$ are presented in Table~\ref{tab:sd5on2-Cs}, and for Fr, Ra$^+$ and Ac$^{2+}$ in Table~\ref{tab:sd5on2-Fr}. For ease of comparison we present both the reduced matrix elements, defined in Eq.~(\ref{eq:sd-pnc-rme}), and the $z$-components. The $s$-$d_{5/2}$ are typically between one and two orders of magnitude smaller than the corresponding $s$-$d_{3/2}$ transitions, due primarily to the absence of $s$-$p_{1/2}$ weak mixing. The largest amplitudes presented are in Fr, consistent with its very large $s$-$s$ and $s$-$d_{3/2}$ transitions. The amplitudes are large in fact for all the Fr-like ions, and are also large in Cs and Yb$^+$. As well as the $s$-$d_{5/2}$ transitions, which have no SI contribution, we have also performed calculations for several $s$-$d_{3/2}$ transitions for which both SI and SD contributions are non-zero. We express these amplitudes in the form $E_{\rm PNC} = P(1+R),$ where $P$ is the SI PNC amplitude (including $Q_W$), and $R$ is the ratio of the SD to SI parts. Here we calculate both parts concurrently, using the same method and wavefunctions. This approach has the advantage that the relative sign difference between the SI and SD parts is fixed, ensuring no ambiguity in the sign of $\varkappa$~\cite{Dzuba2011Yb}. There is also typically a significant improvement in accuracy for the ratio over that for each of the amplitudes individually, due to the fact that the atomic calculations for both components are very similar and much of the theoretical uncertainty cancels in the ratio~\cite{Dzuba2011}. We present these amplitudes for Rb and Cs in Table~\ref{tab:sd3on2-Cs}, and for Fr and Ac$^{2+}$ in Table~\ref{tab:sd3on2-Fr}. We don't present amplitudes for Ba$^{+}$, Yb$^{+}$ or Ra$^{+}$ since these have been performed in our recent work Ref.~\cite{Dzuba2011}. \begin{table \centerin \caption{ PNC amplitudes ($z$~components) of the $\ket{5sF_a}\to\ket{4d_{3/2}F_b}$ transition in Rb, and the $\ket{6sF_a}\to\ket{5d_{3/2}F_b}$ transitions in Cs. Units: $10^{-11}ea_0$. } \begin{ruledtabular \begin{tabular}{ldlllr} & Q_W & $I$ & $F_a$ & $F_b$ & \multicolumn{1}{c}{$E_{\rm PNC}$} \\ \hline \smallspace $^{87}$Rb & -46.8 & 1.5 & 1 & 0 & $ -0.301 \times[1+0.0805\varkappa]$ \\ & & & 1 & 1 & $ -0.337 \times[1+0.0796\varkappa]$ \\ & & & 1 & 2 & $ 0.261 \times[1+0.0779\varkappa]$ \\ & & & 2 & 1 & $ -0.117 \times[1-0.0439\varkappa]$ \\ & & & 2 & 2 & $ -0.301 \times[1-0.0457\varkappa]$ \\ & & & 2 & 3 & $ 0.301 \times[1-0.0483\varkappa]$ \\ \smallspace $^{133}$Cs & -73.2 & 3.5 & 3 & 2 & $-2.05 \times[1+0.0444\varkappa] $\\ & & & 3 & 3 & $-3.14 \times[1+0.0431\varkappa]$ \\ & & & 3 & 4 & $ 1.35 \times[1+0.0412\varkappa] $\\ & & & 4 & 3 & $ -0.923 \times[1-0.0305\varkappa]$ \\ & & & 4 & 4 & $ -2.86 \times[1-0.0323\varkappa] $\\ & & & 4 & 5 & $ 1.87 \times[1-0.0345\varkappa] $\\ \end{tabular \end{ruledtabular \label{tab:sd3on2-Cs \end{table \begin{table \centerin \caption{ PNC amplitudes of the $\ket{7sF_a}\to\ket{6d_{3/2}F_b}$ transitions in Fr and Ac$^{2+}$. Units: $10^{-11}ea_0$. } \begin{ruledtabular \begin{tabular}{ldlllr} & Q_W & $I$ & $F_a$ & $F_b$ & \multicolumn{1}{c}{$E_{\rm PNC}$} \\ \hline \smallspace $^{223}$Fr & -128.3 & 1.5 & 1 & 0 & $ -38.4 \times[1+0.0273\varkappa]$ \\ & & & 1 & 1 &$ -43.0 \times[1+0.0278\varkappa] $\\ & & & 1 & 2 &$ 33.3 \times[1+0.0288\varkappa]$ \\ & & & 2 & 1 &$ -14.9 \times[1-0.0189\varkappa]$ \\ & & & 2 & 2 &$ -38.4 \times[1-0.0179\varkappa]$ \\ & & & 2 & 3 &$ 38.4 \times[1-0.0164\varkappa] $\\ \smallspace $^{227}$Ac$^{2+}$ & -130.1 & 1.5 & 1 & 0 & $ -28.7 \times[1+0.0250\varkappa] $\\ & & & 1 & 1 &$ -32.0 \times[1+0.0241\varkappa]$ \\ & & & 1 & 2 &$ 24.8 \times[1+0.0223\varkappa]$ \\ & & & 2 & 1 &$ -11.1 \times[1-0.0105\varkappa]$ \\ & & & 2 & 2 &$ -28.7 \times[1-0.0123\varkappa]$ \\ & & & 2 & 3 &$ 28.7 \times[1-0.0150\varkappa] $\\ \end{tabular \end{ruledtabular \label{tab:sd3on2-Fr \end{table \subsection{Suitability for measurements} A method has been proposed by Fortson for measuring PNC in a single atomic ion that has been laser trapped and cooled~\cite{Fortson}. Originally proposed with measuring the $6s$-$5d_{3/2}$ transition of Ba$^+$ in mind, work has begun to use this method for the $7s$-$6d_{3/2}$ transition in Ra$^+$ at KVI~\cite{KVI}. The use of this or a similar method to study spin-dependent PNC in $s$-$d_{5/2}$ transitions has been previously discussed~\cite{Geetha1998,Sahoo2011,propose}. Though these transitions have significantly smaller PNC signals than the corresponding $s$-$d_{3/2}$ transitions, the main advantage here is that there is no SI contribution. This is beneficial for the extraction of the nuclear anapole moment since the (larger) SI contribution would not need to be subtracted, and it would limit the possibility of spurious SI-PNC acting as a false signal. In~\cite{Fortson} it was shown that to ensure accurate PNC measurements of a single trapped ion both the upper and lower levels of the transition should be long lived. The only significant contribution to the decay rate of the $5,6d_{5/2}$ states in Ba$^{+}$, Ra$^{+}$ are the E2 transitions to the $s$ ground state. There are also M1 and E2 $d_{5/2}$-$d_{3/2}$ contributions, though these are highly suppressed. Both E2 transitions are suppressed in the case of Ac$^{2+}$, so we include both in the calculation. We calculate the lifetimes of the relevant $d_{5/2}$ states in Ba$^{+}$, Ra$^{+}$ and Ac$^{2+}$ to be 35.9 s, 0.302 s, and 247 s respectively. These results are in good agreement with other recent calculations, e.g.~\cite{Pal2009,Sahoo2006}. The upper states of the other elements presented here are unstable as they have allowed E1 transitions to lower levels. This is not a problem for neutral Cs or Fr where atomic-beam-type experiments could be used. In the $s$-$d_{5/2}$ transitions considered here it is possible that the contribution to the amplitude coming from the combination of the weak charge and hyperfine interaction may not be as small as in other systems, due to the $d_{3/2}$-$d_{5/2}$ and $p_{1/2}$-$p_{3/2}$ hyperfine mixing. The ratio of the hyperfine to fine structure splitting goes as \begin{equation} \frac{1}{Z}\frac{m_e}{m_p}\sim10^{-5}. \end{equation} The PNC amplitude of the $s$-$d_{5/2}$ transitions due to the combined weak charge and hyperfine interaction would therefore be of the order \begin{equation} E_{\rm PNC}^{Q_W+{\rm hf}}(s{\rm-}d_{5/2}) \sim 10^{-5}\, E_{\rm PNC}^{Q_W}(s{\rm-}d_{3/2}). \end{equation} For Cs, this leads to a $Q_W+{\rm hf}$ contribution on the order of $10^{-16}$ (including $Q_W$), whereas the anapole moment contribution to this transition is $10^{-13}\varkappa\sim 10^{-14}$. Similarly for Ba$^+$, Fr and Ra$^+$ the $Q_W+{\rm hf}$ contribution is between one and two orders of magnitude smaller than the contribution from the anapole moment. This is smaller than the assumed accuracy here, so this contribution can be safely neglected for now. An accurate calculation of this contribution is beyond the scope of the current work, and will be the focus of a future study. \section{Conclusion} We have presented order-of-magnitude calculations of nuclear-spin-dependent PNC amplitudes for the $s$-$d_{5/2}$ transitions of several heavy atoms and ions. Also presented are PNC amplitudes of the $s$-$d_{3/2}$ transitions of the same ions (where not presented previously) that are accurate to the $\sim10\%$ level. These calculations could be used to extract an experimental value of the nuclear anapole moment, which in turn could be used to study parity violating nuclear forces. The accuracy of these calculations could be improved with the inclusion of higher order correlation corrections, such as the double-core-polarization~\cite{DCP}, structure radiation~\cite{CPM2} and ladder-diagrams~\cite{ladder}, as well as other small corrections such as the Breit~\cite{Breit} and QED~\cite{QED} corrections. \acknowledgments One of the authors (V.A.D.) would like to express a special thanks to the Mainz Institute for Theoretical Physics (MITP) for its hospitality and support. We extend our thanks to D. S. Elliot for stimulating this work. The work was also supported by the Australian Research Council.
\section{\protect\smallskip Introduction} Atomic coherence lies in the center of many novel effects in quantum optics and laser physics. Electromagnetically induced transparency [1,2], coherent population trapping [2], Hanel-effect laser [3] and quantum beat laser [4] are such examples. Besides these, the correlation between the photons can also be induced by atomic coherence [5-12]. One such example is the generation of squeezed light in a three-level cascade laser using atomic coherence [5-8]. The atomic coherence can be created by preparing the atoms initially in a coherent superposition state of the two states which are dipole-forbidden [5-8] or driving the two states by a strong coherent field [9-11] or Raman coupling the two states through the third auxiliary atomic states [12]. For two-photon correlated-spontaneous-emission laser with injected atomic coherence, it exhibits complete spontaneous-emission noise quenching and phase squeezing simultaneously [5]. It has also been pointed out that atomic coherence in a two-photon correlated emission laser system can be used to generate a macroscopic two-mode entangled state and this system can be treated as an entanglement amplifier [12]. Recently, the topic of continuous-variable entanglement has attracted much attention as it is the base of all branches of quantum information and communication protocols [13]. Among various entanglement generation schemes, entanglement induced by atomic coherence has been extensively researched [14-16]. For a nondegenerate three-level cascade laser with a subthreshold nondegenerate parametric oscillator coupled to a vacuum reservoir, the entanglement and squeezing for the two cavity modes in this combined system is induced by the injected atomic coherence [14]. In a two-mode single-atom laser with the atomic coherence exhibited by two classical laser fields, entanglement between two field modes is demonstrated [15]. Later, it was shown that in a three-level $\Lambda $ or V atomic system with two classical driving fields and two cavity modes coupling corresponding transitions, by exploring the two-channel interaction mechanism and using the squeeze-transformed modes, continuous-variable entanglement between the two modes is obtained and the best achievable entangled state approaches the original EPR state [16]. The above work has mainly been confined to two-partite systems. With the progress in continuous-variable entanglement, the generation of more than two partite entanglement has been paid much attention as it may be the key ingredient for advanced multiparty quantum communication such as quantum teleportation network [17], telecloning [18] and controlled dense coding [19]. Among various generation schemes for tripartite systems, few work has been done to generate tripartite entanglement using atomic coherence. Most recently, a scheme to generate three-mode-entangled light fields via the interaction between the four-level atoms and the cavity has been proposed [20]. Three cavity modes are generated through three successive transitions in the four-level cascade atoms. In addition to the cavity modes, two strong classical fields drive a pair of two-photon transitions in the four-level atoms. They show that the entanglement could only be obtained in a short time as all the mean photons are amplified as time elapses. Thus at steady time, the entanglement does not exist. In this paper, we present a scheme to generate tripartite entanglement for three cavity modes via the interaction for the three-level lambda atoms with the three cavity modes and two classical fields. As the classical fields are strong, the effective interaction is resonant interaction in the dressed-state picture. We deduce the master equation of the three cavity modes by means of the atomic dressed states and linear theory. The sufficient inseparability criterion for continuous-variable entanglement is used to demonstrate the entanglement properties of the three cavity modes and the results show that our system can be used as a source to generate tripartite entangled light even at steady state. It is should be noted that, up to now many schemes have been proposed to generate tripartite entanglement using linear optics or nonlinearities [17, 21-26]. It was theoretically predicted that using single-mode squeezed state and linear optics, a truly N-partite entangled state can be generated [17]. Later, a continuous-variable tripartite entangled state was experimental realized by combing three independent squeezed vacuum states [21]. At first, the production of continuous-variable tripartite entanglement was presented by mixing squeezed beams on unbalanced beamsplitters [21,22]. Recently, generation of tripartite entanglement are focused on using cascade nonlinear interaction in an optical cavity [23-25] or in a quasiperiodic superlattice [26]. Among the latter are systems using parametric down-conversion with sum frequency generation [23,25,26] or using single nonlinearity [24]. During these nonlinear processes, the cavity modes couple with each other directly. As these nonlinear processes are related to the higher-order polarization, the efficiency of these processes are relatively small compared with the processes related to linear polarization. In this way, these nonlinear processes are not the best choice for the generation of high efficiency tripartite entangled states. Compared with the schemes based on the nonlinear processes [23-26], our scheme is more effective as the generation process is resonant interaction in the dressed states and it is only related to linear polarization. What's more, the linear process provides much more parameters to choose than that of the nonlinear processes, as the atomic parameters can be varied. Compared with the scheme in Ref. [20], our scheme can provide steady state tripartite entanglement while the entanglement produced in scheme [20] is just kept in a quite limited time. And in Ref. [20], they use four-level cascade atomic system and the effective processes in the dressed states of the driving fields are all two-photon transitions. High excited states are involved in their scheme. In our scheme we use three-level $\Lambda $ atomic system, and the effective processes in the dressed states of the driving fields are all single-photon transitions. When we take into the account of the atomic spontaneous emission, their schemes seems to have more obstacles than ours. The paper is organized as follows. In Sec. II, we discuss the essential ingredients of the model and deduce the density-matrix equation for the cavity fields in a dressed-state picture. In Sec. III, we present the output correlation spectra by solving the equations of the cavity fields and analyze the output tripartite continuous-variable entanglement characteristics by using a sufficient criterion proposed by van Loock and Furusawa. In Sec. IV, we give a brief conclusion. \section{Model and equation} We consider $N$ three-level lambda-type atoms in a three-mode cavity as shown in Fig. 1(a). Two laser fields of frequencies $\omega _{l1,l2}$ drive the transitions $|1,2\rangle \leftrightarrow $ $|3\rangle $, respectively. Two cavity modes $a_{1,2}$ of frequencies $\omega _{c1,c2}$ couple the atomic transition $|1\rangle \leftrightarrow $ $|3\rangle $, while the cavity mode $a_{3}$ with frequency $\omega _{c3}$ couples the transition |2\rangle \leftrightarrow $ $|3\rangle $. $\gamma _{l}$ ($l=1,2$) are the atomic decay rates from level $|3\rangle $ to levels $|1,2\rangle $ and \kappa _{l}$ ($l=1,2,3$) are the cavity loss rates. For simplicity, we assume that $\gamma _{1}=\gamma _{2}=\gamma $ and $\kappa _{1}=\kappa _{2}=\kappa _{3}=\kappa $. The three cavity modes are assumed to be in their vacuum state initially. In the frame of the frequencies of the laser fields and under the dipole and the rotating-wave approximations, the total Hamiltonian is \begin{figure}[tbph] \centering \includegraphics[width=13 cm]{fig1.eps} \caption{(a) Atomic energy level scheme and the coupling of the cavity fields and the classic fields. (b) Equivalent resonant transitions in the picture dressed by the classical fields.} \label{fig1} \end{figure} \begin{eqnarray} H &=&H_{1}+H_{2}+H_{3}, \nonumber \\ H_{1} &=&\sum_{j=1}^{3}\hbar \delta _{j}a_{j}^{\dagger }a_{j}, \nonumber \\ H_{2} &=&-\hbar \Delta \left( \sigma _{11}-\sigma _{22}\right) +\hbar \Omega \left( \sigma _{31}+\sigma _{32}+H.c.\right) , \label{1} \\ H_{3} &=&i\hbar g\left( a_{1}\sigma _{31}+a_{2}\sigma _{31}+a_{3}\sigma _{32}\right) +H.c., \nonumber \end{eqnarray} H.c. symbols the Hermitian conjugate. $H_{1}$ denotes the free energy for three cavity fields, $H_{2}$ describes the interaction of the laser fields with the atoms, and $H_{3}$ indicates the interaction of the cavity fields with the atoms. $\sigma _{jk}=|j\rangle \langle k|$ ($j,k=1,2,3$) are atomic dipole operators for $j\neq k$ and atomic projection operators for $j=k$. The cavity detunings are defined as $\delta _{j}=\omega _{cj}-\omega _{l1}$ $j=1,2$), and $\delta _{3}=\omega _{c3}-\omega _{l2}$. The detunings of the laser fields are defined as $\Delta _{j}=\omega _{3j}-\omega _{lj}$ ($j=1,2 ), where $\omega _{31}$ and $\omega _{32}$ are the resonance frequencies of transitions $|1,2\rangle \leftrightarrow $ $|3\rangle $. We have assumed equal coupling coefficients $g$ for three cavity modes, equal Rabi frequency $\Omega $ for the two laser fields, and opposite detunings of the two laser fields $\Delta _{1}=-\Delta _{2}=\Delta $. We assume that the laser fields are much stronger than the cavity fields, i.e., $\Omega \gg g\langle a_{l}\rangle $, ($l=1,2,3)$. The laser fields can be viewed as dressing fields for the atoms. Therefore, by diagonalizing the Hamiltonian $H_{2}$, we find the so-called semiclassical dressed states as \begin{eqnarray} |0\rangle &=&-\frac{c}{\sqrt{2}}|1\rangle +\frac{c}{\sqrt{2}}|2\rangle +s|3\rangle , \nonumber \\ |+\rangle &=&\frac{1+s}{2}|1\rangle +\frac{1-s}{2}|2\rangle +\frac{c}{\sqrt{ }}|3\rangle , \label{2} \\ |-\rangle &=&\frac{1-s}{2}|1\rangle +\frac{1+s}{2}|2\rangle -\frac{c}{\sqrt{ }}|3\rangle , \nonumber \end{eqnarray} where $c=\frac{\sqrt{2}\Omega }{d}$, $s=-\frac{\Delta }{d}$, and $d=\sqrt \Delta ^{2}+2\Omega ^{2}}$. Now, we use the Hamiltonian $H_{0}=\hbar d\left( \sigma _{++}-\sigma _{--}\right) +H_{1}$ to perform the unitary dressing transformation. By choosing the cavity detunings as $\delta _{1}=d=-\delta _{2}=-\delta _{3}$, and neglecting the fast-oscillating terms such as $e^{\pm i2dt}$, we obtain the resonant interaction Hamiltonian as \begin{equation} V=ig\hbar \left( c_{3}a_{1}^{\dagger }+c_{2}a_{2}+c_{1}a_{3}\right) \sigma _{0+}+ig\hbar \left( c_{1}a_{1}-c_{3}a_{2}^{\dagger }+c_{3}a_{3}^{\dagger }\right) \sigma _{0-}+H.c., \label{3} \end{equation} where $c_{1}=\frac{1}{2}s(1-s)$, $c_{2}=\frac{1}{2}s(1+s)$, and $c_{3}=\frac 1}{2}c^{2}$. The resonant transitions in the dressed states are shown in Fig. 1(b). The master equation for the cavity modes is obtained by using the usual approach [2], starting from $\frac{d}{dt}\rho =-\frac{i}{\hbar }\left[ V,\rho \right] {\cal +L}_{a}\rho +{\cal L}_{c}\rho $, where ${\cal L _{c}\rho =\frac{_{\kappa }}{2}\sum_{l=1}^{3}\left( 2a_{l}\rho a_{l}^{\dagger }-a_{l}^{\dagger }a_{l}\rho -\rho a_{l}^{\dagger }a_{l}\right) $ and ${\cal }_{a}\rho $ describes the atomic decay in the dressed states picture and its expression is very complicated. The detailed form of atomic decay term {\cal L}_{a}\rho $ is given in Appendix A. The master equation for the cavity modes is obtained by tracing out the atomic states, which gives \frac{d}{dt}\rho _{c}=g\left( c_{3}a_{1}^{\dagger }+c_{2}a_{2}+c_{1}a_{3}\right) \rho _{+0}+g\left( c_{1}a_{1}-c_{3}a_{2}^{\dagger }+c_{3}a_{3}^{\dagger }\right) \rho _{-0}+H.c , where $\rho _{jk}=tr_{atom}(\sigma _{kj}\rho )$ ($j,k=0,+,-$). As the atomic variables vary much faster than the cavity fields, it is possible to express $\rho _{jk}=tr_{atom}(\sigma _{kj}\rho )$ ($jk=+0,-0,0+,0-$) in terms of $\rho _{c}$, $a_{l}$ and $a_{l}^{\dagger }$ ($l=$1-3) from the quasi-steady-state solution of the coupled equations for $\rho _{jk}=tr_{atom}(\sigma _{kj}\rho )$ ($jk=+0,-0,0+,0-$). By using $\rho _{jj}\simeq \rho _{jj}^{s}\rho _{c}$ ($j=0,+,-$) and $\rho _{+-}^{s}\simeq 0 , where ``s'' implies the steady-state solutions of the density matrix equations in the dressed state picture without the quantum fields $a_{l}$ and $a_{l}^{\dagger }$ ($l=$1-3). The steady state populations is obtained as $\rho _{00}^{s}=\frac{c^{4}}{1+3s^{4}}$ and $\rho _{++}^{s}=\rho _{--}^{s}=\frac{s^{2}\left( 1+s^{2}\right) }{1+3s^{4}}$. The master equation for the cavity modes is obtained as \begin{eqnarray} \frac{d}{dt}\rho _{c} &=&\sum_{l=1}^{3}\left\{ A_{ll}\left[ a_{l}^{\dagger },\rho _{c}a_{l}\right] -\left( B_{ll}+\frac{\kappa _{l}}{2}\right) \left[ a_{l}^{\dagger },a_{l}\rho _{c}\right] \right\} \nonumber \\ &&+\sum_{l=2}^{3}\left\{ A_{1l}\left[ a_{1}^{\dagger },\rho _{c}a_{l}^{\dagger }\right] -B_{1l}\left[ a_{1}^{\dagger },a_{l}^{\dagger }\rho _{c}\right] +A_{l1}\left[ a_{l}^{\dagger },\rho _{c}a_{l}^{\dagger }\right] -B_{l1}\left[ a_{l}^{\dagger },a_{1}^{\dagger }\rho _{c}\right] \right\} \label{4} \\ &&+\sum_{l,k=2;l\neq k}^{3}\left\{ A_{lk}\left[ a_{l}^{\dagger },\rho _{c}a_{k}\right] -B_{lk}\left[ a_{l}^{\dagger },a_{k}\rho _{c}\right] \right\} +H.c.. \nonumber \end{eqnarray} The explicit expressions for $A_{lk}$ and $B_{lk}$ ($l,k=1,2,3$) are given in Appendix B. Here the terms $A_{ll}$ ($l$=1-3) and $B_{ll}$ ($l$=1-3) represent the gain term and the absorption of mode $a_{l}$, respectively. And the terms $A_{lk}$ and $B_{lk}$ ($l\neq k$) represent the coupling between the two modes $a_{l}$ and $a_{k}$, and we will show that these quantities are responsible for entanglement among three cavity fields. It is easy to see that without these coupling terms between different cavity fields, the quantum correlation can not be introduced among the three cavity modes. Thus entanglement among the three cavity fields is attributed to the atomic coherence created through the interaction between the fields and the atoms. \section{Correlation spectra} The master equation (4) enables us to derive equations of motion for the cavity modes: \begin{eqnarray} \tau \frac{d}{dt}a_{1}^{\dagger } &=&\left( A_{11}-B_{11}-\frac{\kappa _{1}} 2}\right) a_{1}^{\dagger }+\left( A_{12}-B_{12}\right) a_{2}+\left( A_{13}-B_{13}\right) a_{3}+\sqrt{\kappa _{1}}a_{1}^{\dagger in}, \nonumber \\ \tau \frac{d}{dt}a_{2} &=&\left( A_{21}-B_{21}\right) a_{1}^{\dagger }+\left( A_{22}-B_{22}-\frac{\kappa _{2}}{2}\right) a_{2}+\left( A_{23}-B_{23}\right) a_{3}+\sqrt{\kappa _{2}}a_{2}^{in}, \label{5} \\ \tau \frac{d}{dt}a_{3} &=&\left( A_{31}-B_{31}\right) a_{1}^{\dagger }+\left( A_{32}-B_{32}\right) a_{2}+\left( A_{33}-B_{33}-\frac{\kappa _{3}}{ }\right) a_{3}+\sqrt{\kappa _{3}}a_{3}^{in}, \nonumber \end{eqnarray} where $\tau $ is the round-trip time of light in the cavity and assumed to be the same for three cavity modes. $a_{j}^{in}$ and $a_{j}^{\dagger in}$ ( j=$1-3) are annihilation and creation operators of the input fields to the cavity. This is a set of linear equations. In order to solve this equation, we use the Fourier transformation and the boundary conditions at the mirror between the output quantities and the input quantities a_{j}^{in}+a_{j}^{out}=\sqrt{\kappa _{j}}a_{j}$ ($j=$1-3) to obtain the equation in the frequency domain as \begin{equation} a^{out}\left( \omega \right) =-\left( I+BD_{0}^{-1}B\right) a^{in}\left( \omega \right) , \label{6} \end{equation} where \begin{equation} \begin{tabular}{l} $a^{out}\left( \omega \right) =\left( a_{1}^{\dagger out}\left( -\omega \right) ,a_{2}^{out}\left( \omega \right) ,a_{3}^{out}\left( \omega \right) \right) ^{T},$ \\ $a^{in}\left( \omega \right) =\left( a_{1}^{\dagger in}\left( -\omega \right) ,a_{2}^{in}\left( \omega \right) ,a_{3}^{in}\left( \omega \right) \right) ^{T}, \end{tabular} \label{7} \end{equation} \begin{equation} D_{0}=\left( \begin{array}{ccc} A_{11}-B_{11}-\frac{\kappa _{1}}{2}-i\omega \tau & A_{12}-B_{12} & A_{13}-B_{13} \\ A_{21}-B_{21} & A_{22}-B_{22}-\frac{\kappa _{2}}{2}-i\omega \tau & A_{23}-B_{23} \\ A_{31}-B_{31} & A_{32}-B_{32} & A_{33}-B_{33}-\frac{\kappa _{3}}{2}-i\omega \tau \end{array} \right) , \label{8} \end{equation} \begin{equation} B=\left( \begin{array}{lll} \sqrt{\kappa _{1}} & 0 & 0 \\ 0 & \sqrt{\kappa _{2}} & 0 \\ 0 & 0 & \sqrt{\kappa _{3}} \end{array} \right) ,\text{\qquad }I=\left( \begin{array}{lll} 1 & 0 & 0 \\ 0 & 1 & 0 \\ 0 & 0 & 1 \end{array} \right) . \label{9} \end{equation} where T symbols the matrix transpose. \begin{figure}[tbph] \centering \includegraphics[width=15 cm]{fig2.eps} \caption{The quantum correlations spectra $S_{123}^{out}\left( \omega ^{\prime }\right) $, $S_{231}^{out}\left( \omega ^{\prime }\right) $ and S_{312}^{out}\left( \omega ^{\prime }\right) $ versus the normalized analyzing frequency $\omega ^{\prime }$ are plotted for (a) $\Delta =5$ and (b) $\Delta =10$ by solid, dashed and dotted line, respectively. The other parameters are $\Omega =35$, $g^{2}N=10$, $\gamma =1$ and $\kappa =0.1$.} \label{fig2} \end{figure} \begin{figure}[tbph] \centering \includegraphics[width=15 cm]{fig3.eps} \caption{The quantum correlations spectra $S_{123}^{out}\left( \omega ^{\prime }\right) $, $S_{231}^{out}\left( \omega ^{\prime }\right) $ and S_{312}^{out}\left( \omega ^{\prime }\right) $ versus the detuning $\Delta $ for (a) $\omega ^{\prime }=0$ and (b) $\omega ^{\prime }=1.0$ by solid, dashed and dotted line, respectively. The other parameters are the same as those in Fig. 2.} \label{fig3} \end{figure} In order to study the entanglement properties of output cavity modes, we need to use quadrature amplitude and phase operators defined by \begin{eqnarray} X_{j}^{out} &=&a_{j}^{out}\left( \omega \right) +a_{j}^{\dagger out}\left( -\omega \right) , \nonumber \\ Y_{j}^{out} &=&-i\left[ a_{j}^{out}\left( \omega \right) -a_{j}^{\dagger out}\left( -\omega \right) \right] , \label{10} \end{eqnarray} Using Eq. (6) and $X_{j}^{in}=a_{j}^{in}\left( \omega \right) +a_{j}^{\dagger in}\left( -\omega \right) $, $Y_{j}^{in}=-i\left[ a_{j}^{in}\left( \omega \right) +a_{j}^{\dagger in}\left( -\omega \right) \right] $, we can obtain the relationships between the input fields and the output fields as \smallskip \begin{eqnarray} && \begin{tabular}{l} $X_{1}^{out}\left( \omega ^{\prime }\right) =D_{11}X_{1}^{in}\left( \omega ^{\prime }\right) +D_{12}X_{2}^{in}\left( \omega ^{\prime }\right) +D_{13}X_{3}^{in}\left( \omega ^{\prime }\right) ,$ \\ $Y_{1}^{out}\left( \omega ^{\prime }\right) =D_{11}Y_{1}^{in}\left( \omega ^{\prime }\right) -D_{12}Y_{2}^{in}\left( \omega ^{\prime }\right) -D_{13}Y_{3}^{in}\left( \omega ^{\prime }\right) , \end{tabular} \nonumber \\ && \begin{tabular}{l} $X_{2}^{out}\left( \omega ^{\prime }\right) =D_{21}X_{1}^{in}\left( \omega ^{\prime }\right) +D_{22}X_{2}^{in}\left( \omega ^{\prime }\right) +D_{23}X_{3}^{in}\left( \omega ^{\prime }\right) ,$ \\ $Y_{2}^{out}\left( \omega ^{\prime }\right) =-D_{21}Y_{1}^{in}\left( \omega ^{\prime }\right) +D_{22}Y_{2}^{in}\left( \omega ^{\prime }\right) +D_{23}Y_{3}^{in}\left( \omega ^{\prime }\right) , \end{tabular} \label{11} \\ && \begin{tabular}{l} $X_{3}^{out}\left( \omega ^{\prime }\right) =D_{31}X_{1}^{in}\left( \omega ^{\prime }\right) +D_{32}X_{2}^{in}\left( \omega ^{\prime }\right) +D_{33}X_{3}^{in}\left( \omega ^{\prime }\right) ,$ \\ $Y_{3}^{out}\left( \omega ^{\prime }\right) =-D_{31}Y_{1}^{in}\left( \omega ^{\prime }\right) +D_{32}Y_{2}^{in}\left( \omega ^{\prime }\right) +D_{33}Y_{3}^{in}\left( \omega ^{\prime }\right) , \end{tabular} \nonumber \end{eqnarray} where we have defined the normalized analyzing frequency $\omega ^{\prime }=\omega \tau /\kappa $. The explicit expressions for $D_{jk}$ ($j,k$=1-3) are presented in Appendix C. The presence of entanglement between the three cavity modes can be investigated using the sufficient criterion for continuous-variable tripartite system proposed by van Loock and Furusawa [27]. The sufficient inseparability criterion for continuous variable tripartite entanglement is that if any one of the following inequalities is satisfied, genuine tripartite entanglement is demonstrated. The inequalities are \begin{eqnarray} S_{123} &=&V\left[ X_{1}+\left( X_{2}+X_{3}\right) /\sqrt{2}\right] +V\left[ Y_{1}-\left( Y_{2}+Y_{3}\right) /\sqrt{2}\right] <4, \nonumber \\ S_{231} &=&V\left[ X_{2}+\left( X_{3}+X_{1}\right) /\sqrt{2}\right] +V\left[ Y_{2}-\left( Y_{3}+Y_{1}\right) /\sqrt{2}\right] <4, \label{12} \\ S_{312} &=&V\left[ X_{3}+\left( X_{1}+X_{2}\right) /\sqrt{2}\right] +V\left[ Y_{3}-\left( Y_{1}+Y_{2}\right) /\sqrt{2}\right] <4, \nonumber \end{eqnarray} where $V\left( A\right) =<A^{2}>-<A>^{2}$. From the above definition, the correlation spectra of the quadratures of three output cavity fields are obtained as \begin{eqnarray} S_{123}^{out}\left( \omega ^{\prime }\right) &=&|\sqrt{2 D_{11}-D_{21}-D_{31}|^{2}+|\sqrt{2}D_{12}-D_{22}-D_{32}|^{2}+|\sqrt{2 D_{13}-D_{23}-D_{33}|^{2}, \nonumber \\ S_{231}^{out}\left( \omega ^{\prime }\right) &=&\frac{1}{2}\left( |\sqrt{2 D_{21}-D_{11}-D_{31}|^{2}+|\sqrt{2}D_{22}-D_{12}-D_{32}|^{2}+|\sqrt{2 D_{23}-D_{13}-D_{33}|^{2}\right) \nonumber \\ &&+\frac{1}{2}\left( |\sqrt{2}D_{21}-D_{11}+D_{31}|^{2}+|\sqrt{2 D_{22}-D_{12}+D_{32}|^{2}+|\sqrt{2}D_{23}-D_{13}+D_{33}|^{2}\right) , \label{13} \\ S_{312}^{out}\left( \omega ^{\prime }\right) &=&\frac{1}{2}\left( |\sqrt{2 D_{31}-D_{11}-D_{21}|^{2}+|\sqrt{2}D_{32}-D_{12}-D_{22}|^{2}+|\sqrt{2 D_{33}-D_{13}-D_{23}|^{2}\right) \nonumber \\ &&+\frac{1}{2}\left( |\sqrt{2}D_{31}+D_{11}-D_{21}|^{2}+|\sqrt{2 D_{32}-D_{12}+D_{22}|^{2}+|\sqrt{2}D_{33}-D_{13}+D_{23}|^{2}\right) . \nonumber \end{eqnarray} The quantum correlations spectra $S_{123}^{out}\left( \omega ^{\prime }\right) $, $S_{231}^{out}\left( \omega ^{\prime }\right) $ and S_{312}^{out}\left( \omega ^{\prime }\right) $ for three output cavity fields described in Eq. (13) versus the normalized analyzing frequency \omega ^{\prime }$ are plotted in Fig. $2$ for (a) $\Delta =5$ and (b) \Delta =10$ by solid, dashed and dotted line, respectively. The other parameters are $\Omega =35$, $g^{2}N=10$, $\gamma =1$ and $\kappa =0.1$. The satisfaction of one of the three inequalities $S_{123}^{out}\left( \omega ^{\prime }\right) <4$, $S_{231}^{out}\left( \omega ^{\prime }\right) <4$ and $S_{312}^{out}\left( \omega ^{\prime }\right) <4$ is sufficient to demonstrate genuine tripartite entanglement. In order to analyze the entanglement properties of the three cavity modes, we present all three correlations $S_{ijk}^{out}\left( \omega ^{\prime }\right) $ and find that the indices of the three cavity modes are crucial. When the cavity modes are symmetric, the indices of the cavity modes are not important as the three correlations give the same result. But when the cavity modes are asymmetric, the indices are crucial in that the three correlations will give different results. As shown in Fig. 2(a), all three correlations are below 4 in a wide frequency range thus all three inequalities are satisfied. So the three output cavity modes are entangled. Among the three correlations, S_{123}^{out}\left( \omega ^{\prime }\right) $ gives the minimum values with the same parameters. When the inequalities are satisfied, the smaller the values of correlations are the larger the correlation degree. When we increase the detuning $\Delta $ to $10$ and keep other parameters unchanged as shown in Fig. 2(b), correlations $S_{123}^{out}\left( \omega ^{\prime }\right) $ and $S_{231}^{out}\left( \omega ^{\prime }\right) $ are always below $4$ in a wide frequency range while the correlation S_{312}^{out}\left( \omega ^{\prime }\right) $ is larger than $4$ in a frequency zone around the central analyzing frequency $\omega ^{\prime }=0$. Thus tripartite entanglement is also demonstrated between the three output cavity modes. Compared with Fig. 2(a), the minimum value of S_{123}^{out}\left( \omega ^{\prime }\right) $ is smaller, which means that the correlation degree is also increased with the detuning. For both cases, we also see that the large correlation can be obtained at low analyzing frequency $\omega ^{\prime }$. In Fig. $3$, we plot $S_{123}^{out}\left( \omega ^{\prime }\right) $, S_{231}^{out}\left( \omega ^{\prime }\right) $ and $S_{312}^{out}\left( \omega ^{\prime }\right) $ as a function of detuning $\Delta $ for (a) \omega ^{\prime }=0$ and (b) $\omega ^{\prime }=1.0$ by solid, dashed and dotted line, respectively. The remain parameters are the same as those in Fig. 2. We also see that the correlation $S_{123}^{out}\left( \omega ^{\prime }\right) $ gives the minimum values with the same parameters. It is seen from Fig. 3(a) and 3(b) that, correlations $S_{123}^{out}\left( \omega ^{\prime }\right) $ and $S_{231}^{out}\left( \omega ^{\prime }\right) $ always satisfy the inequalities while $S_{312}^{out}\left( \omega ^{\prime }\right) $ only satisfy the inequality in a small frequency range. Thus tripartite entanglement between the three output cavity modes is demonstrated again. It is worthwhile to point out that when the analyzing frequency $\omega ^{\prime }=0$, the system reaches its steady state. Thus at steady state, we can also obtain entangled tripartite light. This is in contrast with the results in Ref. [20], where the entanglement between the three cavity modes is time dependent. In that case all the mean photon numbers are amplified as time increases. Thus, the entanglement for three cavity modes can not be kept for a long time. And among the three correlations, $S_{123}^{out}\left( \omega ^{\prime }\right) $ decreases with the increasing detuning, while correlations $S_{231}^{out}\left( \omega ^{\prime }\right) $ and $S_{312}^{out}\left( \omega ^{\prime }\right) $ first decrease than increase with increasing detuning. So correlation S_{123}^{out}\left( \omega ^{\prime }\right) $ is the best choice when we investigate the entanglement properties of the three cavity modes. Compared with Fig. 3(a) and 3(b), we find that the minimal values of correlations in Fig. 3(a) are smaller than those in Fig. 3(b). This indicates that the correlation degree is large when the analyzing frequency $\omega ^{\prime }$ is small. \section{Conclusion} In conclusion, we have examined the entanglement properties of three cavity modes interacting with three-level $\Lambda $ atomic system coupled by two extra classical fields. As the classical fields are stronger than the cavity fields, we adopt the dressed-atom approach to calculate the equation for the cavity fields. After tracing out the atomic variables, we obtain the master equation of the cavity modes and analyze the entanglement properties of the output fields. The tripartite entanglement of the three output fields is demonstrated theoretically by a sufficient inseparability criterion and the entanglement characteristics are presented. This scheme of three-mode continuous variable entanglement generation using atomic coherence is useful in quantum information processing. {\bf Acknowledgments} This work is supported by the Scientific Research Plan of the Provincial Education Department in Hubei (Grant No. Q20101304) and NSFC under Grant No. 11147153.
\section{Introduction} \label{S:intro} The Lambert $W$ function, implicitly defined by $W(x)\, e^{W(x)}=x$, has a long and quite convoluted 250-year history, but only recently has it become common to view this function as one of the standard special functions~\cite{Corless}. Applications range widely~\cite{Corless, Valluri:00}, from combinatorics (for instance in the enumeration of rooted trees)~\cite{Corless}, to delay differential equations~\cite{Corless}, to falling objects subject to linear drag~\cite{Vial:12}, to the evaluation of the numerical constant in Wien's displacement law~\cite{Stewart:11, Stewart:12}, to quantum statistics~\cite{Valluri:09}, to constructing the ``tortoise'' coordinate for Schwarzschild black holes~\cite{tortoise}, \emph{etcetera}. In this brief note I will indicate some apparently new applications of the Lambert $W$ function to the distribution of primes, specifically to the prime counting function $\pi(x)$ and estimating the $n$'th prime $p_n$. \section{The prime counting function $\pi(x)$}\label{S:Pi} \subsection{Upper bound\label{SS:Pi-U}} \begin{theorem} The prime counting function $\pi(x)$ satisfies \begin{equation} \pi(x) < {x\over W(x)} = e^{W(x)}; \qquad (\forall x \geq 0). \end{equation} \end{theorem} \paragraph{Proof:} First note that $ x \geq p_{\pi(x)}$. Second recall the standard result that $p_n > n\ln n$ for $n \geq 1$. (See Rosser~\cite{Rosser:38}, or any standard reference book on prime numbers, for example~\cite{Ribenboim:91, Ribenboim:96}.) Then $x \geq p_{\pi(x)} > \pi(x) \ln\pi(x)$, so we have $x > \pi(x) \ln\pi(x)$. Invert, noting that the RHS is monotone increasing, to see that $\pi(x) < x/ W(x)$, certainly for $\pi(x)\geq 1$ (corresponding to $x\geq2$). Then explicitly check validity of the inequality on the domain $x\in[0,2)$. Finally, use the definition of the Lambert $W$ function to note $ x/ W(x) = e^{W(x)}$. \subsection{Asymptotics\label{SS:Pi-A}} \begin{theorem} The prime number theorem, $\pi(x) \sim x/\ln x$, is equivalent to \begin{equation} \pi(x) \sim {x\over W(x)} = e^{W(x)}; \qquad (x\to\infty). \end{equation} \end{theorem} \paragraph{Proof:} Trivial. Note that asymptotically $W(x) \sim \ln x$. (The only potential subtlety is that we are using the principal branch of the $W$ function, denoted $W_0(x)$ whenever there is any risk of confusion~\cite{Corless}.) \noindent So we have \emph{both} a strict upper bound on $\pi(x)$ \emph{and} an asymptotic equality. See figure~\ref{F:pi-x-versus-W}. \begin{figure}[!htbp] \begin{center} \includegraphics[scale=0.75]{pi-x-versus-W.pdf} \caption{Prime counting function $\pi(x)$ and its upper bound $x/W(x)$ for $x\in(0,100)$. } \label{F:pi-x-versus-W} \end{center} \end{figure} \subsection{Lower bound\label{SS:Pi-L}} When it comes to developing an analogous lower bound on $\pi(x)$ the situation is more subtle. Consider the well-known upper bound on $p_n$~\cite{Rosser:41, Ribenboim:91, Ribenboim:96}: \begin{equation} p_n < n(\ln n+ \ln\ln n) = n \ln( n \ln n ); \qquad (n\geq 6). \label{E:lnln} \end{equation} Using logic similar to that of Theorem 1, we first note that $x < p_{\pi(x)+1}$, so it is easy to convert the above into the inequality \begin{equation} x< p_{\pi(x)+1} < [\pi(x)+1] \ln\{ [\pi(x)+1] \ln [\pi(x)+1] \}; \qquad (\pi(x) \geq 5). \end{equation} The function $z\to z \ln(z\ln z)$ is monotonic increasing on $z\in(1,\infty)$, and maps onto the range $(-\infty,\infty)$. So the inverse function $U(x)$, implicitly defined by \begin{equation} U(x) \ln\{ U(x) \ln U(x) \} = x \end{equation} certainly exists over the entire domain $(-\infty,\infty)$, and satisfies \begin{equation} \pi(x) > U(x)-1; \qquad (\pi(x) \geq 5; \;\; x \geq 11). \end{equation} Unfortunately while the inverse function $U(x)$ certainly exists, it has no simple closed form, either in terms of elementary functions or in terms of special functions (not even in terms of Lambert $W$). So to obtain a useful upper bound in terms of Lambert $W$ our strategy must be more indirect. Let us start with the elementary inequality \begin{equation} \ln x \leq {x\over e}, \end{equation} with equality only at $x=e$, and note that this implies \begin{equation} \ln x = {\ln(x^\epsilon)\over\epsilon} \leq {x^\epsilon\over\epsilon\; e}; \qquad (\forall \epsilon > 0), \end{equation} now with equality only at $x= e^{1/\epsilon}$. This inequality explicitly captures the well-known fact that the logarithm grows less rapidly than any positive power. Applied to the $n$'th prime this now implies \begin{equation} p_n < n \ln( n \ln n ) \leq n \ln\left(n^{1+\epsilon}\over\epsilon\; e\right) = n \{(1+\epsilon)\ln n -1-\ln\epsilon \}; \qquad (n\geq 6; \;\;\forall \epsilon>0). \end{equation} This inequality is weaker than previously, equation (\ref{E:lnln}), but much more tractable. Again using logic similar to that of Theorem 1, it is easy to convert this into the inequality \begin{equation} x< [\pi(x)+1] \ln\left\{ [\pi(x)+1]^{1+\epsilon}\over \epsilon\; e \right\}; \qquad (\pi(x) \geq 5; \;\;\forall \epsilon>0). \end{equation} This is now easily inverted to obtain: \begin{theorem} The prime counting function $\pi(x)$ satisfies \begin{equation} \pi(x) > { \displaystyle {x\over 1+\epsilon} \over W\left( \displaystyle{ x\over 1+\epsilon} \; (\epsilon\; e)^{-1/(1+\epsilon)} \right)} - 1. \end{equation} This inequality holds $\forall \epsilon>0$, \emph{and at least for $\pi(x)\geq 5$, corresponding to $x\geq 11$}. But, depending on $\epsilon$, the domain of validity may actually be larger. That is, the inequality holds for $\pi(x)\geq n_0(\epsilon)$ with $n_0(\epsilon)\leq 5$, corresponding to $x\geq x_0(\epsilon)$ with $x_0(\epsilon)<11$. \end{theorem} If for example we choose the specific case $\epsilon = e^{-1}$ we have (see figure~\ref{F:pi-x-versus-W2}): \begin{equation} \pi(x) > { \displaystyle {x\over 1+e^{-1}} \over W\left( \displaystyle{ x\over 1+e^{-1}} \; \right)} - 1; \qquad (x\geq 5). \end{equation} If instead we set $\epsilon = e^{-3}$ the domain of validity is now the entire positive half line (see figure~\ref{F:pi-x-versus-W2}): \begin{equation} \pi(x) > { \displaystyle {x\over 1+e^{-3}} \over W\left( \displaystyle{ x\over 1+e^{-3}} \; e^{2/(1+e^{-3})}\right)} - 1; \qquad (x\geq 0). \end{equation} This second special case, ($\epsilon=e^{-3}$), exhibits somewhat poorer bounding performance at intermediate values of $x$, but eventually overtakes the first special case, ($\epsilon=e^{-1}$), once $x\approx e^{2e+3}\approx 4600$, and then asymptotically provides a better bound. \begin{figure}[!htbp] \begin{center} \includegraphics[scale=0.75]{pi-x-versus-W2.pdf} \caption{Prime counting function $\pi(x)$, its upper bound $x/W(x)$, and two specific lower bounds corresponding to $\epsilon = e^{-1}$ and $\epsilon=e^{-3}$, for $x\in(0,100)$. } \label{F:pi-x-versus-W2} \end{center} \end{figure} Numerous variations on this theme can also be constructed, amounting to different ways of approximating the logarithms in equation (\ref{E:lnln}). \subsection{Implicit bounds\label{SS:p_n-implicit}} It is an old result (see for example Rosser~\cite{Rosser:41}) that $\forall \epsilon >0, \; \exists N(\epsilon): \forall n \geq N(\epsilon)$ \begin{equation} n\{\ln n + \ln\ln n -1 +\epsilon\} < p_n < n\{\ln n + \ln\ln n -1 +\epsilon\}. \end{equation} Without an explicit calculation/specification of $N(\epsilon)$ these bounds are qualitative, rather than quantitative. Nevertheless it may be of interest to proceed as follows: Note that $\ln x$ is convex, and therefore \begin{equation} \ln\ln x < \ln\ln x_0 + {\ln x - \ln x_0\over \ln x_0} \qquad (x >1; \;\;x_0 > 1). \end{equation} Consequently \begin{equation} p_n < n \left\{ \ln n + \ln\ln x_0 + {\ln n - \ln x_0\over \ln x_0} -1 +\epsilon \right\} \qquad (n \geq N(\epsilon); \;\;x_0 > 1). \end{equation} That is \begin{equation} p_n < n \left\{ \left[1 + {1\over \ln x_0} \right] \ln n + \left[\ln\ln x_0 - 2+\epsilon\right] \right\} \qquad (n \geq N(\epsilon); \;\;x_0 > 1). \end{equation} Now \emph{choose} $x_0$ to make the constant term vanish, that is set \begin{equation} \ln \ln x_0 = 2-\epsilon; \qquad \ln x_0 = e^{2-\epsilon}. \end{equation} Then we have \begin{equation} p_n < \left\{ 1 + e^{-2+\epsilon} \right\} n \ln n; \qquad (n \geq N(\epsilon); \;\;\epsilon > 0). \end{equation} Inverting this, a minor variant of the arguments above immediately yields: \begin{theorem} $\forall \epsilon >0, \; \exists M(\epsilon): \forall x \geq M(\epsilon)$ \begin{equation} \pi(x) > { \displaystyle {x\over 1+e^{-2+\epsilon}} \over W\left( \displaystyle{ x\over 1+e^{-2+\epsilon}} \; \right)} - 1. \end{equation} \end{theorem} It is now ``merely'' a case of estimating $M(\epsilon)$ to turn this into an explicit bound. For instance, it is known that for $\epsilon=3$ we have $N(\epsilon=3) = 55$, see Rosser~\cite{Rosser:41}. Explicitly checking the domain of validity, this then corresponds to \begin{corollary} \begin{equation} \pi(x) > { \displaystyle {x\over 1+e} \over W\left( \displaystyle{ x\over 1+e} \; \right)} - 1; \qquad (x \geq 3). \end{equation} \end{corollary} Unfortunately, the bound is not particularly tight once $x$ is large. A more recent result is that for $\epsilon=1/2$ we have $N(\epsilon=1/2) = 20$, see Rosser--Schoenfeld~\cite{Rosser:62}, equation (3.11) on page 69. This then corresponds to \begin{corollary} \begin{equation} \pi(x) > { \displaystyle {x\over 1+e^{-3/2}} \over W\left( \displaystyle{ x\over 1+e^{-3/2}} \; \right)} - 1; \qquad (x \geq 60). \end{equation} \end{corollary} In summary, we have used the Lambert $W$ function to obtain a number of bounds, and some general classes of bounds, on the prime counting function $\pi(x)$ in terms of the Lambert $W$ function $W(x)$. We shall now turn attention to the $n$'th prime $p_n$. \section{The $n$'th prime}\label{S:p_n} \subsection{Upper bound\label{SS:p_n-L}} \begin{theorem} The $n$'th prime $p_n$ satisfies \begin{equation} p_n < - n \;W_{-1}\left(-{1\over n}\right) \qquad (n \geq 4). \end{equation} Here $ W_{-1}(x)$ is the second real branch of the Lambert $W$ function, defined on the domain $x\in [-1/e, 0)$. \end{theorem} \paragraph{Proof:} We start from the fact that for $x\in{\mathbb{R}}$ we have $\pi(x) > x/\ln x$ for $x\geq 17$. See Rosser--Schoenfeld~\cite{Rosser:62}, equation (3.5) on page 69.\footnote{Observe that since $\pi(x)$ changes in a stepwise fashion, and $x/\ln x$ is monotone, this is equivalent for $n\in Z$ to $\pi(n) > (n+1)/\ln(n+1)$ for $n\geq 17$.} Then, since $x\geq p_{\pi(x)}$ and $x/\ln x$ is monotone increasing for $x>e$, we have $\pi(x) > x/\ln x \geq p_{\pi(x)}/\ln p_{\pi(x)}$. This can be written as $n > p_n /\ln p_n$, this inequality certainly being valid for $p_n\geq 17$, corresponding to $n\geq 7$. Inverting, (and appealing to the monotonicity of $x/\ln x$), we have $p_n < - n \;W_{-1}(-1/n)$, certainly for $n \geq 7$. Explicitly inspecting $n\in\{1,2,3,4,5,6\}$ shows that the actual range of validity is $n\geq4$. \begin{corollary} \begin{equation} p_n < - n \;W_{-1}\left(-{1\over n+e}\right) \qquad (n \geq 1). \end{equation} \end{corollary} \paragraph{Proof:} Note that $-W_{-1}(x)$ is monotone increasing on $[-e^{-1},0)$. So we see that $-W_{-1}(-1/[n+e]) > -W_{-1}(-1/n)$, and the claimed inequality certainly holds for $n\geq 4$. For $n\in\{1,2,3\}$ verify the claimed inequality by explicit computation. The virtue of this corollary is that it now holds for all positive integers. See figure~\ref{F:p_n-versus-W}. There are many other variations on this theme that one could construct. \begin{figure}[!htbp] \begin{center} \includegraphics[scale=0.75]{p_n-versus-W.pdf} \caption{ Two upper bounds on the $n$'th prime $p_n$. } \label{F:p_n-versus-W} \end{center} \end{figure} \subsection{Asymptotics\label{SS:p_n-A}} \begin{theorem} The prime number theorem, which can be written in the form $p_n \sim n\ln n$, is equivalent to \begin{equation} p_n \sim - n \;W_{-1}\left(-{1\over n}\right) ; \qquad (n\to\infty). \end{equation} \end{theorem} \paragraph{Proof:} Trivial. Consider the asymptotic result \begin{equation} W_{-1}(x) = \ln(-x) - \ln(-\ln(-x)) + o(1) \qquad (x\to 0^-). \end{equation} Then \begin{equation} - n \;W_{-1}\left(-{1\over n}\right) = n \{ \ln n + \ln\ln n + o(1)\}. \end{equation} \paragraph{Comments:} Note that use of the Lambert $W$ function automatically yields the first two terms of the Cesaro--Cippola asymptotic expansion~\cite{Cesaro, Cippola}: \begin{equation} p_n = n \{ \ln n + \ln\ln n -1 + o(1)\}. \end{equation} We can even obtain the first \emph{three} terms of the Cesaro--Cippola asymptotic expansion by refining the prime number theorem slightly as follows: \begin{theorem} \begin{equation} p_n \sim - n \;W_{-1}\left(-{e\over n}\right) ; \qquad (n\to\infty). \end{equation} \end{theorem} Finally, we note that use of the Lambert $W$ function yields \emph{both} a strict upper bound \emph{and} an asymptotic result. \subsection{Lower bound\label{SS:p_n-U}} To obtain an lower bound on $p_n$ we start with an upper bound on $\pi(x)$. Consider for instance the standard result~\cite{Rosser:62}: \begin{equation} \pi(x) < {x\over\ln x - {3\over2}}; \qquad (x>e^{3/2}) \end{equation} Note that the RHS of this inequality is monotone increasing for $x>e^{5/2}$. Now we always have $p_{\pi(x)} \leq x<p_{\pi(x)+1}$, so \begin{equation} n < {p_{n+1} \over \ln p_{n+1} - {3\over2}}. \end{equation} This holds at the very least for $p_n > e^{5/2}$, corresponding to $n \geq 6$, but an explicit check shows that it in fact holds for $n\geq 2$. This is perhaps more clearly expressed as \begin{equation} n -1 < {p_{n} \over \ln p_{n} - {3\over2}}; \qquad (n \geq 3). \end{equation} Inverting, and noting the constraint arising from the domain of definition of $W_{-1}$, we now obtain: \begin{theorem} \begin{equation} p_n > - (n-1) \; W_{-1}\left(-{e^{3/2}\over n-1}\right); \qquad ( n \geq 14). \end{equation} \end{theorem} \begin{figure}[!htbp] \begin{center} \includegraphics[scale=0.75]{p_n-versus-W2.pdf} \caption{ Lower bound on the $n$'th prime $p_n$. } \label{F:p_n-versus-W2} \end{center} \end{figure} See figure (\ref{F:p_n-versus-W2}). There are many other variations on this theme that one could construct. \subsection{Implicit bounds\label{SS:p_n-implicit}} It is an old result (see for example Rosser~\cite{Rosser:41}) that $\forall \epsilon >0, \; \exists N(\epsilon): \forall n \geq N(\epsilon)$ \begin{equation} {x\over\ln x - 1 + \epsilon } <\pi(x) < {x\over\ln x - 1 - \epsilon }. \end{equation} Again note that without an explicit calculation of $N(\epsilon)$ these bounds are qualitative, rather than quantitative. Nevertheless it may be of interest to point out that a minor variant of the arguments above immediately yields: \begin{theorem} $\forall \epsilon >0, \; \exists M(\epsilon): \forall n \geq M(\epsilon)$ \begin{equation} - n \; W_{-1}\left(-{e^{1-\epsilon}\over n}\right) > p_n > - (n-1) \; W_{-1}\left(-{e^{1+\epsilon}\over n-1}\right). \end{equation} \end{theorem} It is now ``merely'' a case of estimating $M(\epsilon)$ to turn these into explicit bounds. We have already seen that $\epsilon =1$ provides a widely applicable upper bound, and $\epsilon=1/2$ a widely applicable lower bound. Taking $\epsilon\to 0$ now makes it clear why \begin{equation} p_n \sim - n \;W_{-1}\left(-{e\over n}\right) ; \qquad (n\to\infty), \end{equation} is such a good asymptotic estimate for $p_n$. \section{Discussion}\label{S:Discussion} While the calculations carried out above are very straightforward, almost trivial, it is perhaps the shift of viewpoint that is more interesting. The Lambert $W$ function provides (in this context) a ``new'' special function to work with, one which may serve to perhaps simplify and unify many otherwise disparate results. It is perhaps worth noting that the infamous ``$\ln\ln x$'' terms that infest the analytic theory of prime numbers will automatically appear as the sub-leading terms in asymptotic expansions of the Lambert $W$ function. \section*{Acknowledgments} This research was supported by the Marsden Fund, and by a James Cook fellowship, both administered by the Royal Society of New Zealand.
\section{Introduction} Considered here is the construction and study of fixed point algorithms for the numerical resolution of nonlinear systems of the form \begin{eqnarray} L{ u}=N({ u}),\quad u\in \mathbb{R}^{m}, \quad m>1,\label{m1} \end{eqnarray} where $L$ is a nonsingular $m\times m$ real matrix and $N:\mathbb{R}^{m}\rightarrow \mathbb{R}^{m}$ is an homogeneous function of the components of $u$ with degree $p, |p|>1$. These systems are typical in many applications, including the approximation to equilibria in mechanical systems and the numerical generation of traveling waves and ground states in nonlinear dispersive systems for water waves and nonlinear optics. {(In this last context, $m$ would represent the number of discretization points.)} More generally, (\ref{m1}) may appear when generating relative equilibria or coherent structures, \cite{champneyss}. {We denote by $u^{\ast}$ a solution of (\ref{m1}), that is \begin{eqnarray} Lu^{\ast}=N(u^{\ast}).\label{m1a} \end{eqnarray} } The classical fixed point algorithm for (\ref{m1}) has the following formulation. If $u_{0}\neq 0$, the approximation to $u^{*}$ in (\ref{m1a}) at the $(n+1)$-th iteration is given by the recurrence \begin{eqnarray} Lu_{n+1}=N(u_{n}), \quad n=0,1,\ldots\label{m1b} \end{eqnarray} { The method (\ref{m1b}) is not usually convergent for this kind of problems. Note that if \begin{eqnarray} S=L^{-1}N^{\prime}(u^{\ast}),\label{m2b} \end{eqnarray} stands for the iteration matrix at $u^{\ast}$ (and where $N^{\prime}(u)$ denotes the Jacobian of $N$ at $u$), then the homogeneous character of $N$ implies that $ N^{\prime}(u^{\ast})u^{\ast}=pN(u^{\ast}); $ therefore, using (\ref{m1a}), \begin{eqnarray*} S(u^{\ast})u^{\ast}=L^{-1}N^{\prime}(u^{\ast})u^{\ast}=pL^{-1}N(u^{\ast})=pu^{\ast}. \end{eqnarray*} Thus, $u^{\ast}$ is an eigenvector of $S$ associated to an eigenvalue $\lambda=p$ with $|p|>1$. } The methods presented here generalize the so-called Petviashvili method. From a starting iteration $u_{0}\neq 0$, the Petviashvili method generates the recurrence \begin{eqnarray} m(u_{n})&=&\frac{\langle Lu_{n},u_{n}\rangle}{\langle N(u_{n}),u_{n}\rangle},\label{m2}\\ Lu_{n+1}&=&m(u_{n})^{\gamma}N(u_{n}), \quad n=0,1,\ldots\label{m3} \end{eqnarray} where {here and in the rest of the paper} $\langle\cdot,\cdot\rangle$ stands for the Euclidean inner product and $\gamma$ is a free real parameter. The term (\ref{m2}) is called stabilizing factor and, in the case of convergence, must tend to one. The origin of the method is in \cite{petviashvili}, focused on the search for lump solitary waves of the Kadomtsev-Petviashvili I (KPI) equation \begin{eqnarray} \left(u_{t}+2uu_{x}+u_{xxx}\right)_{x}=u_{yy},\quad t>0, x,y\in \mathbb{R}\label{crete1} \end{eqnarray} of the form $u(x,y,t)=c\varphi(X,Y)=c\varphi(\sqrt{c}(x-ct),cy), c>0$. The profile $\varphi$ must satisfy \begin{eqnarray*} \partial_{XX}\left(-\varphi+\partial_{XX}\varphi\right)-\partial_{YY}\varphi= -\partial_{XX}\varphi^{2}, \end{eqnarray*} which, in terms of the $2$-D Fourier Transform, \begin{eqnarray*} \widehat{\varphi}({k_{x}, k_{y}})=\int_{-\infty}^{\infty}\int_{-\infty}^{\infty} \varphi(x,y)e^{-ik_{x}x}e^{-ik_{y}y}dxdy, \end{eqnarray*} is converted into an algebraic system \begin{eqnarray} &&\widehat{\varphi}({k_{x}, k_{y}})=G({k_{x}, k_{y}})A({k_{x}, k_{y}}),\nonumber\\ &&G({k_{x}, k_{y}})=\frac{k_{x}^{2}}{k_{x}^{4}+k_{x}^{2}+k_{y}^{2}},\quad A({k_{x}, k_{y}})=\widehat{\varphi^{2}}({k_{x}, k_{y}})\label{crete3} \end{eqnarray} The divergence of the classical fixed point algorithm, applied to (\ref{crete3}) forces to consider a new iteration system, of the form \begin{eqnarray*} \widehat{\varphi}({k_{x}, k_{y}})=m(\varphi)^{\gamma}G({k_{x}, k_{y}})A({k_{x}, k_{y}}), \end{eqnarray*} where the stabilizing factor $m(\varphi)$ is defined as \begin{eqnarray*} m(\varphi)=\frac{s_{1}}{s_{2}},\quad s_{1}=\int\int|\widehat{\varphi}|^{2}dk_{x}dk_{y},\, s_{2}=\int\int\frac{k_{x}^{2}}{k_{x}^{4}+k_{x}^{2}+k_{y}^{2}}\widehat{\varphi^{2}} \overline{\widehat{\varphi}}dk_{x}dk_{y}, \end{eqnarray*} {($\overline{\widehat{\varphi}}$ denotes the complex conjugate of ${\widehat{\varphi}}$)} where $\gamma$ is a free real parameter. In the case of (\ref{crete1}), $\gamma$ is taken approximately $2$, \cite{petviashvili}. On the other hand, for the exact profile $\varphi$, $m(\varphi)=1$. The Petviashvili method has become popular as a technique to generate special solutions in partial differential equations of interest in water waves and nonlinear optics. It takes part of a large family of methods designed to this goal, which includes variants of the Newton's method, \cite{yang}, {modified conjugate gradient methods applied to nonlinear problems, \cite{lakoba}}, squared operator methods, \cite{yangl2}, imaginary-time evolution methods, \cite{yangl1} or different variational procedures, \cite{garciap,baod,caliario} . Some literature about (\ref{m2}), (\ref{m3}), from the original paper, \cite{petviashvili}, is now briefly reviewed. Pelinovsky and Stepanyants, \cite{pelinovskys}, analyze the continuous version of the method to approximate solitary wave profiles of the nonlinear dispersive models \begin{eqnarray*} u_{t}-\mathcal{L}u_{x}+pu^{p-1}u_{x}&=&0,\quad p>1, \quad t>0, \quad x\in\mathbb{R}, \end{eqnarray*} where $\mathcal{L}$ is a pseudodifferential operator with positive Fourier symbol. On the other hand, Lakoba and Yang, \cite{lakobay,lakobay2}, introduce a generalized version of the procedure, for more general systems of the form \begin{eqnarray*} -Mu+F(x,u)=0,\quad u\rightarrow 0, \quad |x|\rightarrow\infty \end{eqnarray*} where $M$ is positive definite, self-adjoint operator and $F$ is nonlinear (see also \cite{yang2}). Finally, Ablowitz and Musslimani, \cite{ablowitzm}, (see also \cite{ablowitzm2}) propose an alternative of the algorithm, the spectral renormalization method, with application to generate numerically ground state profiles for systems of NLS type \begin{eqnarray*} iU_{z}+\Delta U-V(x)U+f(|U|^{2})U=0. \end{eqnarray*} Some new results contained in this paper are described below. \begin{itemize} \item Based on the philosophy the Petviashvili method was devised with, new fixed point methods are derived. They can be considered as a Petviashvili type family of methods. \item From the view point of the classical algorithm, the corresponding iteration functions are designed to filter the harmful directions of the errors leading to divergence, in such a way that convergence results are obtained under the same hypotheses as those of the Petviashvili method. Here it is worth mentioning two types of convergence. The first one has the classical sense, with the requirement (among others) of isolated fixed points. However, in traveling wave generation, it is very typical that the system of equations admits a symmetry group {(usually related to translational or rotational invariance of the system)}. In this case, fixed points cannot be isolated and the convergence must be understood in the orbital sense, that is, for the orbits of fixed points. Convergence results for both cases will be given. This study complements some previous results of convergence presented in the literature for the Petviashvili method, \cite{pelinovskys,lakobay,lakobay2}. \end{itemize} The structure of the paper is as follows: in Section \ref{sec2} and starting from the Petviashvili method (\ref{m2}), (\ref{m3}), the new family of fixed point algorithms are constructed and analyzed. A comparison of efficiency of some of them is also carried out. Section \ref{sec3} will treat the derivation of general conditions for the local convergence of the methods. The first part of the study assumes the existence of a neighborhood where the fixed point $u^{*}$ is unique. The spectral analysis of the iteration matrix (\ref{m2b}) of the classical fixed point algorithm (or, equivalently, the pencil $A(\lambda)=\lambda L-N^{\prime}(u^{*})$) is used. The local convergence of the methods can be achieved even when (\ref{m2b}) admits eigenvalues with modulus greater than or equals one. The results are illustrated with several numerical examples, concerning the generation of localized ground state solutions of nonlinear Schr\"{o}dinger type equations with potentials. On the other hand, some other applications of the methods suggest to analyze a case where the hypothesis of local uniqueness of the fixed point does not hold, in the sense that the system (\ref{m1}) admits a group of symmetries, generating orbits of solutions. From the point of view of the analysis, the existence of a symmetry group in (\ref{m1}) is associated to the formation of the eigenvalue one in the pencil, \cite{champneyss}, and leads, in a natural way, to the concept of orbital convergence. Section \ref{sec3} is finished off with the corresponding results of convergence for this case and they will be illustrated by the generation of soliton solutions of the nonlinear Schr\"{o}dinger equation. The present paper is a first part of a study of the methods carried out by the same authors. It will be followed by a second part, in which some particular, relevant cases of systems (\ref{m1}) are emphasized and where the effect of the introduction of acceleration techniques is studied. \section{Derivation and convergence analysis of the algorithms} \label{sec2} \subsection{Derivation} The following fixed point methods for the iterative resolution of (\ref{m1}) are introduced. If ${ u}_{0}\neq 0$, the iterations ${u}_{n}, n=1,2,\ldots$ are generated by a formula of the form \begin{eqnarray} Lu_{n+1}=s(u_{n})N(u_{n}), \quad n=0,1,\ldots\label{m4} \end{eqnarray} where $s:\mathbb{R}^{m}\rightarrow \mathbb{R}$ is a $C^{1}$ function satisfying the following properties: \begin{itemize} \item[(P1)] A set of fixed points of the iteration operator \begin{eqnarray} F(u)=s(u)L^{-1}N(u),\label{iterop} \end{eqnarray} coincides with a set of fixed points of (\ref{m1}). This means that: (a) if $u^{*}$ is a solution of (\ref{m1}) then $s(u^{*})=1$; (b) inversely, if the sequence $\{u_{n}\}_{n=0}^{\infty}$, generated by (\ref{m4}), converges to some $y$, then $s(y)=1$ (and, consequently, $y$ is a solution of (\ref{m1})). \item[(P2)] $s$ is homogeneous with degree $q$ such that $|p+q|<1$. \end{itemize} Note that, in particular, the choice \begin{eqnarray} s(u)=\left(\frac{\langle Lu,u\rangle}{\langle N(u),u\rangle}\right)^{\gamma},\quad q=\gamma(1-p),\label{m5} \end{eqnarray} leads to the Petviashvili method (\ref{m2}), (\ref{m3}). Then, (\ref{m4}) can be considered as a generalization and justifies that $s$ will be also called a stabilizing factor. Several examples are the following: \begin{itemize} \item The term (\ref{m5}) can be generalized by considering any $C^{1}$ homogeneous function $f:\mathbb{R}^{m}\rightarrow \mathbb{R}$ with degree greater than or equals one and taking \begin{eqnarray} s_{f}(u)=\left(\frac{\langle Lu,f(u)\rangle}{\langle N(u),f(u)\rangle}\right)^{\gamma},\quad q=\gamma(1-p), \quad |p+q|<1.\label{m6a} \end{eqnarray} \item Another alternative is the use of norms, with \begin{eqnarray} s_{r}(u)=\left(\frac{||Lu||_{r}}{||N(u)||_{r}}\right)^{\gamma},\quad q=\gamma(1-p), \quad |p+q|<1,\label{m6b} \end{eqnarray} where if $u=(u_{1},\ldots,u_{m})^{T}$ then $||u||_{r}=\left(|u_{1}|^{r}+\ldots+|u_{m}|^{r}\right)^{1/r}, 1\leq r\leq +\infty$, with $r=+\infty$ standing for the usual maximum norm. The case $r=1$ was considered in \cite{ablowitzm2}. \end{itemize} \subsection{First numerical experiments} \label{sec22} Displayed here are some numerical experiments concerning the performance of the methods, according to the choice of the stabilizing factor. As an example we consider the problem of generating lump solitary waves in the 2D Benjamin equation \begin{eqnarray} \label{ben2da} \left(\eta_{t}+\alpha (\eta^{2})_{x}-\beta \mathcal{H}(\eta_{xx})+\delta \eta_{xxx} \right)_{x}-\eta_{zz}=0, \end{eqnarray} where $\alpha, \beta, \delta\geq 0$ and $\mathcal{H}$ stands for the Hilbert transform with respect to $x$: \begin{eqnarray*} \mathcal{H}f(x)=\frac{1}{\pi}P.V. \int_{-\infty}^{\infty} \frac{f(y)}{x-y}dy. \end{eqnarray*} Equation (\ref{ben2da}) is analyzed in \cite{kim,kima1,kima2}. It appears as an extension of the one-dimensional equation derived by Benjamin, \cite{ben0,ben1,ben2}, and modeling the propagation of waves at the interface of two ideal fluids, with a bounded upper layer and the heavier one with infinite depth, and under the presence of interfacial tension. The two-dimensional version incorporates weak transverse variations. The form (\ref{ben2da}) contains particular cases, such as the Kamdotsev-Petviashvili (KP-I) equation, \cite{kadomtsevp,manakovzbim} ($\beta=0, \delta>0$) and the two-dimensional Benjamin-Davis-Ono (BDO) equation, corresponding to $\delta=0, \beta>0$, \cite{ablowitzs}. A normalized form of (\ref{ben2da}) \begin{eqnarray} \label{ben2db} \left(\eta_{t}+ (\eta^{2})_{x}-2{\Gamma} \mathcal{H}(\eta_{xx})+\eta_{xxx} \right)_{x}-\eta_{zz}=0, \end{eqnarray} is derived in \cite{kima2} and will be adopted here. The parameter $\Gamma\geq 0$ is related to the interfacial tension and the densities of the fluids. Finally, for localized solutions, the constraint \begin{eqnarray} \label{constraint} \int_{-\infty}^{\infty} \eta(x,z,t)dx=0, \end{eqnarray} (zero total mass condition) is assumed, as in the KP and BDO equations \cite{katsisa}. \begin{figure}[htbp] \centering \subfigure{\subfigure{ \includegraphics[width=5.5cm]{sw2d1r.pdf}} \subfigure{ \includegraphics[width=5.5cm]{sw2d2r.pdf}} } \end{figure} \begin{figure}[htbp] \centering \subfigure{\subfigure{ \includegraphics[width=5.5cm]{sw2d3r.pdf}} \subfigure{ \includegraphics[width=5.5cm]{sw2d4r.pdf}} } \end{figure} \begin{figure}[htbp] \centering \subfigure{\subfigure{ \includegraphics[width=5.5cm]{sw2d5r.pdf}} \subfigure{ \includegraphics[width=5.5cm]{sw2d6r.pdf}} } \end{figure} \begin{figure}[htbp] \centering \subfigure{\subfigure{ \includegraphics[width=5.5cm]{sw2d7r.pdf}} \subfigure{ \includegraphics[width=5.5cm]{sw2d8r.pdf}} } \caption{Solitary wave generation of (\ref{ben2da}) with Petviashvili method. The approximated profiles correspond to $\Gamma=0.5,0.9,0.95,0.99$ (left). On the right, the corresponding $X$ and $Z$ cross sections {are shown} (solid and dashed-dotted lines, respectively) .} \label{fexample41} \end{figure} The search for lump solitary wave solutions of (\ref{ben2db}) \begin{eqnarray*} \eta(x,z,t)=\eta(X,Z),\quad X=x-c_{s}t, \quad Z=z, \end{eqnarray*} leads to the equation \begin{eqnarray} \label{lumpsw} \left(-c_{s}\eta+\eta^{2}-2\gamma \mathcal{H}(\eta_{X})+\eta_{XX}\right)_{XX}-\eta_{ZZ}=0, \end{eqnarray} for the profile $\eta$. In \cite{kima2} it is shown that (\ref{lumpsw}) admits lumps of wavepacket type, as well as lumps of KP-I type {(see Figures \ref{fexample41}(d) and \ref{fexample41}(a) respectively)}. The bifurcation point corresponds to $\Gamma=1$. As an alternative to the numerical procedure performed in that paper, some lumps will be here generated numerically, by using numerical continuation in $\Gamma$ from the known lump solitary wave solution of the KP-I equation (corresponding to $\Gamma=0$), \cite{manakovzbim}. Taking two-dimensional Fourier transform in (\ref{lumpsw}) we have \begin{eqnarray} \left({k_{x}}^{2}\left(c_{s}+2\Gamma |k_{x}|+k_{x}^{2}\right)+{k_{y}}^{2}\right)\widehat{\eta}(k_{x},k_{y})={k_{x}^{2}}\widehat{\left(\eta^{2}\right)}(k_{x},k_{y}).\label{lumpsw2} \end{eqnarray} The continuation algorithm until the computation of the profile at certain value $\Gamma^{*}$ consists of defining an homotopic path \begin{eqnarray*} \Gamma_{0}=0<\Gamma_{1}<\cdots <\Gamma_{M}=\Gamma^{*}<1, \end{eqnarray*} and solving numerically (\ref{lumpsw2}) with the method (\ref{m4}) at each $\Gamma_{j}$ and initial iteration given by the last computed iterate at the previous stage $\Gamma_{j-1}$. The procedure starts with the exact KP-I lump at $\Gamma=\Gamma_{0}=0$. {The numerical resolution of (\ref{lumpsw2}) with (\ref{m4}) is now described. Note that (\ref{lumpsw2}) for $k_{x}=k_{y}=0$ leaves the $(0,0)$-Fourier component free and the value $\widehat{\eta}(0,0)=0$ is set by the zero total mass condition (\ref{constraint}). System (\ref{lumpsw2}) is discretized by using a Fourier collocation scheme and, with the notation of (\ref{m1}), the corresponding discrete system leads to a singular matrix $L$. In order to solve this, the $(0,0)$-Fourier component of the approximation is set to zero, then the resulting system for the rest of the Fourier components is not singular and is therefore iteratively solved with (\ref{m4}), \cite{pelinovskys}.} The numerical results are shown in Figure \ref{fexample41}. {(They correspond to $c_{s}=1$.)} On the left, different lump profiles, associated to several values of $\Gamma$ are displayed. On the right, the corresponding $X-$ and $Z-$ cross sections are represented. The convergence of the procedure is illustrated by Figure \ref{fexample42}. For the case $\Gamma=0.99$, it shows the limiting behaviour of the stabilizing factor (left) and the residual errors (right) \begin{eqnarray} RE_{n}=||Lx_{n}-N(x_{n})||,\label{res} \end{eqnarray} {(where here and in the rest of the paper $||\cdot ||$ stands for the Euclidean norm)} both as functions of the number of iterations as for several choices of the stabilizing factor, from the two families (\ref{m6a}), (\ref{m6b}). \begin{figure}[htbp] \centering \subfigure[]{ \includegraphics[width=7cm]{sw2d_stab_r.pdf} } \subfigure[]{ \includegraphics[width=7cm]{sw2d_error_r.pdf} } \caption{Convergence results of the Petviashvili type methods (\ref{m4}) for (\ref{lumpsw}): (a) Discrepancy between the stabilizing factor vs number of iterations (semilog scale). (b) Logarithm of the residual errors vs number of iterations. Solid line: Petviashvili method (\ref{m6a}) with {$f(x)=x$}; dashed line: (\ref{m6b}) with $r=2$; dashed-dotted line: (\ref{m6b}) with $r=1$.} \label{fexample42} \end{figure} As a representative of (\ref{m6a}), the original Petviashvili method {($f(x)=x$)} has been compared with two methods with stabilizing factors of the form (\ref{m6b}), corresponding to $r=1,2$. Figure \ref{fexample42}(a) shows that the convergence of the stabilizing factor to one is more efficient with the Petviashvili method: it provides an error with less iterations and, for a fixed number of iterations, it gives a smaller discrepancy. This is also the conclusion when analyzing Figure \ref{fexample42}(b), concerning the behaviour of the residual error with the number of iterations. However, this better performance of the Petviashvili method in this example is not big enough to be conclusive and to rule the rest of the methods out in a general situation. {We have the impression} that in terms of the computational effort, the methods are more or less equivalent, with a slight superiority of (\ref{m2}), (\ref{m3}). For that reason, this will be considered as a representative of (\ref{m4}) for the rest of the experiments in this paper (see Section \ref{sec3}). {Finally, although the main goal of the example is illustrating a comparison between some methods of the family (\ref{m4}), it is worth mentioning that (\ref{ben2db}) is translationally invariant. Thus, the convergence must be understood in the sense analyzed in Section \ref{sec33} (orbital convergence).} \section{Analysis of convergence} \label{sec3} As mentioned in the Introduction, the convergence of the methods (\ref{m4}) can be divided in two cases, depending on the character of $u^{*}$ as fixed point of (\ref{m1}). {In what follows, the Jacobian of the iteration operator (\ref{iterop}) at a fixed point $u^{*}$ satisfying (\ref{m1a}), \begin{eqnarray} F^{\prime}(u^{*})=S+u^{*}\left(\nabla s(u^{*})\right),\label{m6c} \end{eqnarray} will be used. {(In (\ref{m6c}), the gradient $\nabla s(u^{*})$ is taken as a row vector.)}} \subsection{Convergence (classical sense)} We first define the pencil {$A(\lambda)=\lambda L-N^{\prime}(u^{*})$, where $u^{*}$ satisfies (\ref{m1a}). Note that, since $L$ is nonsingular, the zeros of $A(\lambda)$ coincide with the spectrum of the iteration matrix (\ref{m2b}), see \cite{Demmel}, Section 4.5 and \cite{GolubV}, Section 7.7. (In particular, $\lambda=p$ is a zero of $A(\lambda)$ with $A(p)u^{*}=0$.) We also remind that an eigenvalue $\lambda$ of a matrix is semisimple if the corresponding geometric and algebraic multiplicities are the same; that is, if the dimension of the associated eigenspace coincides with the order of $\lambda$ as zero of the characteristic polynomial.} \begin{thm} \label{theorem1} Assume that \begin{itemize} \item[(H1)] There exists $R>0$ such that $u^{*}$ in (\ref{m1a}) is the unique fixed point of (\ref{m1}) in $B(u^{*},R)=\{u\in \mathbb{R}^{m} / ||u-u^{*}||<R\}$. \end{itemize} Take $u_{0}\neq 0$ and assume the following hypotheses on the zeros of $A(\lambda)$: \begin{itemize} \item[(i)] $\lambda=p$ is simple. \item[(ii)] The rest of $\lambda$ satisfies $|\lambda|\leq 1$. \item[(iii)] If $|\lambda|=1\Rightarrow\left\{\begin{matrix}\lambda \quad \mbox {is semisimple}\\ u_{0} \quad \mbox {does not have component in } Ker A(\lambda)\end{matrix}\right.$ \end{itemize} Then the method (\ref{m4}), with $s$ satisfying (P1) and (P2), is locally convergent, that is, there is a neighborhood $W$ of $u^{*}$ such that if $u_{0}\in W, u_{0}\neq 0$, the sequence $\{u_{n}\}_{n=0}^{\infty}$ generated by (\ref{m4}), converges to $u^{*}$. The optimal rate of convergence is obtained with $q=-p$. \end{thm} { {\em Proof}. The errors $e_{n}=u_{n}-u^{*}, n=0,1,\ldots,$ satisfy \begin{eqnarray} e_{n+1} =F^{\prime}(u^{*})e_{n} +O(||e_{n}||^{2}),\quad n=0,1,\ldots,\label{m7} \end{eqnarray} where $F^{\prime}(u^{*})$ is given by (\ref{m6c}). According to the hypotheses (i)-(iii), $e_{n}$ can be decomposed \begin{eqnarray} e_{n}=\alpha_{n}u^{*}+z_{n}, \quad \alpha_{n}\in \mathbb{R}, \quad z_{n}\in V, \quad S(V)\subset V,\label{m8} \end{eqnarray} where $V$ is a $S$ invariant supplementary subspace of $span(u^{*})$. ($V$ is the sum of the $S$-invariant subspaces associated to the eigenvalues of $S$ different from $\lambda=p$.) Substituting (\ref{m8}) into (\ref{m7}) and neglecting second order terms the system \begin{eqnarray} \alpha_{n+1}&=&\alpha_{n}(p+\left(\nabla s(u^{*})\right)u^{*})+\left(\nabla s(u^{*})\right)z_{n}\nonumber\\ &=&\alpha_{n}(p+q)+\left(\nabla s(u^{*})\right)z_{n}\label{m9}\\ z_{n+1}&=&Sz_{n},\label{m10} \end{eqnarray} is obtained. (Last equality in (\ref{m9}) comes from (P1) and (P2), which imply that $\left(\nabla s(u^{*})\right)u^{*}=qs(u^{*})=q$.) Due to hypotheses (ii) and (iii), the sequence $\{z_{n}\}_{n=0}^{\infty}$ in (\ref{m10}) converges to zero. This and property (P2) imply then that $\alpha_{n}\rightarrow 0$ in (\ref{m9}), leading to local convergence. Finally, the fastest rate of convergence occurs when the factor $p+q$ in (\ref{m9}) is zero.$\Box$ } In summary, under the hypotheses of Theorem \ref{theorem1}, the iteration map (\ref{iterop}) is contractive in a neighborhood of the fixed point, with the fastest rate of convergence when $q=-p$. Condition (iii) was already obtained in \cite{pelinovskys}, for the equations treated there and the continuous version of the Petviashvili method. In this sense, Theorem \ref{theorem1} establishes the fact that (iii) is one of the sufficient conditions for the local convergence for more general methods and in more general systems. { Assumption (iii) also suggests a dependence of the convergence on the choice of the initial iterate, not only in the sense required by the local convergence, but also because $u_{0}$ must contain the correct directions. The contribution to the iteration error of the components of $u_{0}$ in these \lq harmful\rq\ eigendirections that (iii) is concerned with, can be sketched as follows. Assume for simplicity that $S$ contains one semisimple eigenvalue $\lambda_{0}$ with $|\lambda_{0}|=1$ and the rest of the spectrum (except $\lambda=p$) is below one in modulus. In (\ref{m8}), we can decompose the term $z_{n}$ in the form $z_{n}=v_{n}+w_{n}$ with $v_{n}\in Ker(\lambda_{0} I-S), w_{n}\in V\backslash Ker(\lambda_{0} I-S)$. (Thus, $v_{n}$ can be written as a linear combination of a basis of $Ker(\lambda_{0} I-S)$, with the coordinates depending on $n$.) Now, (\ref{m9}) and (\ref{m10}) can be written as \begin{eqnarray} &&\alpha_{n+1}=\alpha_{n}(p+q)+\left(\nabla s(u^{*})\right)v_{n} +\left(\nabla s(u^{*})\right)w_{n},\label{m91}\\ && v_{n+1}=Sv_{n}=\lambda_{0}v_{n},\label{m92}\\ && w_{n+1}=Sw_{n}.\label{m101} \end{eqnarray} Therefore, due to (\ref{m101}) and the previous assumptions on the spectrum of $S$, $w_{n}$ goes to zero, while (\ref{m92}) implies $v_{n}=\lambda_{0}^{n}v_{0}, n=0,1,\ldots,$ being $v_{0}$ the component of $e_{0}$ in $Ker(\lambda_{0}I-S)$ (which is to say the component of $u_{0}$ in $Ker(\lambda_{0}I-S)$). Then (\ref{m91}) becomes \begin{eqnarray*} \alpha_{n+1}=\alpha_{n}(p+q)+\lambda_{0}^{n}\left(\nabla s(u^{*})\right)v_{0} +\left(\nabla s(u^{*})\right)w_{n}. \end{eqnarray*} Thus, in general, $\alpha_{n}$ would be $O(||v_{0}||)$ as $n\rightarrow\infty$. As proved by Theorem \ref{theorem1}, if $v_{0}=0$ (condition (iii)) and using (P2), then $\alpha_{n}$ tends to zero as $n\rightarrow\infty$. The previous arguments also say that the errors would behave as the size of the component $v_{0}$. } The comparison between the matrices $S$ and $F^{\prime}(u^{*})$ reveals that the stabilizing factor acts like a filter for the harmful direction of the error that leads to the nonconvergence of the classical fixed-point algorithm in this case. The spectrum of $F^{\prime}(x^{*})$ differs from that of $S$ in the dominant eigenvalue $p$, which is transformed to some less than one (or, eventually, to zero eigenvalue if the optimal case is taken), leading to convergence if the rest of the spectrum of $S$, with probably the help of the initial iteration, behaves in the way described in Theorem \ref{theorem1} {see the numerical experiments in sections \ref{sec32} and \ref{sec34})}. \subsection{Some examples} \label{sec32} As a first example, the application of the Petviashvili method to generate localized ground state solutions of the nonlinear Schr\"{o}dinger (NLS) model \begin{eqnarray} \label{doub_well11} iu_{t}+\partial_{xx} u+V(x)u-|u|^{2}u=0, \end{eqnarray} with {a potential $V(x)$} is considered. A ground state solution has the form $u(x,t)=e^{i\mu t}U(x)$, where $\mu\in \mathbb{R}$ and the profile $U(x)$ is assumed to be real and localized ($U\rightarrow 0,\; |x|\rightarrow\infty$) and then must satisfy \begin{eqnarray} \label{doub_well12} U^{\prime\prime}(x)+V(x)U(x)-\mu U(x)-U^{3}(x)=0. \end{eqnarray} The Petviashvili method (as a representative of the family (\ref{m4}), see Section \ref{sec22}) can be applied to a discretization of (\ref{doub_well12}). One way to treat numerically the problem is approximating (\ref{doub_well12}) on a sufficiently long interval $(-l,l)$ and then discretizing the corresponding system for the profile. As an illustration, the discretization based on a Fourier collocation method for the periodic problem is taken, in such a way that the corresponding discrete equations have the form (\ref{m1}) with \begin{eqnarray*} L=D^{2}+{\rm diag}(V)-\mu I,\quad N(U_{h})=-U_{h}.^{3}, \end{eqnarray*} where $D$ is the pseudospectral differentiation matrix, (see \cite{Boyd}, chapter 6 and \cite{Canutohqz}, chapter 2), ${\rm diag}(V)$ is the diagonal matrix with elements $V_{j}=V(x_{j}), x_{j}=-l+jh, j=0,\ldots,m-1$, $I$ is the $m\times m$ identity matrix and the dot in the nonlinearity $N$ stands for the Hadamard product from the approximation $U_{h}\in \mathbb{R}^{m}$ to the exact values of the profile at the grid points $x_{j}$. \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|} \hline\hline \multicolumn{2}{|c|} {$V(x)={\rm sech}^{2}(x)$} & \multicolumn{2}{|c|}{$V(x)=-6({\rm sech}^{2}(x-1)+{\rm sech}^{2}(x+1))$}\\ \multicolumn{2}{|c|} {$\mu=1.3$} & \multicolumn{2}{|c|}{$\mu=1.43$}\\ \hline eigs $S$&eigs $(F^{\prime}(u^{*}))$&eigs $S$&eigs $(F^{\prime}(u^{*}))$\\\hline 2.9999E+00&7.0640E-01&8.0032E+00&8.0032E+00\\ 7.0640E-01&3.2731E-01&-5.6760E+00&-5.6760E+00\\ 3.2731E-01&1.9060E-01&2.9999E+00&-1.5841E+00\\ 1.9060E-01&1.2518E-01&-1.5841E+00&1.1350E+00\\ 1.2518E-01&8.8644E-02&1.1350E+00&-9.7207E-01\\ 8.8644E-02&6.6133E-02&-9.7207E-01&-5.7730E-01\\ \hline\hline \end{tabular} \end{center} \caption{Ground state generation for (\ref{doub_well11}) with $V(x)={\rm sech}^{2}(x), \mu=1.3$ and $V(x)=-6({\rm sech}^{2}(x-1)+{\rm sech}^{2}(x+1)), \mu=1.43$. Six largest magnitude eigenvalues of the approximate iteration matrix (\ref{m2b}) and of the Jacobian (\ref{m6c}). Both are evaluated at the last computed iterate $U_{f}$ of the Petviashvili method for (\ref{doub_well12}) and AITEM for (\ref{doub_well12}), respectively.}\label{tav1} \end{table} { The ground state generation of (\ref{doub_well11}) is illustrated with two potentials (see \cite{lakobay,yang2} and references therein for applications). The first one is $V(x)={\rm sech}^{2}(x)$.} For $\mu=1.3$ and a Gaussian profile as initial iteration, the Petviashvili method has been run. Table \ref{tav1} (first column) shows the six largest magnitude eigenvalues of the approximated iteration matrix (\ref{m2b}) at $u^{*}=U_{f}$, where $U_{f}$ is the last computed iterate (an analytical expression for the ground state profile is not known). The dominant eigenvalue $p=3$ (corresponding to the degree of homogeneity for this case) is observed, with the rest below one. The effect of the method is observed in the second column of Table \ref{tav1}, that displays the dominant eigenvalues of the Jacobian (\ref{m6c}). The magnitude of the eigenvalues is less than one, guaranteeing the convergence of the method, which is illustrated in Figure \ref{fexample221} (a). This shows the logarithm of the residual error (\ref{res}) as function of the number of iterations. In approximately $25$ iterations, a residual error of about $1.5\times 10^{-12}$ is obtained. The ground state profile is shown in Figure \ref{fexample221} (b). {The convergence of the stabilizing factor to one has also been checked, with a final discrepancy, in $25$ iterations, of about $3\times 10^{-14}$.} \begin{figure}[htbp] \centering \subfigure[]{ \includegraphics[width=6.6cm]{f223r.pdf} } \subfigure[]{ \includegraphics[width=6.6cm]{f221r.pdf} } \caption{Convergence results of the Petviashvili method for (\ref{doub_well11}) and $V(x)={\rm sech}^{2}(x)$: (a) Logarithm of the residual errors vs number of iterations. (b) Approximate profile.} \label{fexample221} \end{figure} { As a second example, we consider (\ref{doub_well11}) with a double-well potential $V(x)=-6({\rm sech}^{2}(x-1)+{\rm sech}^{2}(x+1))$. As indicated in \cite{lakobay}, the Petviashvili method fails in the search for an anti-symmetric solution of (\ref{doub_well12}) (see the profile in Figure \ref{fexample222}, obtained for $\mu=1.43$ with the AITEM method, \cite{yangl1}). This case of divergence can be justified using the previous results.} For $\mu=1.43$, Table \ref{tav1} (third column) shows the six largest magnitude eigenvalues of the iteration matrix evaluated at the profile obtained with the AITEM method. Besides the eigenvalue $\lambda=3$, associated to the degree of homogeneity of the nonlinear part, Table \ref{tav1} reveals the existence of other eigenvalues with magnitude above one. This divergence is also confirmed by the eigenvalues of the Jacobian (\ref{m6c}), shown in the fourth column of Table \ref{tav1}. \begin{figure}[htbp] \centering \includegraphics[width=0.7\textwidth]{nls_1r.pdf} \caption{Antisymmetric solution of (\ref{doub_well12}): numerical profile obtained with AITEM, $\mu=1.43$.} \label{fexample222} \end{figure} \subsection{Systems with symmetries. Orbital convergence} \label{sec33} In many situations, the system (\ref{m1}) admits symmetries, \cite{olver,marsdenr}. This means that there is an ${\nu}$-parameter group of transformations ($\nu\geq 1$) \begin{eqnarray} \mathcal{G}=\{G_{\alpha}:\mathbb{R}^{m}\rightarrow\mathbb{R}^{m}, \alpha=(\alpha_{1},\ldots,\alpha_{\nu})\in\mathbb{R}^{\nu}\},\label{m11} \end{eqnarray} with the property of transforming solutions of (\ref{m1}) into other solutions: \begin{eqnarray} Lu^{*}=N(u^{*})\Rightarrow L(G_{\alpha}u^{*})=N(G_{\alpha}u^{*}),\quad \alpha\in\mathbb{R}^{\nu}.\label{m12} \end{eqnarray} For simplicity, we assume that the transformations $G_{\alpha}$ in (\ref{m11}) are smooth and $\mathcal{G}$ is Abelian. {(In traveling wave generation, typical examples are, as mentioned before, translations and phase rotations, see Section \ref{sec34}.)}. The existence of a symmetry group for (\ref{m1}) has several consequences. We emphasize two of them: \begin{itemize} \item The group (\ref{m11}) defines orbits of solutions of (\ref{m1}): \begin{eqnarray*} \mathcal{G}(u^{*})=\{G_{\alpha}u^{*}:\alpha\in\mathbb{R}^{\nu}\}. \end{eqnarray*} The space of solutions of (\ref{m1}) is partitioned into these orbits, in such a way that assumption (H1) in Theorem \ref{theorem1} does not hold in this case: a fixed point $u^{*}$ cannot be isolated and the concept of convergence for the iterative methods must be redefined. Under these conditions, it is said that the iteration (\ref{m4}) is orbitally convergent to $u^{*}$ if $u_{n}$ converges to $G_{\alpha}u^{*}$ for some $\alpha\in \mathbb{R}^{\nu}$. \item The pencil $A(\lambda)$ admits $\lambda=1$ as eigenvalue, since differentiation with respect to $\alpha$ in (\ref{m12}) implies \begin{eqnarray*} \left(L-N^{\prime}(u^{*})\right)\frac{\partial}{\partial \alpha_{j}}\big|_{\alpha=0}G_{\alpha}(u^{*})=0,\quad j=1,\ldots, \nu, \end{eqnarray*} and the infinitesimal generators of the group, \cite{olver} \begin{eqnarray} u\mapsto v_{j}(u)=\frac{\partial}{\partial \alpha_{j}}\big|_{\alpha=0}G_{\alpha}(u),\quad j=1,\ldots, \nu,\label{igg} \end{eqnarray} evaluated at $u=u^{*}$, are associated eigenvectors. \end{itemize} {The convergence result in Theorem \ref{theorem1} can be adapted to this case as follows. First, hypothesis (H1) is substituted by \begin{itemize} \item[(H1)'] $\mathcal{G}$ is a symmetry group of (\ref{m1}) with $dim Ker A(1)=\nu$. \end{itemize} As far as the spectrum of $A(\lambda)$ is concerned, we still assume (i) and (ii) of Theorem \ref{theorem1} while the third condition is now \begin{itemize} \item[(iii)] If $|\lambda|=1\Rightarrow\left\{\begin{matrix}\lambda \quad \mbox {is semisimple}\\\mbox{if} \quad \lambda\neq 1\Rightarrow u_{0} \quad \mbox {does not have } \\ \mbox {component in}\quad Ker A(\lambda)\end{matrix}\right.$ \end{itemize} In this case the convergence is orbital, in the sense above defined. Note that now the errors $e_{n}, n=0,1,\ldots,$ can be decomposed in the form (cf. (\ref{m8})) \begin{eqnarray*} e_{n}=\alpha_{n}u^{*}+\sum_{k=1}^{\nu}\beta_{n,k}v_{k}(u^{*})+z_{n}, \quad \alpha_{n},\beta_{n,k}\in \mathbb{R}, k=1,\ldots,\nu,\quad z_{n}\in V, \end{eqnarray*} where $\{v_{k}(u^{*})\}_{k=1}^{\nu}$ is the basis (\ref{igg}) of $Ker A(1)$ and now $V$ is the (unique) supplementary $(m-\nu-1)$ dimensional space of $span(u^{*})+Ker A(1)$ with $S(V)\subset V$. Now, the sequences $\alpha_{n}, \beta_{n,k}, =1,\ldots,\nu$ and $z_{n}$ satisfy \begin{eqnarray} \alpha_{n+1}&=&\alpha_{n}(p+q)+\sum_{k=1}^{\nu}\beta_{n,k}\left(\nabla s(u^{\ast})\right)v_{k}(u^{*})+\left(\nabla s(u^{\ast})\right)z_{n},\label{mb1}\\ z_{n+1}&=&Sz_{n},\label{mb2}\\ \beta_{n+1,k}&=&\beta_{n,k}, \quad k=1,\ldots,\nu.\label{mb3} \end{eqnarray} Note, on the other hand, that (P1) and (\ref{m12}) imply \begin{eqnarray} s(G_{\alpha}(u^{*}))=1,\quad \alpha=(\alpha_{1},\ldots,\alpha_{\nu})\in\mathbb{R}^{\nu}.\label{ese} \end{eqnarray} Differentiating (\ref{ese}) with respect to each $\alpha_{j}, j=1,\ldots,\nu$ and evaluating at $\alpha=0$ we have \begin{eqnarray*} \left(\nabla s(u^{\ast})\right)v_{j}=0,\quad j=1,\ldots,\nu, \end{eqnarray*} and (\ref{mb1})-(\ref{mb3}) can be written as \begin{eqnarray*} \alpha_{n+1}&=&\alpha_{n}(p+q)+\left(\nabla s(u^{\ast})\right)z_{n},\\ z_{n+1}&=&Sz_{n},\\ \beta_{n+1,k}&=&\beta_{n,k}, \quad k=1,\ldots,\nu\\ &&\Rightarrow \beta_{n,k}=\beta_{0,k},\quad k=1,\ldots,\nu, \quad n\geq 0. \end{eqnarray*} Consequently, \begin{eqnarray*} G_{(\beta_{0,1},\ldots,\beta_{0,\nu})}(u^{*})+\alpha_{n}x^{*}+z_{n} \end{eqnarray*} differs from \begin{eqnarray*} u_{n}=u^{*}+e_{n}=u^{*}+\sum_{k=1}^{\nu}\beta_{0,k}v_{k}(u^{*})+z_{n}, \end{eqnarray*} in $O(||e_{n}||^{2})$ terms. Under the above mentioned hypotheses, the convergence is to the element $G_{(\beta_{0,1},\ldots,\beta_{0,\nu})}(u^{*})$ of the orbit of $u^{*}$, determined by the component of the initial iteration in Ker$(I-S)$.As in Theorem \ref{theorem1}, the fastest rate of convergence occurs when $q=-p$. } \subsection{Some examples} \label{sec34} This case is illustrated by the generation of solitary wave solutions of nonlinear Schr\"{o}dinger equations of the form (see e.~g. \cite{sulems} and references therein) \begin{eqnarray} iu_{t}+u_{xx}+|u|^{2\sigma}u=0,\quad -\infty<x<\infty,\quad t>0,\label{nlse1} \end{eqnarray} where $\sigma>0$. The symmetry group for (\ref{nlse1}) consists of gauge transformations and translations \begin{eqnarray} G_{{\theta_{0},x_{0}}}(u(x))=e^{i\theta_{0}}u(x+x_{0}),\quad \theta_{0},x_{0}\in \mathbb{R}.\label{nlse2} \end{eqnarray} Solitary wave solutions of (\ref{nlse1}) can be obtained from profiles {$U(x)$ satisfying} \begin{eqnarray} U^{\prime\prime}+|U|^{2\sigma}u-\lambda_{1}U-i\lambda_{2}U^{\prime}=0,\label{nlse3} \end{eqnarray} for some real parameters $\lambda_{1},\lambda_{2}$. This leads to the explicit formulas \begin{eqnarray} U(x)&=&\rho(x)e^{i\theta(x)}\label{nlse4}\\ \rho(x)&=&(a(\sigma+1))^{1/2\sigma}\left({\rm sech}(\sigma\sqrt{a}x)\right)^{1/\sigma},\quad a=\lambda_{1}-(\lambda_{2}^2)/4,\label{nlse4b}\\ \theta(x)&=&\frac{\lambda_{2}}{2}x.\label{nlse4c} \end{eqnarray} Due to the symmetry group (\ref{nlse2}), the two-parameter orbit of the solution given by (\ref{nlse4})-(\ref{nlse4c}) is of the form \begin{eqnarray} \mathcal{G}(\rho,\theta)=\{\varphi=\rho(x-x_{0})e^{i\theta(x-x_{0})+i\theta_{0}}: x_{0},\theta_{0}\in \mathbb{R}\}.\label{nlse41} \end{eqnarray} The four-parameter family of solitary wave solutions of (\ref{nlse1}) is finally of the form \begin{eqnarray*} \psi(x,t,a,c,x_{0},\theta_{0})=G_{(t\lambda_{1},t\lambda_{2})}(\varphi)=\rho(x-ctx_{0})e^{i\theta(x-ct-x_{0})+i\theta_{0}+i(a+(c^{2}/4))t},\label{nlse5} \end{eqnarray*} (where $c=\lambda_{1}$). As far as the discretization is concerned, the corresponding Fourier collocation approximation of (\ref{nlse3}) \begin{eqnarray*} D^{2}U_{h}+|(U_{h})|.^{2\sigma}.U_{h}-\lambda_{1}U_{h}-i\lambda_{2}DU_{h}=0, \end{eqnarray*} (the dot stands for the Hadamard product) inherites the symmetry group infinitesimally generated by (see (\ref{igg})) \begin{eqnarray*} U_{h}\mapsto v_{1}(U_{h})=iU_{h},\quad U_{h}\mapsto v_{2}(U_{h})=DU_{h}. \end{eqnarray*} They are associated, respectively, to phase rotations and spatial translations. \begin{table} \begin{center} \begin{tabular}{|c|c|} \hline\hline $\sigma=1$ & $\sigma=2$\\ \hline 2.9999E+00&4.9999E+00\\ 9.9999E-01&9.9999E-01\\ 9.9999E-01&9.9999E-01\\ 4.9999E-01&4.2857E-01\\ 3.3333E-01&2.3810E-01\\ 2.9999E-01&1.9999E-01\\ \hline\hline \end{tabular} \end{center} \caption{Solitary wave generation of (\ref{nlse1}). Six largest magnitude eigenvalues of the approximate iteration matrix (\ref{m2b}) evaluated at the exact profile (\ref{nlse4}) with $\lambda_{1}=\lambda_{2}=1, x_{0}=\theta_{0}=0$ for $\sigma=1$ (first column) and $\sigma=2$ (second column).}\label{tav2} \end{table} Local convergence of the Petviashvili method is first checked by Table \ref{tav2}. This shows, for $\sigma=1,2$, the first largest magnitude eigenvalues of the iteration matrix $S$ at the exact profile (\ref{nlse4})-(\ref{nlse4c}) with parameters $\lambda_{1}=\lambda_{2}=1$, (where $x_{0}=\theta_{0}=0$). The results guarantee the satisfaction of the conditions for local convergence. (Note that, in this case, since the symmetry group is two dimensional, the eigenvalue $\lambda=1$ has geometric multiplicity equals two.) { The following results illustrate the orbital convergence. An exact profile of the form (\ref{nlse4})-(\ref{nlse4c}) with $\lambda_{1}=\lambda_{2}=1, x_{0}=\theta_{0}=0$, denoted by $U_{\rm exact}$, is perturbed in the form \begin{eqnarray} U_{0}=U_{\rm exact}+\epsilon_{1} i U_{\rm exact}+\epsilon_{2} D U_{\rm exact},\label{m36} \end{eqnarray} (with small parameters $\epsilon_{1},\epsilon_{2}$). Now the Petviashvili method is run with (\ref{m36}) as initial iteration. The results are illustrated by two experiments. Figure \ref{fnls1}(a) (resp. Figure \ref{fnls1}(b)) compares the real part (resp. the modulus) of the exact profile $U_{\rm exact}$ with that of the last computed iterate, denoted by $U_{f}$ and obtained with a residual error below $10^{-13}$ and for $\epsilon_{1}=0.2, \epsilon_{2}=0$. } \begin{figure}[htbp] \centering \subfigure[]{ \includegraphics[width=6.6cm]{nlsr1.pdf} } \subfigure[]{ \includegraphics[width=6.6cm]{nlsr2.pdf} } \subfigure[]{ \includegraphics[width=6.6cm]{nlsr3.pdf} } \caption{{Solitary wave generation of (\ref{nlse1}). The parameters are $\lambda_{1}=\lambda_{2}=1, x_{0}=\theta_{0}=0, \epsilon_{1}=0.2, \epsilon_{2}=0$. (a) Real part of the last computed iterate $U_{f}$ obtained by the Petviashvili scheme (solid line) and of the exact profile $U_{\rm exact}$ (dashed line). (b) Modulus of the last computed iterate $U_{f}$ obtained by the Petviashvili scheme (solid line) and of the exact profile $U_{\rm exact}$ (dashed line). (c) Fitting line to the phase of the computed profile $U_{f}$.} \label{fnls1}} \end{figure} {While moduli are practically indistinguishable, Figure \ref{fnls1}(a) reveals a phase displacement of the computed profile with respect to the exact one. The phase of $U_{f}$ (computed as $Im\left(log\left(U_{f}/|U_{f}|\right)\right)$, modulo $2\pi$) has been calculated. The resulting data are fitted to a line $y=mx+n$, see Figure \ref{fnls1}(c). The computed slope is $m=4.9934\times 10^{-1}$ (approximating the corresponding value $\lambda_{2}/2$ of (\ref{nlse4})-(\ref{nlse4c}) for this case) while $n=1.2764\times 10^{1}$, which modulo $2\pi$ is $1.9751\times 10^{-1}$, an approximation to the value of $\epsilon_{1}$. These results suggest that the computed profile is closer to the element of the orbit (\ref{nlse41}) of $U_{\rm exact}$ with new phase $\theta_{0}+\epsilon_{1}=\epsilon_{1}$ and the same translational parameter $x_{0}=0$. } \begin{figure}[htbp] \centering \subfigure[]{ \includegraphics[width=6.6cm]{nlsr4.pdf} } \subfigure[]{ \includegraphics[width=6.6cm]{nlsr5.pdf} } \subfigure[]{ \includegraphics[width=6.6cm]{nlsr6.pdf} } \caption{{Solitary wave generation of (\ref{nlse1}). The parameters are $\lambda_{1}=\lambda_{2}=1, x_{0}=\theta_{0}=0, \epsilon_{1}=0.2, \epsilon_{2}=0.2$. (a) Real part of the last computed iterate $U_{f}$ obtained by the Petviashvili scheme (solid line) and of the exact profile $U_{\rm exact}$ (dashed line). (b) Modulus of the last computed iterate $U_{f}$ obtained by the Petviashvili scheme (solid line) and of the exact profile $U_{\rm exact}$ (dashed line). (c) Logarithm of the residual error against number of iterations.}} \label{fnls2} \end{figure} {A second experiment is performed with the values $\epsilon_{1}=0.2, \epsilon_{2}=0.2$. Figure \ref{fnls2}(a) (resp. Figure \ref{fnls2}(b)) compares the real part (resp. the modulus) of the exact profile $U_{\rm exact}$ with that of the last computed iterate $U_{f}$, after $35$ iterations and with a residual error (\ref{res}) below $10^{-11}$ (see Figure \ref{fnls2}(c)). The error in the stabilizing factor is of the order of the machine precision (below $10^{-15}$). In this case, besides the phase shift, also the modulus is affected by a displacement (cf. Figure \ref{fnls1}(b)). The corresponding fitting line to the phase data of the computed profile $U_{f}$, $y=mx+n$ has a slope $m=4.9933\times 10^{-1}$ while $n=1.2852\times 10^{1}$, which modulo $2\pi$ is approximately $2.8553\times 10^{-1}$. This last value approximates $\epsilon_{1}+(\lambda_{2}/2)\epsilon_{2}=0.3$, suggesting that the computed profile is close to the exact one of the form (\ref{nlse4}), (\ref{nlse41}) with group parameters $\theta_{0}+\epsilon_{1}, x_{0}+\epsilon_{2}$. } \section*{Acknowledgements} This research has been supported by projects MTM2010-19510/MTM (MCIN), MTM2012-30860(MECC) and VA118A12-1 (JCYL). The authors want to thank the reviewers for their fruitful suggestions and comments.
\section{Introduction} In the modern digital period, we are facing a rapid growth of available datasets in science and technology. In most computing tasks (e.g. storing and searching in such datasets), large datasets are a burden and require more computation. However, for learning tasks the situation is radically different. A simple observation is that more data can never hinder you from performing a task. If you have more data than you need, just ignore it! A basic question is how to learn from ``big data''. The statistical learning literature classically studies questions like ``how much data is needed to perform a learning task?" or ``how does accuracy improve as the amount of data grows?" etc. In the modern, ``data revolution era'', it is often the case that the amount of data available far exceeds the information theoretic requirements. We can wonder whether this, seemingly redundant data, can be used for other purposes. An intriguing question in this vein, studied recently by several researchers (\citep{DecaturGoRo98,Servedio00,ShalevShTr12,BerthetRi13,ChandrasekaranJo13}), is the following \begin{quotation}\label{quest} {\em Question 1:} Are there any learning tasks in which more data, {\em beyond the information theoretic barrier}, can {\em provably} be leveraged to speed up computation time? \end{quotation} The main contributions of this work are: \begin{itemize} \item Conditioning on the hardness of refuting random $\mathrm{3CNF}$ formulas, we give the first example of a {\em natural supervised learning problem} for which the answer to Question 1 is positive. \item To prove this, we present a novel technique to establish computational-statistical tradeoffs in supervised learning problems. To the best of our knowledge, this is the first such a result that is not based on cryptographic primitives. \end{itemize} Additional contributions are non trivial efficient algorithms for learning halfspaces over $2$-sparse and $3$-sparse vectors using $\tilde{O}\left(\frac{n}{\epsilon^2}\right)$ and $\tilde{O}\left(\frac{n^2}{\epsilon^2}\right)$ examples respectively. The natural learning problem we consider is the task of learning the class of {\em halfspaces over $k$-sparse vectors}. Here, the instance space is the {\em space of $k$-sparse vectors}, \[ C_{n,k}=\{x\in \{-1,1,0\}^n\mid |\{i\mid x_i\ne 0\}|\le k\} ~, \] and the hypothesis class is {\em halfspaces over $k$-sparse vectors,} namely \[ {\mathcal H}_{n,k} = \{h_{w,b} :C_{n,k}\to\{\pm 1\} \mid h_{w,b}(x)= \textrm{sign}(\inner{w,x} + b),w\in \mathbb R^n,b\in\mathbb R\} ~, \] where $\inner{\cdot,\cdot}$ denotes the standard inner product in $\mathbb R^n$. We consider the standard setting of agnostic PAC learning, which models the realistic scenario where the labels are not necessarily fully determined by some hypothesis from ${\mathcal H}_{n,k}$. Note that in the realizable case, i.e. when some hypothesis from ${\mathcal H}_{n,k}$ has zero error, the problem of learning halfspaces is easy even over $\mathbb{R}^n$. In addition, we allow improper learning (a.k.a. representation independent learning), namely, the learning algorithm is not restricted to output a hypothesis from ${\mathcal H}_{n,k}$, but only should output a hypothesis whose error is not much larger than the error of the best hypothesis in ${\mathcal H}_{n,k}$. This gives the learner a lot of flexibility in choosing an appropriate representation of the problem. This additional freedom to the learner makes it much harder to prove lower bounds in this model. Concretely, it is not clear how to use standard reductions from NP hard problems in order to establish lower bounds for improper learning (moreover, \cite{ApplebaumBaXi08} give evidence that such simple reductions do not exist). The classes ${\mathcal H}_{n,k}$ and similar classes have been studied by several authors (e.g. \cite{LongSe13}). They naturally arise in learning scenarios in which the set of all possible features is very large, but each example has only a small number of active features. For example: \begin{itemize} \item {\em Predicting an advertisement based on a search query:} Here, the possible features of each instance are all English words, whereas the active features are only the set of words given in the query. \item {\em Learning Preferences~\citep{HazanKaSh12}:} Here, we have $n$ players. A {\em ranking} of the players is a permutation $\sigma:[n]\to [n]$ (think of $\sigma(i)$ as the rank of the $i$'th player). Each ranking induces a {\em preference} $h_{\sigma}$ over the ordered pairs, such that $h_{\sigma}(i,j)=1$ iff $i$ is ranked higher that $j$. Namely, \[ h_{\sigma}(i,j)= \begin{cases} 1 & \sigma (i)>\sigma (j)\\ -1 & \sigma (i)<\sigma (j) \end{cases} \] The objective here is to learn the class, ${\mathcal P}_n$, of all possible preferences. The problem of learning preferences is related to the problem of learning ${\mathcal H}_{n,2}$: if we associate each pair $(i,j)$ with the vector in $C_{n,2}$ whose $i$'th coordinate is $1$ and whose $j$'th coordinate is $-1$, it is not hard to see that ${\mathcal P}_{n}\subset {\mathcal H}_{n,2}$: for every $\sigma$, $h_{\sigma}=h_{w,0}$ for the vector $w\in \mathbb R^n$, given by $w_i=\sigma(i)$. Therefore, every upper bound for ${\mathcal H}_{n,2}$ implies an upper bound for ${\mathcal P}_{n}$, while every lower bound for ${\mathcal P}_{n}$ implies a lower bound for ${\mathcal H}_{n,2}$. Since $\VC({\mathcal P}_n)=n$ and $\VC({\mathcal H}_{n,2})=n+1$, the information theoretic barrier to learn these classes is $\Theta\left(\frac{n}{\epsilon^2}\right)$. In \cite{HazanKaSh12} it was shown that ${\mathcal P}_n$ can be {\em efficiently} learnt using $O\left(\frac{n\log^3(n)}{\epsilon^2}\right)$ examples. In section \ref{sec:upper_bounds}, we extend this result to ${\mathcal H}_{n,2}$. \end{itemize} We will show a positive answer to Question 1 for the class ${\mathcal H}_{n,3}$. To do so, we show\footnote{In fact, similar results hold for every constant $k \ge 3$. Indeed, since ${\mathcal H}_{n,3} \subset {\mathcal H}_{n,k}$ for every $k \ge 3$, it is trivial that item $3$ below holds for every $k \ge 3$. The upper bound given in item $1$ holds for every $k$. For item 2, it is not hard to show that ${\mathcal H}_{n,k}$ can be learnt using a sample of $\Omega \left(\frac{n^k}{\epsilon^2}\right)$ examples by a naive improper learning algorithm, similar to the algorithm we describe in this section for $k=3$.} the following: \begin{enumerate} \item Ignoring computational issues, it is possible to learn the class ${\mathcal H}_{n,3}$ using $O\left(\frac{n}{\epsilon^2}\right)$ examples. \item It is also possible to \emph{efficiently} learn ${\mathcal H}_{n,3}$ if we are provided with a larger training set (of size $\tilde{\Omega}\left(\frac{n^2}{\epsilon^2}\right)$). This is formalized in Theorem \ref{thm:main}. \item It is impossible to \emph{efficiently} learn ${\mathcal H}_{n,3}$, if we are only provided with a training set of size $O\left(\frac{n}{\epsilon^2}\right)$ under Feige's assumption regarding the hardness of refuting random $\mathrm{3CNF}$ formulas \citep{Feige02}. Furthermore, for every $\alpha \in [0,0.5)$, it is impossible to learn efficiently with a training set of size $O\left(\frac{n^{1+\alpha}}{\epsilon^2}\right)$ under a stronger hardness assumption. This is formalized in Theorem \ref{thm:learning_H_new}. \end{enumerate} A graphical illustration of our main results is given below: \begin{center} \begin{tikzpicture}[scale=2.3] \draw[blue,very thick] (0.20,0.70) --(0.25,0.70) --(0.30,0.70) --(0.35,0.70) --(0.40,0.70) --(0.45,0.70) --(0.50,0.70) --(0.55,0.70) --(0.60,0.70) --(0.65,0.70) --(0.70,0.70) --(0.75,0.70) --(0.80,0.70) --(0.85,0.69) --(0.90,0.69) --(0.95,0.68) --(1.00,0.67) --(1.05,0.65) --(1.10,0.62) --(1.15,0.57) --(1.20,0.51) --(1.25,0.44) --(1.30,0.35) --(1.35,0.26) --(1.40,0.19) --(1.45,0.13) --(1.50,0.08) --(1.55,0.05) --(1.60,0.03) --(1.65,0.02) --(1.70,0.01) --(1.75,0.01) --(1.80,0.00) ; \draw[|->,very thick,black] (0.4,0.8) -- (0.4,0.7); \draw[|->,very thick,black] (1,0.5) -- (1,0.6); \draw[|->,very thick,black] (1.6,0.2) -- (1.6,0.1); \draw[->,very thick,black] (0,-0.1) -- (0,1) node[above=5pt] {runtime}; \draw[-,thick,black] (0.06,0.8) -- (-0.06,0.8) node[left] {$2^{O(n)}$}; \draw[-,thick,black] (0.06,0.5) -- (-0.06,0.5) node[left] {$>\poly(n)$}; \draw[-,thick,black] (0.06,0.2) -- (-0.06,0.2) node[left] {$n^{O(1)}$}; \draw[->,very thick,black] (0,-0.1) -- (1.9,-0.1) node[right] {examples}; \draw[-,thick,black] (1.6,-0.04) -- (1.6,-0.16) node[below] {$n^2$}; \draw[-,thick,black] (1,-0.04) -- (1,-0.16) node[below] {$n^{1.5}$}; \draw[-,thick,black] (0.4,-0.04) -- (0.4,-0.16) node[below] {$n$}; \end{tikzpicture} \end{center} \vskip -0.4cm The proof of item 1 above is easy -- simply note that $H_{n,3}$ has VC dimension $n+1$. Item 2 is proved in section \ref{sec:upper_bounds}, relying on the results of \cite{HazanKaSh12}. We note, however, that a weaker result, that still suffices for answering Question 1 in the affirmative, can be proven using a naive improper learning algorithm. In particular, we show below how to learn ${\mathcal H}_{n,3}$ efficiently with a sample of $\Omega\left(\frac{n^3}{\epsilon^2}\right)$ examples. The idea is to replace the class ${\mathcal H}_{n,3}$ with the class $\{\pm 1\}^{C_{n,3}}$ containing \emph{all} functions from $C_{n,3}$ to $\{\pm 1\}$. Clearly, this class contains $H_{n,3}$. In addition, we can efficiently find a function $f$ that minimizes the empirical training error over a training set $S$ as follows: For every $x \in C_{n,k}$, if $x$ does not appear at all in the training set we will set $f(x)$ arbitrarily to $1$. Otherwise, we will set $f(x)$ to be the majority of the labels in the training set that correspond to $x$. Finally, note that the VC dimension of $\{\pm 1\}^{C_{n,3}}$ is smaller than $n^3$ (since $|C_{n,3}| < n^3$). Hence, standard generalization results (e.g. \cite{Vapnik95}) implies that a training set size of $\Omega\left(\frac{n^3}{\epsilon^2}\right)$ suffices for learning this class. Item 3 is shown in section \ref{sec:main} by presenting a novel technique for establishing statistical-computational tradeoffs. {\bf The class ${\mathcal H}_{n,2}$.} Our main result gives a positive answer to Question 1 for the task of improperly learning ${\mathcal H}_{n,k}$ for $k\ge 3$. A natural question is what happens for $k=2$ and $k=1$. Since $\VC({\mathcal H}_{n,1})= \VC({\mathcal H}_{n,2})=n+1$, the information theoretic barrier for learning these classes is $\Theta\left(\frac{n}{\epsilon^2}\right)$. In section \ref{sec:upper_bounds}, we prove that ${\mathcal H}_{n,2}$ (and, consequently, ${\mathcal H}_{n,1}\subset {\mathcal H}_{n,2}$) can be learnt using $O\left(\frac{n\log^3(n)}{\epsilon^2}\right)$ examples, indicating that significant computational-statistical tradeoffs start to manifest themselves only for $k\ge 3$. \subsection{Previous approaches, difficulties, and our techniques} \citep{DecaturGoRo98} and \citep{Servedio00} gave positive answers to Question 1 in the realizable PAC learning model. Under cryptographic assumptions, they showed that there exist binary learning problems, in which more data can provably be used to speed up training time. \citep{ShalevShTr12} showed a similar result for the agnostic PAC learning model. In all of these papers, the main idea is to construct a hypothesis class based on a one-way function. However, the constructed classes are of a very synthetic nature, and are of almost no practical interest. This is mainly due to the construction technique which is based on one way functions. In this work, instead of using cryptographic assumptions, we rely on the hardness of refuting random $\mathrm{3CNF}$ formulas. The simplicity and flexibility of $\mathrm{3CNF}$ formulas enable us to derive lower bounds for natural classes such as halfspaces. Recently, \citep{BerthetRi13} gave a positive answer to Question 1 in the context of unsupervised learning. Concretely, they studied the problem of sparse PCA, namely, finding a \emph{sparse} vector that maximizes the variance of an unsupervised data. Conditioning on the hardness of the planted clique problem, they gave a positive answer to Question 1 for sparse PCA. Our work, as well as the previous work of \cite{DecaturGoRo98,Servedio00,ShalevShTr12}, studies Question 1 in the supervised learning setup. We emphasize that unsupervised learning problems are radically different than supervised learning problems in the context of deriving lower bounds. The main reason for the difference is that in supervised learning problems, the learner is allowed to employ improper learning, which gives it a lot of power in choosing an adequate representation of the data. For example, the upper bound we have derived for the class of sparse halfspaces switched from representing hypotheses as halfspaces to representation of hypotheses as tables over $C_{n,3}$, which made the learning problem easy from the computational perspective. The crux of the difficulty in constructing lower bounds is due to this freedom of the learner in choosing a convenient representation. This difficulty does not arise in the problem of sparse PCA detection, since there the learner must output a good sparse vector. Therefore, it is not clear whether the approach given in \citep{BerthetRi13} can be used to establish computational-statistical gaps in supervised learning problems. \section{Background and notation} For hypothesis class ${\mathcal H}\subset \{\pm 1\}^{X}$ and a set $Y\subset X$, we define the {\em restriction of ${\mathcal H}$ to $Y$} by ${\mathcal H}|_{Y}=\{h|_Y\mid h\in {\mathcal H}\}$. We denote by $J=J_n$ the all-ones $n\times n$ matrix. We denote the $j$'th vector in the standard basis of $\mathbb R^n$ by $e_j$. \subsection{Learning Algorithms} For $h:C_{n,3}\to\{\pm 1\}$ and a distribution ${\mathcal D}$ on $C_{n,3}\times\{\pm 1\}$ we denote the {\em error of $h$ w.r.t. ${\mathcal D}$} by $\Err_{\mathcal D}(h)=\Pr_{(x,y)\sim {\mathcal D}}\left(h(x)\ne y\right)$. For ${\mathcal H}\subset \{\pm 1\}^{C_{n,3}}$ we denote the {\em error of ${\mathcal H}$ w.r.t. ${\mathcal D}$} by $\Err_{{\mathcal D}}({\mathcal H})=\min_{h\in{\mathcal H}}\Err_{{\mathcal D}}(h)$. For a sample $S\in \left(C_{n,3}\times \{\pm 1\}\right)^m$ we denote by $\Err_{S}(h)$ (resp. $\Err_{S}({\mathcal H})$) the error of $h$ (resp. ${\mathcal H}$) w.r.t. the empirical distribution induces by the sample $S$. A {\em learning algorithm}, $L$, receives a sample $S\in \left(C_{n,3}\times \{\pm 1\}\right)^m$ and return a hypothesis $L(S):C_{n,3}\to \{\pm 1\}$. We say that $L$ {\em learns ${\mathcal H}_{n,3}$ using $m(n,\epsilon)$ examples} if,\footnote{For simplicity, we require the algorithm to succeed with probability of at least $9/10$. This can be easily amplified to probability of at least $1-\delta$, as in the usual definition of agnostic PAC learning, while increasing the sample complexity by a factor of $\log(1/\delta)$. } for every distribution ${\mathcal D}$ on $C_{n,3}\times \{\pm 1\}$ and a sample $S$ of more than $m(n,\epsilon)$ i.i.d. examples drawn from ${\mathcal D}$, \[ \Pr_{S}\left(\Err_{{\mathcal D}}(L(S))>\Err_{{\mathcal D}}({\mathcal H}_{3,n})+\epsilon\right)<\frac{1}{10} \] The algorithm $L$ is {\em efficient} if it runs in polynomial time in the sample size and returns a hypothesis that can be evaluated in polynomial time. \subsection{Refuting random $\mathrm{3SAT}$ formulas} We frequently view a boolean assignment to variables $x_1,\ldots,x_n$ as a vector in $\mathbb R^n$. It is convenient, therefore, to assume that boolean variables take values in $\{\pm 1\}$ and to denote negation by $``-"$ (instead of the usual $``\neg"$). An $n$-variables $\mathrm{3CNF}$ clause is a boolean formula of the form \[ C(x)=(-1)^{j_1}x_{i_1}\vee (-1)^{j_2}x_{i_2}\vee (-1)^{j_1}x_{i_3},\;\;x\in \{\pm 1\}^n \] An $n$-variables $\mathrm{3CNF}$ formula is a boolean formula of the form \[ \phi(x)=\wedge_{i=1}^m C_i(x) ~, \] where every $C_i$ is a $\mathrm{3CNF}$ clause. Define the {\em value}, $\val (\phi)$, of $\phi$ as the maximal fraction of clauses that can be simultaneously satisfied. If $\val(\phi)=1$, we say the $\phi$ is {\em satisfiable}. By $\mathrm{3CNF}_{n,m}$ we denote the set of $\mathrm{3CNF}$ formulas with $n$ variables and $m$ clauses. Refuting random $\mathrm{3CNF}$ formulas has been studied extensively (see e.g. a special issue of TCS \cite{DubiosMoSeZe}). It is known that for large enough $\Delta$ ($\Delta=6$ will suffice) a random formula in $\mathrm{3CNF}_{n,\Delta n}$ is not satisfiable with probability $1-o(1)$. Moreover, for every $0\le \epsilon<\frac{1}{4}$, and a large enough $\Delta=\Delta(\epsilon)$, the value of a random formula $\mathrm{3CNF}_{n,\Delta n}$ is $\le 1-\epsilon$ with probability $1-o(1)$. The problem of refuting random $\mathrm{3CNF}$ concerns efficient algorithms that provide a proof that a random $\mathrm{3CNF}$ is not satisfiable, or far from being satisfiable. This can be thought of as a game between an adversary and an algorithm. The adversary should produce a $\mathrm{3CNF}$-formula. It can either produce a satisfiable formula, or, produce a formula uniformly at random. The algorithm should identify whether the produced formula is random or satisfiable. Formally, let $\Delta:\mathbb N\to\mathbb N$ and $0\le \epsilon <\frac{1}{4}$. We say that an efficient algorithm, $A$, {\em $\epsilon$-refutes random $\mathrm{3CNF}$ with ratio $\Delta$} if its input is $\phi\in \mathrm{3CNF}_{n,n\Delta(n)}$, its output is either $\mathrm{``typical"}$ or $\mathrm{``exceptional"}$ and it satisfies: \begin{itemize} \item {\em Soundness:} If $\val(\phi)\ge 1-\epsilon$, then \[ \Pr_{\text{Rand. coins of }A}\left(A(\phi)=\mathrm{``exceptional"}\right) \ge \frac{3}{4} \] \item {\em Completeness:} For every $n$, \[ \Pr_{\text{Rand. coins of }A,\;\phi\sim \mathrm{Uni}(\mathrm{3CNF}_{n,n\Delta (n)})}\left(A(\phi)=\mathrm{``typical"}\right) \ge 1-o(1) \] \end{itemize} By a standard repetition argument, the probability of $\frac{3}{4}$ can be amplified to $1-2^{-n}$, while efficiency is preserved. Thus, given such an (amplified) algorithm, if $A(\phi)=``\mathrm{typical}"$, then with confidence of $1-2^{-n}$ we know that $\val(\phi)< 1-\epsilon$. Since for random $\phi\in \mathrm{3CNF}_{n,n \Delta (n)}$, $A(\phi)=``\mathrm{typical}"$ with probability $1-o(1)$, such an algorithm provides, for most $\mathrm{3CNF}$ formulas a proof that their value is less that $1-\epsilon$. Note that an algorithm that $\epsilon$-refutes random $\mathrm{3CNF}$ with ratio $\Delta$ also $\epsilon'$-refutes random $\mathrm{3CNF}$ with ratio $\Delta$ for every $0\le \epsilon'\le\epsilon$. Thus, the task of refuting random $\mathrm{3CNF}$'s gets easier as $\epsilon$ gets smaller. Most of the research concerns the case $\epsilon=0$. Here, it is not hard to see that the task is getting easier as $\Delta$ grows. The best known algorithm \citep{FeigeOf07} $0$-refutes random $\mathrm{3CNF}$ with ratio $\Delta (n)=\Omega(\sqrt{n})$. In \cite{Feige02} it was conjectured that for constant $\Delta$ no efficient algorithm can provide a proof that a random $\mathrm{3CNF}$ is not satisfiable: \begin{conjecture}[R3SAT hardness assumption -- \citep{Feige02}]\label{hyp:feige_basic} For every $\epsilon>0$ and for every large enough integer $\Delta>\Delta_0(\epsilon)$ there exists no efficient algorithm that $\epsilon$-refutes random $\mathrm{3CNF}$ formulas with ratio $\Delta$. \end{conjecture} In fact, for all we know, the following conjecture may be true for every $0 \le \mu \le 0.5$. \begin{conjecture}[$\mu$-R3SAT hardness assumption]\label{hyp:mu_conj} For every $\epsilon>0$ and for every integer $\Delta>\Delta_0(\epsilon)$ there exists no efficient algorithm that $\epsilon$-refutes random $\mathrm{3CNF}$ with ratio $\Delta\cdot n^\mu$. \end{conjecture} Note that Feige's conjecture is equivalent to the $0$-R3SAT hardness assumption. \section{Lower bounds for learning ${\mathcal H}_{n,3}$}\label{sec:main} \begin{theorem}[main]\label{thm:main} Let $0\le \mu\le 0.5$. If the $\mu$-R3SAT hardness assumption (conjecture \ref{hyp:mu_conj}) is true, then there exists no efficient learning algorithm that learns the class ${\mathcal H}_{n,3}$ using $O\left(\frac{n^{1+\mu}}{\epsilon^2}\right)$ examples. \end{theorem} In the proof of Theorem \ref{thm:main} we rely on the validity of a conjecture, similar to conjecture \ref{hyp:mu_conj} for $3$-variables majority formulas. Following an argument from \citep{Feige02} (Theorem~\ref{hyp:feige_maj}) the validity of the conjecture on which we rely for majority formulas follows the validity of conjecture \ref{hyp:mu_conj}. Define \[ \forall (x_1,x_2,x_3)\in \{\pm 1\}^3,\;\textrm{MAJ}(x_1,x_2,x_3):=\textrm{sign}(x_1+x_2+x_3) \] An $n$-variables $\textrm{3MAJ}$ clause is a boolean formula of the form \[ C(x)=\textrm{MAJ}((-1)^{j_1}x_{i_1}, (-1)^{j_2}x_{i_2},(-1)^{j_1}x_{i_3}),\;\;x\in \{\pm 1\}^n \] An $n$-variables $\textrm{3MAJ}$ formula is a boolean formula of the form \[ \phi(x)=\wedge_{i=1}^m C_i(x) \] where the $C_i$'s are $\textrm{3MAJ}$ clauses. By $\textrm{3MAJ}_{n,m}$ we denote the set of $\textrm{3MAJ}$ formulas with $n$ variables and $m$ clauses. \begin{theorem}[\citep{Feige02}]\label{hyp:feige_maj} Let $0\le \mu \le 0.5$. If the $\mu$-R3SAT hardness assumption is true, then for every $\epsilon>0$ and for every large enough integer $\Delta>\Delta_0(\epsilon)$ there exists no efficient algorithm with the following properties. \begin{itemize} \item Its input is $\phi\in \textrm{3MAJ}_{n,\Delta n^{1+\mu}}$, and its output is either $\mathrm{``typical"}$ or $\mathrm{``exceptional"}$. \item If $\val(\phi)\ge \frac{3}{4}-\epsilon$, then \[ \Pr_{\text{Rand. coins of }A}\left(A(\phi)=\mathrm{``exceptional"}\right) \ge \frac{3}{4} \] \item For every $n$, \[ \Pr_{\text{Rand. coins of }A,\;\phi\sim \mathrm{Uni}(\textrm{3MAJ}_{n,\Delta n^{1+\mu}})}\left(A(\phi)=\mathrm{``typical"}\right) \ge 1-o(1) \] \end{itemize} \end{theorem} Next, we prove Theorem \ref{thm:main}. In fact, we will prove a slightly stronger result. Namely, define the subclass ${\mathcal H}_{n,3}^d\subset {\mathcal H}_{n,3}$, of homogenous halfspaces with binary weights, given by ${\mathcal H}_{n,3}^d=\left\{h_{w,0}\mid w\in \{\pm 1\}^n\right\}$. As we show, under the $\mu$-R3SAT hardness assumption, it is impossible to efficiently learn this subclass using only $O\left(\frac{n^{1+\mu}}{\epsilon^2}\right)$ examples. {\em Proof idea:} We will reduce the task of refuting random $\mathrm{3MAJ}$ formulas with linear number of clauses to the task of (improperly) learning ${\mathcal H}^d_{n,3}$ with linear number of samples. The first step will be to construct a transformation that associates every $\mathrm{3MAJ}$ clause with two examples in $C_{n,3}\times \{\pm 1\}$, and every assignment with a hypothesis in ${\mathcal H}^d_{n,3}$. As we will show, the hypothesis corresponding to an assignment $\psi$ is correct on the two examples corresponding to a clause $C$ if and only if $\psi$ satisfies $C$. With that interpretation at hand, every $\mathrm{3MAJ}$ formula $\phi$ can be thought of as a distribution ${\mathcal D}_\phi$ on $C_{n,3}\times \{\pm 1\}$, which is the empirical distribution induced by $\psi$'s clauses. It holds furthermore that $\Err_{{\mathcal D}_\phi}({\mathcal H}^d_{n,3})=1-\val(\phi)$. Suppose now that we are given an efficient learning algorithm for ${\mathcal H}^d_{n,3}$, that uses $\kappa \frac{n}{\epsilon^2}$ examples, for some $\kappa>0$. To construct an efficient algorithm for refuting $\mathrm{3MAJ}$-formulas, we simply feed the learning algorithm with $\kappa \frac{n}{0.01^2}$ examples drawn from ${\mathcal D}_{\phi}$ and answer ``exceptional'' if the error of the hypothesis returned by the algorithm is small. If $\phi$ is (almost) satisfiable, the algorithm is guaranteed to return a hypothesis with a small error. On the other hand, if $\phi$ is far from being satisfiable, $\Err_{{\mathcal D}_\phi}({\mathcal H}^d_{n,3})$ is large. If the learning algorithm is proper, then it must return a hypothesis from ${\mathcal H}^d_{n,3}$ and therefore it would necessarily return a hypothesis with a large error. This argument can be used to show that, unless $NP=RP$, learning ${\mathcal H}^d_{n,3}$ with a {\em proper} efficient algorithm is impossible. However, here we want to rule out {\em improper} algorithms as well. The crux of the construction is that if $\phi$ is random, {\em no algorithm} (even improper and even inefficient) can return a hypothesis with a small error. The reason for that is that since the sample provided to the algorithm consists of only $\kappa \frac{n}{0.01^2}$ samples, the algorithm won't see most of $\psi$'s clauses, and, consequently, the produced hypothesis $h$ will be {\em independent of them}. Since these clauses are random, $h$ is likely to err on about half of them, so that $\Err_{D_{\phi}}(h)$ will be close to half! To summarize we constructed an efficient algorithm with the following properties: if $\phi$ is almost satisfiable, the algorithm will return a hypothesis with a small error, and then we will declare ``exceptional'', while for random $\phi$, the algorithm will return a hypothesis with a large error, and we will declare ``typical''. Our construction crucially relies on the restriction to learning algorithm with a small sample complexity. Indeed, if the learning algorithm obtains more than $n^{1+\mu}$ examples, then it will see most of $\psi$'s clauses, and therefore it might succeed in ``learning'' even when the source of the formula is random. Therefore, we will declare ``exceptional'' even when the source is random. \begin{proof} (of theorem \ref{thm:main}) Assume by way of contradiction that the $\mu$-R3SAT hardness assumption is true and yet there exists an efficient learning algorithm that learns the class ${\mathcal H}_{n,3}$ using $O\left(\frac{n^{1+\mu}}{\epsilon^2}\right)$ examples. Setting $\epsilon=\frac{1}{100}$, we conclude that there exists an efficient algorithm $L$ and a constant $\kappa>0$ such that given a sample $S$ of more than $\kappa\cdot n^{1+\mu}$ examples drawn from a distribution ${\mathcal D}$ on $C_{n,3}\times \{\pm 1\}$, returns a classifier $L(S):C_{n,3}\to \{\pm 1\}$ such that \begin{itemize} \item $L(S)$ can be evaluated efficiently. \item W.p. $\ge\frac{3}{4}$ over the choice of $S$, $\Err_{{\mathcal D}}(L(S))\le \Err_{{\mathcal D}}({\mathcal H}_{n,3})+\frac{1}{100}$. \end{itemize} Fix $\Delta$ large enough such that $\Delta>100 \kappa$ and the conclusion of Theorem \ref{hyp:feige_maj} holds with $\epsilon=\frac{1}{100}$. We will construct an algorithm, $A$, contradicting Theorem \ref{hyp:feige_maj}. On input $\phi\in \textrm{3MAJ}_{n,\Delta n^{1+\mu}}$ consisting of the $\textrm{3MAJ}$ clauses $C_1,\ldots, C_{\Delta n^{1+\mu}}$, the algorithm $A$ proceeds as follows \begin{enumerate} \item Generate a sample $S$ consisting of $\Delta n^{1+\mu}$ examples as follows. For every clause, $C_k=\textrm{MAJ}((-1)^{j_1}x_{i_1},(-1)^{j_2}x_{i_2},(-1)^{j_3}x_{i_3})$, generate an example $(x_k,y_k)\in C_{n,3}\times \{\pm 1\}$ by choosing $b\in \{\pm 1\}$ at random and letting $$(x_k,y_k)=b\cdot\left(\sum_{l=1}^3 (-1)^{j_l}e_{i_l},1\right)\in C_{n,3}\times \{\pm 1\}~.$$ For example, if $n=6$, the clause is $\textrm{MAJ}(- x_2,x_3,x_6)$ and $b=-1$, we generate the example $$\left((0,1,-1,0,0,-1),-1 \right)$$ \item Choose a sample $S_1$ consisting of $\frac{\Delta n^{1+\mu}}{100}\ge \kappa\cdot n^{1+\mu}$ examples by choosing at random (with repetitions) examples from $S$. \item Let $h=L(S_1)$. If $\Err_{S}(h)\le \frac{3}{8}$, return $\mathrm{``exceptional"}$. Otherwise, return $\mathrm{``typical"}$. \end{enumerate} We claim that $A$ contradicts Theorem \ref{hyp:feige_maj}. Clearly, $A$ runs in polynomial time. It remains to show that \begin{itemize} \item If $\val(\phi)\ge \frac{3}{4}-\frac{1}{100}$, then \[ \Pr_{\text{Rand. coins of }A}\left(A(\phi)=\mathrm{``exceptional"}\right) \ge \frac{3}{4} \] \item For every $n$, \[ \Pr_{\text{Rand. coins of }A,\;\phi\sim \mathrm{Uni}(\textrm{3MAJ}_{n,\Delta n^{1+\mu}})}\left(A(\phi)=\mathrm{``typical"}\right) \ge 1-o(1) \] \end{itemize} Assume first that $\phi\in \textrm{3MAJ}_{n,\Delta n^{1+\mu}}$ is chosen at random. Given the sample $S_1$, the sample $S_2:=S\setminus S_1$ is a sample of $|S_2|$ i.i.d. examples which are independent from the sample $S_1$, and hence also from $h=L(S_1)$. Moreover, for every example $(x_k,y_k)\in S_2$, $y_k$ is a Bernoulli random variable with parameter $\frac{1}{2}$ which is independent of $x_k$. To see that, note that an example whose instance is $x_k$ can be generated by exactly two clauses -- one corresponds to $y_k=1$, while the other corresponds to $y_k=-1$ (e.g., the instance $(1,-1,0,1)$ can be generated from the clause $\textrm{MAJ}(x_1,-x_2,x_4)$ and $b=1$ or the clause $\textrm{MAJ}(-x_1,x_2,-x_4)$ and $b=-1$). Thus, given the instance $x_k$, the probability that $y_k=1$ is $\frac{1}{2}$, independent of $x_k$. It follows that $\Err_{S_2}(h)$ is an average of at least $\left(1-\frac{1}{100}\right)\Delta n^{1+\mu}$ independent Bernoulli random variable. By Chernoff's bound, with probability $\ge 1-o(1)$, $\Err_{S_2}(h)>\frac{1}{2}-\frac{1}{100}$. Thus, $$\Err_{S}(h)\ge \left(1-\frac{1}{100}\right)\Err_{S_2}(h)\ge \left(1-\frac{1}{100}\right)\cdot \left(\frac{1}{2}-\frac{1}{100}\right)>\frac{3}{8}$$ And the algorithm will output $\mathrm{``typical"}$. Assume now that $\val(\phi)\ge\frac{3}{4}-\frac{1}{100}$ and let $\psi\in\{\pm 1\}^n$ be an assignment that indicates that. Let $\Psi\in {\mathcal H}_{n,3}$ be the hypothesis $\Psi(x)=\textrm{sign}\left(\inner{\psi,x}\right)$. It can be easily checked that $\Psi(x_k)=y_k$ if and only if $\psi$ satisfies $C_k$. Since $\val(\phi)\ge\frac{3}{4}-\frac{1}{100}$, it follows that $$\Err_{S}(\Psi)\le \frac{1}{4}+\frac{1}{100} ~.$$ Thus, $$\Err_{S}({\mathcal H}_{n,3})\le \frac{1}{4}+\frac{1}{100} ~.$$ By the choice of $L$, with probability $\ge 1-\frac{1}{4}=\frac{3}{4}$, $$\Err_S(h)\le \frac{1}{4}+\frac{1}{100}+\frac{1}{100}< \frac{3}{8}$$ and the algorithm will return $\mathrm{``exceptional"}$. \end{proof} \section{Upper bounds for learning ${\mathcal H}_{n,2}$ and ${\mathcal H}_{n,3}$}\label{sec:upper_bounds} The following theorem derives upper bounds for learning ${\mathcal H}_{n,2}$ and ${\mathcal H}_{n,3}$. Its proof relies on results from \cite{HazanKaSh12} about learning $\beta$-decomposable matrices, and due to the lack of space is given in the appendix. \begin{theorem}\label{thm:learning_H_new} ~ \begin{itemize} \item There exists an efficient algorithm that learns ${\mathcal H}_{n,2}$ using $O\left(\frac{n\log^3(n)}{\epsilon^2}\right)$ examples \item There exists an efficient algorithm that learns ${\mathcal H}_{n,3}$ using $O\left(\frac{n^2\log^3(n)}{\epsilon^2}\right)$ examples \end{itemize} \end{theorem} \section{Discussion} We formally established a computational-sample complexity tradeoff for the task of (agnostically and improperly) PAC learning of halfspaces over $3$-sparse vectors. Our proof of the lower bound relies on a novel, non cryptographic, technique for establishing such tradeoffs. We also derive a new non-trivial upper bound for this task. {\bf Open questions.} An obvious open question is to close the gap between the lower and upper bounds. We conjecture that ${\mathcal H}_{n,3}$ can be learnt efficiently using a sample of $\tilde{O}\left(\frac{n^{1.5}}{\epsilon^2}\right)$ examples. Also, we believe that our new proof technique can be used for establishing computational-sample complexity tradeoffs for other natural learning problems. \paragraph{Acknowledgements:} Amit Daniely is a recipient of the Google Europe Fellowship in Learning Theory, and this research is supported in part by this Google Fellowship. Nati Linial is supported by grants from ISF, BSF and I-Core. Shai Shalev-Shwartz is supported by the Israeli Science Foundation grant number 590-10. \bibliographystyle{plainnat}
\section{Introduction} Saturn's upper atmosphere is defined as the region at which the neutral molecular species within cease to mix convectively and are thus able to separate according to their scale heights; it is dominated by H, H$_2$, and He and the boundary between the mixing and non-mixing regions is known as the homopause \citep{2009nagy}. Within the upper atmosphere, solar radiation, particularly extreme ultraviolet (UV) is responsible for the production of the Saturnian ionosphere \citep{2009Moore}. In addition to this, the impact of electrons in the polar regions (accelerated along magnetic field lines), is responsible for the creation of the Saturnian aurora \citep[e.g.,][]{lamy2009uvskr}. Study of the Saturnian aurorae is generally divided into two wavelengths, the UV and infrared (IR). The former reveals the impact of highly energetic electrons on the polar ionosphere through the excitation of H and H$_2$ \citep{1983shemansky,gustin09}, whilst the latter is the focus of this study and is observed primarily via the discrete ro-vibrational emission lines of the molecular ion H$_3^+$. First detected on Saturn in late 1992 \citep{1993geb}, H$_3^+$ has served as a useful probe for examining the conditions in the ionosphere of Saturn. H$_3^+$ is produced indirectly from the ionisation of H$_2$, driven in the polar regions by particle precipitation at the boundary between open and closed planetary magnetic field lines \citep{Cow04open}, statistically located near $\sim$15$^{\circ}$ co-latitude \citep{badmansatwid}. Since H$_3^+$ is quasi-thermalized in the upper atmospheres of the giant planets \citep{1990mill}, examining emissions in the polar regions thus allows for the study of both the interactions between the ionosphere and its immediate space environment, and the physical conditions in the surrounding neutral atmosphere. \\ Prior to the findings of this study, measurements of auroral thermosphere temperatures have been limited to time-averaged values as Saturn's auroral H$_3^+$ emissions in the infrared are relatively weak: they are less than a hundredth the intensity of those at Jupiter when observed from Earth owing to Jupiter's higher ionospheric temperature \citep[$\sim$1000 K,][]{lam97} compared to Saturn's as we shall see, and this rising temperature causing an almost exponential increase in H$_3^+$ emission \citep[see][Figure 2]{96line}. \citet{melin07a} performed the most recent ground-based study of auroral thermospheric temperatures at Saturn, combining results from the 3.8-metre UKIRT telescope and CGS4 spectrograph obtained in 1999 and 2004, with exposure times of 210 and 26 mins, to derive a temperature of 450$\pm{50}$ K for the southern spring/summer auroral thermosphere. More recently, data from the Visual and Infrared Mapping Spectrometer (VIMS) on board the Cassini Saturn orbiter \citep{vims}, analysed by \citet{tssras} and \citet{melin11}, showed temperatures in the southern auroral region of 560-620$\pm{30}$ K over a 24-hour period in June 2007, and 440$\pm{50}$ K in measurements in September 2008, respectively. These temperatures are far warmer, by several hundreds of Kelvin, than predicted by models using only the Sun as an energy input \citep{YM}. We shall not heavily dwell on the longer term (months and years) effects here, except to say that both the IR and UV components of auroral emissions are modulated by solar wind conditions which do vary on such time scales, a fact owing to the variable solar wind dynamic pressure exerted on Saturn's magnetosphere \citep[see][and references therein]{kurthbook}.\\ Heating in the auroral region is thought to be dominated by Joule heating and ion drag via ionospheric Pedersen currents \citep{Cow04open,2005smith,aurtrans,2011galand,ingo2012}. The total heating from these two mechanisms generates $\sim$5 TW of power per hemisphere, whilst auroral particle precipitation provides an additional energy input at $\sim$0.1 TW \citep{Cow04open}. Only through a greater understanding of the mechanisms and conditions that cause physical parameters such as temperature, column density, and the total emitted energy over all wavelengths (henceforth, total emission) to persist or vary, can we start to add constraints to models and theories of the ionosphere. However, whilst individual temperature measurements have been made over long time scales, a study in both hemispheres simultaneously has not yet been performed. Here, we present and discuss the results of observations with the Near-Infrared SPECtrometer (NIRSPEC) \citep{nirs} instrument in high-resolution mode using the 10-metre W.M. Keck telescope situated on Mauna Kea, Hawaii. We study the main auroral region H$_3^+$ temperature, column density and total emission in the northern and southern hemispheres of Saturn at the same time with a temporal resolution of 15 minutes, and explore the implications of these measurements. \section{Observations} This study examines 132 minutes of observations obtained on 17 April 2011 using the 10-metre Keck telescope. These data were obtained using the NIRSPEC \citep{nirs} in high-resolution cross-dispersed mode with a resolution of R = $\lambda$/$\Delta\lambda$ $\sim$25,000, providing a minimum resolution of $\Delta\lambda\approx$ 1.59 x 10$^{-4}$ $\mu$m at 3.975 $\mu$m. The wavelength range used in this study is between 3.95 and 4.0 $\mu$m, covering the Q-Branch ($\Delta$J=0) ro-vibrational transition lines of H$_3^+$. The slit of the spectrometer was orientated in a north-south direction on Saturn aligned along the rotational axis, which we note is also co-aligned with the magnetic axis to within measurement uncertainties of $\sim$0.1$^{\circ}$ \citep{satintfield}. The planet is then seen to rotate beneath the slit allowing the acquisition of spectral images at a fixed local time of noon, but with a varying Saturn System III Central Meridian Longitude (CML). We were able to collect data between $\sim$10$^{\circ}$ and $\sim$22$^{\circ}$ co-latitude in both the northern and southern hemispheres given the viewing geometry at the time. The slit measures 0.432$^{\prime\prime}$ width by 24$^{\prime\prime}$ length with a pixel on the CCD corresponding to 0.144$^{\prime\prime}$ squared on the sky. The atmospheric seeing during this period was $\sim$0.4$^{\prime\prime}$. This dataset was recorded between 10:33:42 and 12:46:28 Universal Time (UT), covering a CML on Saturn of 103-176$^{\circ}$. Each set of spectra taken consist of twelve 5-s integrations, creating exposures 60 s long, consisting of object, or Saturn, `A' and sky `B' frames with the telescope slewing between the relevant positions of each in the sky in an ABBA pattern. An example of a typical 60 second `A-B' frame exposure (unreduced apart from the sky subtraction) is shown in Figure 1. The reduced spectral images are co-added over 15 minute segments to improve the signal to noise (S/N) ratio (this is elaborated on in the Data reduction section). During one such segment, Saturn rotates through 8.5$^{\circ}$ of longitude. On 17 April 2011, Saturn's northern hemisphere was tilted towards the Earth with a sub-Earth latitude of 8.2$^{\circ}$: in conditions of Saturn's northern spring. \begin{figure} \centerline{\includegraphics[width=16.5cm]{TIMEfig1_v5.eps}} \caption{A typical sky-subtracted (A-B) spectral image of Saturn taken at Saturn local noon. The wavelength range is shown on the horizontal axis and the angular size in the sky is shown on the vertical axis. A real image of Saturn is shown to the right (taken with the Keck II telescope guide-camera) with arrows indicating the position of the aurorae and rings. Discrete H$_3^+$ emission lines can be seen as white vertical lines within the indicated yellow boxes. Hydrocarbon absorption of solar radiation appears as black between the auroral regions regions; hydrocarbons follow the general formula C$_n$H$_m$ where n and m are integer numbers, for example, methane is CH$_4$. The white bar of emission at -2$^{\prime\prime}$ is the continuum reflection of sunlight from the rings. The remaining white pixels are due to reflected sunlight.} \end{figure} \\ \section{Data reduction and analysis} We examined the data summed between $\sim$10$^{\circ}$ and $\sim$22$^{\circ}$ co-latitude in the northern and southern hemispheres as stated, we found that the errors in parameters such as temperature were too large to perform a study of small-scale ($\sim$500-1000's km) structures known to exist within, for example, in the UV aurorae \citep{2011grod,mere13}. The error in assigning pixels to co-latitudes (pixel registration) was based on fitting a Gaussian curve to the large and very bright rings to find an exact latitude on the planet and then mapping poleward from that position to the planet's limbs, giving a negligible error of $\textless$0.1 pixel for each spectral image. The random error introduced by atmospheric seeing is $\sim$3 pixels, such that the pixel rows containing the main auroral emissions (which are well constrained in position), is able to receive light from adjacent pixels. This `smearing' of data becomes more prevalent towards the poles, leading to $\sim$3-7 degrees of smear for the auroral regions under study. In addition, the main auroral regions themselves vary in width and location within the aforementioned 12$^{\circ}$ co-latitude range. Standard sky subtraction and flux calibration (using the star HR 6035) techniques were applied to the data, accounting for the Earth's atmosphere, and correlating CCD count to physical photon flux. The methane (CH$_4$) hydrocarbon present in Saturn's upper atmosphere acts to both reflect and absorb solar radiation, depending on the wavelength of light. In Figure 1, sunlight is strongly reflected at some wavelengths, but CH$_4$ acts to absorb light at other wavelengths. Due to the increased column depth of CH$_4$ towards the limb, sunlight is absorbed enough to see auroral H$_3^+$ emission with little interference. However, as the reflected sunlight increases equatorward it eventually swamps the signal from most discrete H$_3^+$ emission lines, the low-latitude H$_3^+$ emissions are discussed in \citet{paper1}. The co-latitude range we selected led to a S/N ratio of $\sim$10-30, which is appropriate to give an uncertainties of between 5 and 10\%. Reflected sunlight is removed from the auroral emission by measuring the equatorial solar spectrum then subtracting an empirically calculated proportion of it from our auroral spectra. With this subtraction, the only significant spectral signature that remains is the auroral H$_3^+$ emission. \\ For a given temperature, H$_3^+$ produces a unique spectrum, such that there is a fixed temperature-dependent ratio between emission lines at different wavelengths. The H$_3^+$ temperature, T(H$_3^+$), is found using a fitting routine which uses the spectroscopic line list from \citet{96line} and the latest H$_3^+$ partition function constants from \citet{2010Miller} to produce expected transition-line intensities for a given temperature, this is also described by \citet{Melin2013u} in further detail. The spectral function of H$_3^+$ is varied until the line-ratios match the least squares fit to the observed data, as shown by the 15 minute-integrated spectrum in Figure 2. For a given temperature the emission from a single H$_3^+$ molecule is known, so by dividing the observed emission by the molecular emission, we can find the number of emitting molecules, i.e. the H$_3^+$ column density. We produce a line-of-sight corrected column density, N(H$_3^+$), by measuring the differing observed path-lengths through the atmosphere across the disk of the planet. This varies as the sine of the colatitude as observed from Earth, accounting for the varying line-of-sight column depth as it increases towards the poles, e.g. multiplying N(H$_3^+$) by the sine of the colatitude in which it is located, will reduce the observed density to create a column density. Saturn's sub-Earth latitude was 8.2$^{\circ}$, by the addition of this angle to the observed latitude we correctly locate the position of pixels, e.g. at 80$^{\circ}$ north from Earth's perspective is actually 88.2$^{\circ}$ on Saturn. A measure of wavelength-integrated total emission from a line-of-sight corrected column of H$_3^+$, E(H$_3^+$) can then be calculated by multiplying the total emission per molecule by N(H$_3^+$). This parameter was introduced by \citet{lam97} for Jupiter, to be used as a separate parameter for studying the ionosphere, due to their observation of an anti-correlation between temperature and column density. It is important to note that whilst the temperature is measured using the relative amplitudes of the various spectral peaks, the density is derived subsequent to the determination of temperature. Therefore, even though the least-squares fitting routine described above was used, the derived T(H$_3^+$) and N(H$_3^+$) are relatively independent parameters. E(H$_3^+$) is a useful parameter as it reveals the amount of energy lost by the ionosphere via radiative cooling to space. Throughout this paper the errors shown for the above parameters are $\pm$1 standard deviation (1-sigma), these arise from the uncertainty in the Gaussian fits to each spectral line (see Figure 2). \\ \begin{figure} \centerline{\includegraphics[width=13cm]{TIMEFig2_v2.eps}} \caption{An example model fit to the data. H$_3^+$ emission is shown as a function of wavelength (black crosses), fitted to a model of the expected emission (red) for a given temperature. The data in this figure are from a 0.144$^{\prime\prime}$ (height) by 0.432$^{\prime\prime}$ (width) area of the planet, which corresponds to an area of approximately 3320 km x 2700 km at 19$^{\circ}$ co-latitude. From this fit we obtained a temperature of 523$\pm$13 K. The S/N ratio is $\sim$25 in this typical spectral profile. Low levels of solar reflection from hydrocarbons are visible over all wavelengths, though these levels are much reduced from their original values by the empirically calculated solar reflection subtraction.} \end{figure} \section{Results and discussion} In Figure 3 we present simultaneous measurements of Saturn's H$_3^+$ temperature, column density, and total emission in the northern and southern auroral regions at local noon as a function of time. We view the aurorae as they rotate past the spectrometer slit and so variability is a combination of temporal changes occurring at local noon and longitudinal variations rotating into view. The relationships between the temperature, column density, and total emission between hemispheres are investigated. A summary of the results is also given in Table 1. The northern thermospheric temperature is on average 527$\pm$18 K, while the southern is 583$\pm$13 K. The column density averages for the north and south aurorae are 1.56$\pm$0.32 x10$^{15}$ m$^{-2}$ and 1.16$\pm$0.14 x10$^{15}$ m$^{-2}$, respectively. An anti-correlation between H$_3^+$ temperature and column density is observed in our data. The total emission is $\sim$1.5 times higher in the south, 0.98$\pm$0.02 Wm$^{-2}$str$^{-1}$, compared with 0.65$\pm$0.03 Wm$^{-2}$str$^{-1}$ in the north. This result is similar to previous work based on Cassini VIMS observations in which they examined IR wavelengths associated with H$_3^+$ emission at $\sim$3.6 $\mu$m, which showed the pre-equinox southern main region to be on average $\sim$1.3 times more intense than the northern main auroral region \citep{2011badman}. The higher levels of emission cause the S/N ratio in the south is$\sim$30 whilst the north is $\sim$18 on average, so that the errors in all parameters are lower there relative to the north (see again Table 1). \begin{landscape} \begin{table} \caption{Saturn's main auroral region properties as a function of time. Data obtained on 17 April 2011. All uncertainties shown are one standard deviation (i.e. 1-sigma errors).} \begin{tabular}{c c c c c c c c} Start Time & E(North,H$_3^+$) & T(North,H$_3^+$) & N(North,H$_3^+$) & E(South,H$_3^+$) & T(South,H$_3^+$) & N(South,H$_3^+$)\\ (UT) & (x10$^{-5}$ Wm$^{-2}$sr$^{-1}$) & (Kelvin) & (x10$^{15}$ m$^{-2}$) & (x10$^{-5}$ Wm$^{-2}$sr$^{-1}$) & (Kelvin) & (x10$^{15}$ m$^{-2}$) \\ \hline $10:33$ & 0.81 $\pm$0.03 & 528 $\pm$17 & 1.85 $\pm$0.34 & 0.92 $\pm$0.02 & 592 $\pm$13 & 0.96 $\pm$0.11 \\ $10:49$ & 0.69 $\pm$0.02 & 529 $\pm$17 & 1.59 $\pm$0.30 & 0.90 $\pm$0.02 & 580 $\pm$14 & 1.10 $\pm$0.14 \\ $11:02$ & 0.64 $\pm$0.02 & 539 $\pm$18 & 1.33 $\pm$0.25 & 0.89 $\pm$0.02 & 575 $\pm$14 & 1.16 $\pm$0.15 \\ $11:19$ & 0.66 $\pm$0.03 & 544 $\pm$19 & 1.29 $\pm$0.25 & 0.96 $\pm$0.02 & 588 $\pm$14 & 1.07 $\pm$0.13 \\ $11:36$ & 0.64 $\pm$0.03 & 537 $\pm$19 & 1.41 $\pm$0.28 & 1.00 $\pm$0.02 & 586 $\pm$13 & 1.15 $\pm$0.13 \\ $11:49$ & 0.61 $\pm$0.03 & 528 $\pm$20 & 1.49 $\pm$0.33 & 1.06 $\pm$0.02 & 585 $\pm$13 & 1.22 $\pm$0.14 \\ $12:04$ & 0.57 $\pm$0.03 & 513 $\pm$20 & 1.66 $\pm$0.39 & 1.06 $\pm$0.02 & 580 $\pm$12 & 1.30 $\pm$0.20 \\ $12:17$ & 0.60 $\pm$0.03 & 521 $\pm$20 & 1.51 $\pm$0.35 & 1.02 $\pm$0.02 & 579 $\pm$13 & 1.25 $\pm$0.14 \\ $12:32$ & 0.62 $\pm$0.02 & 517 $\pm$18 & 1.66 $\pm$0.34 & 0.99 $\pm$0.02 & 580 $\pm$13 & 1.19 $\pm$0.14 \\ $12:46$ & 0.65 $\pm$0.02 & 515 $\pm$15 & 1.77 $\pm$0.32 & 0.97 $\pm$0.02 & 579 $\pm$12 & 1.20 $\pm$0.13 \\ \end{tabular} \end{table} \end{landscape} \begin{figure} \noindent\includegraphics[width=34pc]{TIMEFig3_v3.eps} \caption{Temperature, H$_3^+$ column density and total emission of Saturn's main auroral region emissions for each hemisphere integrated over 10-22$^{\circ}$ co-latitude (y-axis), plotted as a function of time (x-axis). Northern data are shown in blue, and southern data in red. The thin dark coloured lines show data values, while the light coloured shading shows their corresponding uncertainty ranges. The error bars show uncertainties of one standard deviation (i.e. 1-sigma error bars). The time at 0 minutes is 10:34 UT on 17 April 2011.} \end{figure} \subsection{Interhemispheric asymmetry in temperature and emission} The average auroral thermospheric temperature in the south, $\sim$583 K, is within the 560-620 K range found by \citet{tssras} using Cassini VIMS data from June 2007. However, this is substantially higher than the ground-based 1999/2004 UKIRT result of 400$\pm$50 K found by \citet{melin07a} and the Cassini VIMS September 2008 result of 440$\pm$50 K by \citet{melin11}. These differences suggest that while temperatures are stable on the short time scales observed here, highly variable auroral temperatures can be seen on longer time scales. The relatively higher temperatures in both hemispheres here may indicate that Saturn is in the midst of a slow `heating event' on time scales greater than hours. Such an event has been observed on Jupiter by \citet{aurtrans} and takes place over a period of 3 Earth days, but this heating event was due to the compression of closed field lines, such that we can expect the effects to be symmetric north and south. Only through further observations taken of the same latitudes can we identify the nature of such long-term trends. The most striking result shown in Figure 3 is that the southern auroral thermosphere is significantly hotter and more emissive than the north over the $\sim$2 hour duration of these observations. Although the observations represent a `snapshot' of the possible conditions in Saturn's ionosphere, the following discussions and conclusions assume this represents conditions that are typical on Saturn at that time in its season. Additional observations are required over time scales of weeks and months, to validate that this asymmetry is not due to short term (hours or days) effects. To investigate the reasons for this unexpected temperature difference, we consider the combined Joule and ion drag heating rate per unit area of the ionosphere, in particular the effect of the hemispheric difference in ionospheric magnetic field strength, where the northern polar field is a factor $\sim$1.2 times the strength of the southern polar field (both integrated between $\sim$10$^{\circ}$-22$^{\circ}$) due to the quadrupole term in the planet's internal field. This is illustrated in Figure 4, where we plot the field strength in the Pedersen layer versus co-latitude from the respective poles for the northern (solid line) and southern (dashed) hemispheres, respectively. Here we have used the latest internal field model based on Cassini data by \citet{satintfield}, consisting of axial dipole, quadrupole and octupole terms, evaluated at an altitude of 1000 km above the IAU 1 bar reference spheroid. The latter 1 bar surface has equatorial and polar radii of 60,268 and 54,364 km, respectively, with the Pedersen layer located $\sim$1000 km above \citep[e.g.,][]{satintfield}. \begin{figure} \noindent\includegraphics[width=33pc]{TIMEfig4.eps} \caption{Saturn's internal magnetic field strength $\mid$B$\mid$ in the ionospheric Pedersen layer using the \citet{satintfield} model, shown plotted as a function of planetocentric co-latitude in degrees from the corresponding pole for the northern (solid) and southern (dashed) hemispheres, respectively. The Pedersen layer is taken to lie 1000 km above the IAU 1 bar pressure reference spheroid. The vertical lines (dotted) indicate the range of auroral region co-latitudes.} \end{figure} The combined Joule and ion drag thermospheric heating rate per unit area in the northern (N) and southern (S) hemispheres is given by (e.g., \citet{2005smith,2005cow}) \begin{equation} q_{N,S}=\Sigma^*_{P_{N,S}}E^2_{eqN,S}~, \end{equation} where $\Sigma^*_{P_{N,S}}$ is the effective height-integrated Pedersen conductivity of the ionosphere, modified from the true value $\Sigma_{P_{N,S}}$ due to drag-induced atmospheric sub-corotation, and $E_{eqN,S}$ is the equatorward-directed ionospheric electric field (\textbf{\textit{E}}= -\textbf{\textit{V}}$\times$\textbf{\textit{B}}) in the rest frame of the planet at a given co-latitude with respect to the rotation/magnetic axis. The latter is given by \begin{equation} E_{eqN,S} = \rho_{iN,S}(\Omega_{Sat} - \omega ) B_{iN,S}~, \end{equation} where $\rho_{iN,S}$ is the perpendicular distance of the Pedersen conducting layer from the axis, $\Omega_{Sat}$ is the angular velocity of Saturn defining the planetary `rest frame' of rigid corotation, $\omega$ is the magnetospheric plasma angular velocity on the field line passing through the ionosphere at that co-latitude, and $B_{iN,S}$ is the corresponding ionospheric magnetic field strength. The field is taken to be uniform and perpendicular to the polar ionosphere to a sufficient approximation, the latter peaking in Pedersen conductivity at an altitude of $\sim$1000 km in the auroral region \citep{2011galand}. The effective Pedersen conductivity is given by \begin{equation} \Sigma^*_{P_{N,S}} = (1-k)\Sigma_{P_{N,S}}~, \end{equation} where \textit{k} is the ratio between the neutral atmosphere angular velocity and the plasma angular velocity in the planet's frame \begin{equation} (\Omega_{Sat} - \Omega^*_{Sat} ) = k(\Omega_{Sat}-\omega)~, \end{equation} where $\Omega^*_{Sat}$ is the angular velocity of the neutral atmosphere. Atmospheric modelling results indicate that \textit{k}$\sim$0.5 at Saturn \citep{2011galand}. Combining Equations (1) and (2) we obtain \begin{equation} q_{N,S} = \Sigma^*_{P N,S}~\rho^2_{iN,S}~(\Omega_{Sat}-\omega)^2~B^2_{iN,S}~. \end{equation} Now to the same order of field approximation (in which terms quadratic in the co-latitude are neglected relative to zeroth order) the magnetic flux threading the ionosphere between the pole and cylindrical distance $\rho_{iN,S}$ from the axis is \begin{equation} \Phi_i = \pi\rho^2_{iN,S}~B_{iN,S}~, \end{equation} such that we can write Equation (5) as \begin{equation} q_{N,S}= \frac{1}{\pi}~\Sigma^*_{P N,S}~(\Omega_{Sat}-\omega)^2~\Phi_iB_{iN,S}~. \end{equation} If we then consider conjugate points in the northern and southern hemispheres joined by a common field line, thus containing equal magnetic flux $\Phi_i$ and having equal plasma angular velocities $\omega$, as required under steady state conditions, it can be seen that the relative heating rates per unit area north and south depend only on the product of the effective height-integrated Pedersen conductivity, $\Sigma^*_{P N,S}$, and the field strength in the ionosphere, $B_{iN,S}$. However, for approximately equal ionospheric Pedersen layer number densities north and south (as addressed in the next section), the Pedersen conductivity is expected to vary approximately inversely as the ionospheric field strength, as reported by \citet{2011galand} for near-equinoctial conditions. In these circumstances the thermospheric heating rates per unit area will be equal in the two hemispheres at conjugate points, independent of the magnetic field strength. This result does not therefore give immediate reason to expect the southern thermosphere to be hotter than the northern, unless the northern ionospheric conductivity is lower than that in the south by an unexpectedly large factor. We note, however, that the above result also implies that the total heat input to the thermosphere from Joule heating and ion drag integrated over the whole polar region will be larger in the south than in the north, because the area of heating is larger in the south than in the north due to the lower field strength. If we consider conjugate circular ionospheric strips north and south with equal magnetic flux $d\Phi_{i}=2\pi B_{iN,S}~\rho_{iN,S}~d\rho_{iN,S}$, then the total heating north and south in the strips is given by \begin{equation} dQ_{N,S}= 2\pi q_{N,S}~\rho_{i,N,S}~d\rho_{iN,S}= \frac{1}{\pi}\Sigma^*_{P N,S}~(\Omega_{Sat}-\omega)^2~\Phi_id\Phi_i~. \end{equation} Thus the total heating rate, obtained by integrating over all the flux strips from the pole to the point where rigid corotation is attained, is then proportional only to $\Sigma^*_{P N,S}$, such that if the latter varies approximately inversely with the field strength as indicated above, the total power dissipated to heat in the polar thermosphere will larger in the southern hemisphere than in the northern. It remains to be investigated by modelling whether such an effect could produce the temperature differences measured here. If not, then some other heating mechanism, such as hemispheric differences in wave driving from below, must be implicated. A comparison with recent modelling work by \citet{2011galand} agrees with the interpretation above in that an asymmetry is present (during equinox) in which the Pedersen and Hall conductivities were 1.2 and 1.3 times higher, respectively, in the southern hemisphere than in the northern. Previous observations by Cassini VIMS analysed by \citet{2011badman} also found the same trend in intensity - and likely therefore in temperature - in their pre-equinox 2006-2009 data. The fact that this asymmetry persists post-equinox in our data suggests that the magnitude of the magnetic field is the dominant effect on Pedersen conductivity rather than the solar extreme-UV ionization, at least in northern spring. In other words, magnetic field strength may dominate over seasonal effects in determining inter-hemispheric auroral thermosphere temperatures, though further observations and modelling are required to test this - in particular whether or not it persists into the northern summer season. \\ In the UV, simultaneous observations of the conjugate northern and southern aurora taken by the HST in 2009 have been analysed by \citet{UVaurora} and \citet{mere13}. The former showed, from the data acquired over a period of $\sim$1 month just pre-equinox, that the northern main auroral region had on average $\sim$17\% higher emitted power than the south, the opposite to the IR case presented here. The latter study found transient eastward-propagating patches of UV emission in the dawn-to-noon sector for 70\% of the 32 visits using the HST, these patches are similar to the small-scale features found by \cite{2011grodgrapes}. In this study, such small-scale features are therefore very likely to be passing by the spectrograph slit, and could lead to small-scale variations in the column density of H$_3^+$. However, UV emissions are a prompt emission in which hydrogen is excited by particle precipitation and immediately releases the newly acquired energy to space via the emission of UV photons. Hydrogen that emits in the UV is not therefore in thermal equilibrium with the surrounding thermosphere. By contrast, H$_3^+$ emission is largely driven by temperature changes due to Joule heating and ion drag, so to a good approximation H$_3^+$ is thermalized with the thermosphere \citep{millsept06}. As a consequence of the differing IR and UV emission production mechanisms, direct comparison is difficult. Given that Joule heating and ion drag is $\sim$50 times greater in power than auroral particle precipitation \citep{Cow04open} (responsible for UV emission), it is understandable in the above UV studies that a stronger northern UV emission or the appearance of small-scale structures/patches need not necessarily correspond to higher temperatures or IR emission. We were unable to resolve small-scale features here with small uncertainties, so we cannot compare individual features. In addition, it should be noted that the above UV observations took place over 2 years earlier than those presented here. \subsection{Altitudinal considerations} The ratio between the average column integrated densities in the north and south auroral regions is 1.35. Broadly speaking, the cause for this asymmetry could be an increase in the northern H$_2$ ionization rate, which itself arises from the larger incident solar photon flux in the north owing to Saturn's 9.1$^{\circ}$ sub-solar latitude. An increase in ionized H$_2$ then leads to a greater production of H$_3^+$. This was also demonstrated using a 1-D model, the Saturn Thermosphere Ionosphere Model (STIM), which was utilised by the work presented here and found a range of north-south H$_3^+$ density ratios between 1.2-3.8 based on solar EUV influx alone (between 10-22$^{\circ}$ co-latitude), for several different values of vibrationally excited H$_2$. \\ Despite having lower column densities, the average total emission is $\sim$1.5 times higher in the south, this is due to the higher temperature there. The total emission is on average not related to the column density as there are fewer H$_3^+$ molecules emitting and yet there is more emission, so there is clearly more emission per molecule because the emitting molecules are hotter. \\ One might assume that a higher H$_3^+$ density and lower temperature (as in the northern auroral thermosphere) could indicate that the column of H$_3^+$ sampled was deeper in the atmosphere: an inverse relationship like this exists at Jupiter \citep{lystrup8}. As previously stated, the H$_3^+$ densities presented herein are column integrated and line of sight corrected, such that altitudinal information is averaged for the observed atmospheric column. If the H$_3^+$ densities were higher in the northern hemisphere relative to the south because of the inverse relationship above, it implies that electrons must penetrate deeper in the north, thus leading to enhancements in H$_3^+$ production (density) in a colder region. For this to occur, the electron precipitation energy must be relatively higher in the northern auroral region, since higher energy electrons penetrate to lower altitudes than lower energy electrons \citep{2011Tao}. To test this, we employed the magnetosphere-ionosphere coupling model known as the `CBO' model, derived by \citet{2003cow} and \citet{Cow04open}, and used updated parameters derived from Cassini spacecraft measurements \citep{2008cow}. It is appropriate to reduce the northern conductivity from 4 mho in \citet{2008cow} by a factor of 1.215 corresponding to the field asymmetry already mentioned, such that it is fixed at 3.3 mho whilst the south remains at 4 mho. Following this, we find that the field-aligned current density, precipitating electron energy flux, and average electron precipitation energy in both hemispheres are closely similar, the latter parameter being $\sim$11.4 keV. Therefore, we have no reason to expect strong differences in the altitude at which auroral electrons are deposited in either hemisphere, nor the overall number density profile at those altitudes. This expectation is based on electrons accelerated planet-ward along closed field lines that require accelerating voltages of $\sim$10 kV to reach the ionosphere. Poleward of this, currents along open field lines can be carried by cool dense magnetosheath plasma that requires accelerating voltages of $\sim$0.1-1 kV \citep{08Bunceopen}, hence the majority of auroral emission is associated with particle precipitation along closed field lines. \\ The previous section concerned itself with an inverse relationship between Pedersen conductivity and field strength, with the caveat being that the number density of the Pedersen layer is the same for both hemispheres, in line with the above expectation. In Figure 3 it appears at first glance that the contrary is true because H$_3^+$ density is higher in the north, which would imply our previous argument is incomplete. However, H$_3^+$ density which peaks at an altitude of 1155 km \citep{tssras} is not wholly representative of the Pedersen layer density which itself peaks at an altitude of 1000 km \citep{2010moore,2011galand}. The ions (and their companion electrons) that create the Pedersen layer are largely hydrocarbon ions, which are dominant below $\sim$1000 km \citep{ingo2012}, and so the H$_3^+$ density is neither a proxy for the conducting layer density and does not therefore have direct implications for the previous derivation. \\ Following from the above, the H$_3^+$ temperature must differ from the Pedersen layer temperature because it is higher in altitude in a region of positive temperature gradients. However, an increase in temperature at 1000 km will lead to an increase in temperature at altitudes above it due to the vertical conduction of heat. So, in contrast to the H$_3^+$ column density, the H$_3^+$ temperature is a useful proxy for the thermal conditions within the conducting layer. \\ \subsection{Correlations between parameters} The northern auroral thermosphere exhibits an anti-correlation between temperature and column density of -0.72. This is less pronounced in the south, with a correlation coefficient of -0.52. These anti-correlations are based on small variations in these parameters - small because they remain within the errors bars of neighbouring values, such that variability is within the uncertainty, particularly for temperatures. Recent work by \citet{melinanti} shows that this is a physical phenomenon provided that the anti-correlation is outside of the range of uncertainties, as opposed to a product of the least-squares fit, i.e. as temperature increases, column density decreases and \textit{vice versa}. If the anti-correlation here is real, the physical ramification may be that increases in the density of H$_3^+$ lead to decreases in temperature, i.e. H$_3^+$ may be acting as a `thermostat' to cool the planet in a small way as it does on Jupiter and Uranus \citep{millsept06}, although recent work by \citet{ingo2012} shows such cooling plays a minor role at Saturn. In the north, there is a positive correlation coefficient of 0.54 between temperature and total emission whilst the correlation coefficient between total emission and column density is significantly weaker at 0.13, suggesting that it is changes in temperature that modulate changes in total emission. In the south, the H$_3^+$ column density and total emission are instead correlated strongly, with a coefficient of 0.76, contrasting with the north, though such correlations are the result of variability of similar size to the uncertainties involved. For both hemispheres, new parts of the main aurorae are passing by the spectrograph slit and there may also be there are large changes in the particle precipitation at local noon during this period. Both such effects have been observed in the H$_3^+$ aurora using Cassini VIMS data by \citet{badman2012a,badman2012b}. Due to the observational techniques employed here (observing only Saturn local noon), it is difficult to distinguish between these processes. \section{Conclusions} Ground based Keck NIRSPEC observations of auroral H$_3^+$ emission from Saturn have been analyzed. During the $\sim$2 hours of data obtained here, temperatures remained effectively constant within the error range achieved. At the same time, column density and total emission vary greatly over time in each hemisphere evidenced by Figure 3. This may be caused by temporal or spatial variation in the aurora, likely due to varying particle precipitation, leading to the small variations seen in all H$_3^+$ parameters. The main auroral region in the south is significantly warmer and more emissive than its northern counterpart. This asymmetry is attributed to an inversely proportional relationship between ionospheric Joule and ion drag heating and magnetic field strength. Somewhat unexpectedly, this effect outweighs the increased heating produced by seasonal enhancements in conductivity, meaning that the southern autumn aurora is 50 K warmer than that in the northern spring hemisphere. This is consistent with model predictions of a higher Pedersen conductivity in the south than the north \citep{2011galand} and an intensity asymmetry observed by Cassini VIMS pre-equinox \citep{2011badman}. A number of correlations exist between parameters that may be significant and we highlight possible causes for them. A dedicated observing campaign of Saturn's aurora is required to verify these relationships and assess the long-term behaviour of Saturn in response to seasonal variations. Although the southern aurora is unfortunately no longer viewable until at least 2023 from Earth-based telescopes, the changing viewing geometry will allow for more comprehensive studies of the northern aurora for several years. \section{Acknowledgements} The data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California, and NASA. The observations were made to support the Cassini auroral campaign in April 2011. Discussions within the international team lead by Tom Stallard on `Comparative Jovian Aeronomy' have greatly benefited this work, this was hosted by the International Space Science Institute (ISSI). We would like to thank Marina Galand for modelling assistance and related discussions. The UK Science and Technology Facilities Council (STFC) supported this work through the PhD Studentship of J.O'D. and consolidated grant support for T.S.S., S.W.H.C. and H.M. whilst S.V.B. was supported by a Royal Astronomical Society Research Fellowship.
\section{Introduction} In theoretical studies of high $T_c$ superconductors, one of the most important challenges is to extract the minimal essence of the material that leads to the strong pairing state. After the discovery of the iron-based superconductors\cite{Hosono}, the nesting between electron and hole Fermi surfaces has been considered as such an essential feature, and therefore, the identity of the family. In fact, several theoretical studies suggested a possibility of spin fluctuation mediated pairing, where the spin fluctuation arises around the nesting vector $(\pi,0)$ \cite{Mazin,Kuroki,Chubukov,Hirshfeld,Ikeda,Daghofer,Thomale,Wang}. The spin fluctuation mediates $s\pm$-wave pairing, where the gap function has $s$-wave symmetry, but its sign is reversed between the electron and hole Fermi surfaces. However, recent experiments suggest that high $T_c$ is obtained when the nesting is degraded, or even in the absence of the nesting\cite{Iimura,KFe2Se2,KFe2Se2ARPES,STO,STO2}. Then, a question of great interest is ``what is the key ingredient for high $T_c$ \textit{peculiar to the iron-based superconductors ?}'' In this context, the so-called hydrogen-doped 1111 systems, \textit{Ln}FeAsO$_{1-x}$H$_x$ (\textit{Ln} = Gd, Sm, Ce, La)\cite{Iimura} where a large amount of electrons can be doped by O~$\rightarrow$~H substitution, provide us with some important clues. For \textit{Ln}~=~La, the $T_c$ vs $x$ (doping rate) phase diagram exhibits a double dome feature, and the second dome has higher $T_c$ than the first (see Fig.~\ref{fig2}). The normal state properties above $T_c$ such as the temperature dependence of the resistivity are also different between the two domes. On the other hand, for \textit{Ln}= Sm, Ce, Gd, the phase diagram exhibits a single dome feature, and very high $T_c$'s close to or exceeding 50 K are observed. These single dome materials share commonalities with the second $T_c$ dome of LaFeAsO$_{1-x}$H$_x$\cite{Iimura}, so that understanding the origin of the second dome directly leads to the origin of the very high $T_c$ in the iron-based superconductors. One can easily expect that the Fermi surface nesting is degraded in the second dome compared to that in the first due to the large amount of doped electrons. In fact, the present study reveals that while the first $T_c$ dome originates from the spin fluctuation induced by the nesting of the Fermi surface having $d_{xz/yz}$ (and also $d_{xy}$ in some cases) orbital components, the second $T_c$ dome is due to the spin fluctuation enhanced by a peculiar motion of electrons within the $d_{xy}$, where the second neighbor diagonal hoppings are larger than the nearest neighbor ones. Such an electron motion is specific to the tetrahedral coordination of the pnictogen atoms, and we conclude that this prioritized diagonal motion is a key factor giving rise to the high $T_c$. In the single dome $T_c$ materials, ``the nesting dominating'' and the ``prioritized diagonal motion'' regimes are not well separated, and the highest $T_c\sim 50$K is attained around the crossover regime. Then, another important key ingredient for high $T_c$ is that the $d_{xy}$ and $d_{xz/yz}$ orbitals both act as driving forces of the same pairing state, namely, $s\pm$-wave pairing. Among various multiorbital systems, this is an unparalleled feature peculiar to the iron-based superconductors. In 1111 systems, electrons are doped into the FeAs layer by substituting O$^{2-}$ with F$^{-}$ or H$^{-}$. The doping actually affects the electronic band structure in two ways; i.e., the increase of positive and negative charges in the \textit{Ln}O and FeAs layers, respectively, and the reduction of the As-Fe-As bond angle. The bond angle reduction occurs linearly with doping as shown in Fig.S1 of the Supplemental Material\cite{sup}. Changing the rare earth element appears as a parallel shift of the bond angle variance against the doping rate. Quite recently, this trend has further been confirmed by partially replacing As by P in SmFeAsO$_{1-x}$H$_x$\cite{Matsuishi}, where a double dome phase diagram is found for sufficient amount of phosphorous substitution. To model these effects, band structure calculations are performed using the VASP code\cite{vasp} for hypothetical variations of LaFeAsO, where we (i) adopt the virtual crystal approximation replacing the oxygen potential by a $1-x:x$ mixture of oxygen and fluorine potentials\cite{Iimura} ($x=0.05\sim 0.5$ with an increment of $0.05$), and (ii) vary the bond angle linearly according to $x$ as $\alpha(x)=-7.48x + 114.36 +\Delta\alpha$, where $\Delta\alpha$ is the amount of parallel shift made with respect to the the actual bond angle variance of LaFeAsO$_{1-x}$H$_x$. We consider $\Delta\alpha=-3, -2, -1, 0, +1, +2^{\circ}$, as shown in Fig.S1 in the Supplemental Material\cite{sup}. Varying $\Delta\alpha$ corresponds to considering materials with different rare earth (\textit{Ln}) or anion (As partially replaced by P) elements\cite{Lee}. To capture the essence, we vary only the bond angle, while fixing the Fe-As bond length. By extracting the bands near the Fermi level using the Wannier90 package\cite{Wannier90}, we construct models consisting of $d_{xy}$, $d_{yz}$, $d_{xz}$, $d_{x^2-y^2}$, and $d_{3z^2-r^2}$ Wannier orbitals\cite{Kuroki}. Considering that the three dimensionality is not essential to the single vs. double dome issue, we omit the interlayer hoppings, and concentrate on two dimensional models in which the Brillouin zone can be unfolded to obtain a five orbital model \cite{Kuroki}. Figure~\ref{fig1} shows the Fermi surface evolution with doping for $\Delta\alpha=+1^{\circ}$ and $-1^{\circ}$. The main difference between the two cases is the presence or absence of the Fermi surface around the wave vector $(\pi, \pi)$, which originates from the $d_{xy}$ orbital\cite{KurokiPRB}. The volume of the electron Fermi surfaces around $(\pi, 0)$ and $(0, \pi)$ increases with doping, and the hole Fermi surfaces around $(0, 0)$, arising from the $d_{xz}/d_{yz}$ orbitals, shrink. On the other hand, The volume of the $d_{xy}$ hole Fermi surface around $(\pi, \pi)$ remains nearly unchanged due to the band structure variation with doping\cite{SuzukiJPSJ,Yamakawa}. In any case, the volume difference between electron and hole Fermi surfaces increases with doping, so that the nesting becomes ill-conditioned. \begin{figure}[t] \centering \includegraphics[width=9cm,clip]{fig1.eps} \caption{ Fermi surfaces for $x=0.05$, $x=0.15$, $x=0.25$, $x=0.35$ with (a) $\Delta\alpha=-1^{\circ}$ or (b) $\Delta\alpha=+1^{\circ}$. \label{fig1}} \end{figure} \begin{figure}[t] \centering \includegraphics[width=9cm,clip]{fig2.eps} \caption{ Eigenvalue of the Eliashberg equation against doping. (a) $\lambda$ against doping for $\Delta\alpha=-1$, $0$, $+1$, $+2^{\circ}$. (b) Similar as in (a) for $\Delta\alpha=-3$, $-2$, $-1^{\circ}$. (c) $\lambda$ for simplified models in which only $t_1$ is varied in conjunction with $x$ so as to maintain the volume of the $(\pi,\pi)$ Fermi surface. The numbers are the values of $t_1$ (in eV). Here, $t_2$ is fixed at 0.106 (0.113) eV for $\Delta\alpha=-1$ $(+1)$. See text for more details. Upper right panel : Experimental result of $T_c$ vs $x$ for \textit{Ln}FeAs(O,H) with \textit{Ln}=Gd, Sm, Ce, and La (from Ref.~\onlinecite{Iimura}).} \label{fig2} \end{figure} Considering intra- and interorbital electron-electron interactions on top of the five orbital band structure, we apply the fluctuation exchange (FLEX) approximation\cite{Ikeda,Bickers,Dahm} to each model, and obtain the eigenvalue of the Eliashberg equation $\lambda$ (at a fixed temperature of $T=0.005$eV), which is taken as a measure of $T_c$. We take intraorbital $U=1.3$eV, the interorbital $U'=U-2J$, Hund's coupling, and the pair hoppings $J=J'=U/6$. In a previous study, we adopted random phase approximation, where the self energy correction was neglected\cite{SuzukiJPSJ}. There, the eigenvalue of the Eliashberg equation was found to be monotonically enhanced with electron doping, which does not agree with the experimental observations. Also, the origin of the material dependence of the phase diagram was not clarified. The calculated eigenvalues of the Eliashberg equation $\lambda$ for $\Delta\alpha=-1^{\circ}\sim +2^{\circ}$ are shown in Fig.~\ref{fig2}(a). For $\Delta\alpha=-1^{\circ}$, the $\lambda$ against $x$ plot shows a ``single dome'' variance. This is already quite interesting in that the magnitude of $\lambda$ (and, hence, $T_c$) is maintained in such a large doping range. Even more interestingly, for $\Delta\alpha=0^{\circ}$, there appears a slight dip in the lightly doped regime, and this feature becomes more pronounced for $\Delta\alpha=+1^{\circ}$ and $+2^{\circ}$. Also, the maximum $T_c$ is obtained at a larger doping rate when $\Delta\alpha$ is increased. Similar results are obtained also for orbital dependent interactions (Supplemental Material\cite{sup} Fig.~S3). In Fig.~\ref{fig2}~(b), we show the doping dependence of $\lambda$ for $\alpha=-1^{\circ}\sim -3^{\circ}$. It can be seen that the increase of $\lambda$ with $x$ in the lightly doped regime becomes more rapid with decreasing $\Delta \alpha$. These results are overall in good agreement with the trend observed experimentally in Ref.~\onlinecite{Iimura} (Fig.~\ref{fig2}, upper right panel) and also Ref.~\onlinecite{Lee}. In Fig.~\ref{fig3}, we show the doping dependence of the intraorbital spin susceptibility $\chi_{xy}$ and $\chi_{xz/yz}$ within the $d_{xy}$ and $d_{xz}/d_{yz}$ orbitals\cite{Comment} for $\Delta\alpha=+1^{\circ}$ and $-1^{\circ}$. Let us first focus on $\Delta\alpha=-1$. For $x=0.05$, there appear peaks around $(\pi,0)$ and $(0,\pi)$ in both $\chi_{xy}$ and $\chi_{xz/yz}$ reflecting the Fermi surface nesting in the lightly doped system. These peak structures are suppressed by electron doping because the nesting is degraded. However, $\chi_{xy}$ is unexpectedly reenhanced beyond $x\sim 0.2$. The reason for this cannot be the Fermi surface nesting in its original sense because the nesting is monotonically degraded by doping. For $\Delta\alpha=+1^{\circ}$, on the other hand, the variance of $\chi_{xy}$ is different in that there is no enhancement in the lightly doped regime. The absence of enhanced $\chi_{xy}$ there is natural because the $d_{xy}$ hole Fermi surface around $(\pi,\pi)$ is absent for $\Delta\alpha=+1^{\circ}$, so that there is no Fermi surface nesting. Conversely, it is surprising to find an enhancement in the largely doped regime. Interestingly, an inelastic neutron scattering experiment for LaFeAsO$_{1-x}$H$_x$ actually observes the suppression of the spin fluctuation around $x\sim 0.2$ and its reenhancement in the largely doped regime\cite{IimuraPRB}. Also, comparing $\Delta \alpha=-1^{\circ}$ and $\Delta \alpha=+1^{\circ}$, the spin fluctuation grows more rapidly with doping for the former than for the latter. \begin{figure}[t] \centering \includegraphics[width=8cm,clip]{fig3.eps} \caption{ Intraorbital spin susceptibilities $\chi_{xy}$ and $\chi_{xz/yz}$ for $x=0.05$, 0.15, and 0.25. (a) $\Delta\alpha=-1^{\circ}$ and (b) $\Delta\alpha=+1^{\circ}$.} \label{fig3} \end{figure} The doping dependence of the Eliashberg equation eigenvalue $\lambda$ and the intraorbital spin fluctuations are strongly correlated, so that understanding the latter directly leads to the understanding of the former. Since the Fermi surface evolution does not seem to be correlated with the doping dependence of the spin fluctuation, we now focus on the real space hopping integrals within the $d_{xy}$ orbitals. In Fig.~\ref{fig4}, we plot the doping dependence of the nearest $(t_1)$ and the next nearest $(t_2)$ neighbor hoppings within the $d_{xy}$ orbitals for $\Delta\alpha=-1^{\circ}$ and $+1^{\circ}$. The nearest neighbor hopping $t_1$ decreases rapidly with doping, and becomes smaller than $t_2$ at a certain doping rate $x_c\sim 0.17$ for $\Delta\alpha=-1^{\circ}$ and $x_c\sim 0.28$ for $\Delta\alpha=+1^{\circ}$. We also show in Fig.~\ref{fig4}(b) the calculation result for the actual La1111 and Sm1111 using the experimentally determined lattice parameters. It is indeed seen that $x_c$ is larger for La than for Sm corresponding to the larger bond angle in the former. We will refer to this peculiar hopping relation $t_2>t_1$ as ``prioritized'' diagonal motion (or hopping) of electrons. The rapid decrease of $t_1$ by doping as compared to $t_2$ can be understood as a combined effect of (i) the increased positive charge in the LaO layer, (ii) reduction of the Fe-Fe distance, and (iii) increase of the pnictogen height, where (ii) and (iii) are the effects of the bond angle reduction. In the five orbital model, we consider Wannier orbitals, which implicitly take into account the Fe $3d$ and the hybridized As $4p$ atomic orbitals. If we consider these atomic orbitals explicitly, the present $t_1$ can be expressed as $t_1=t_1^{\rm direct}+2t_1^{\rm indirect}$, where $t_1^{\rm direct}$ and $t_1^{\rm indirect}$ are contributions from the direct hopping between Fe 3$d_{xy}$ orbitals and the indirect hopping via As $4p$, respectively, as shown in Fig.~\ref{fig4}. The two contributions have opposite signs, and $t_1^{\rm indirect}$ dominates in the lightly doped regime. This cancellation of the direct and indirect hoppings has been discussed in Refs.\cite{Miyake42622,Kotliar}. On the other hand, the next nearest neighbor $t_2$ is mainly governed by $t_2^{\rm indirect}$ because of the larger Fe-Fe distance. As electrons are doped, the energy level of the As $4p$ orbital is lowered and moves away from the Fe $3d$ level due to the effect of (i), so that the indirect hoppings decrease. The indirect contribution is also reduced because of (iii). By contrast, the direct hopping increases due to (ii). The combined effect of increasing $t_1^{\rm direct}$ and decreasing $|t_1^{\rm indirect}|$ results in a rapid decrease of $t_1$ with doping. The effect is weak for $t_2$ because it is mainly dominated by $t_2^{\rm indirect}$. As can be understood from this explanation, $x_c$ is larger for materials with larger $\Delta\alpha$. \begin{figure}[t] \centering \includegraphics[width=7cm,clip]{fig4.eps} \caption{ Upper panels: doping dependence of the real space hopping integrals $t_1$ and $t_2$ against the doping. (a) $\Delta\alpha=+1^{\circ}$ (red or gray) and $\Delta\alpha=-1^{\circ}$ (black). Solid (dashed) lines are $t_2$ ($t_1$). (b) Actual materials La1111(red or gray) and Sm1111(black). Circles (boxes) are $t_1$ ($t_2$). Lower panels : schematic figure of the origin of the prioritized diagonal hopping with doping. \label{fig4}} \end{figure} Intuitively, $t_1<(>)t_2$ corresponds to $J_1<(>)J_2$ in the limit of strong electron correlation\cite{Si,Hu} since $J_i\propto t_i^2/U$, where $U$ is the on-site intraorbital repulsion. Therefore, $t_1 < [ > 0]t_2$ is naively expected to be in favor of the $(\pi, 0)$ [$(\pi, \pi)$] spin fluctuations. More precisely, however, the enhancement of the spin fluctuation in the largely doped regime should be traced back to the band structure (not just the Fermi surface) since we are adopting FLEX, which is essentially a weak coupling approach. In fact, as shown in the Supplemental Material\cite{sup} (Fig.S2), the shape of the band changes with doping, which is mainly due to the reduction of $t_1$. The disappearance of the van Hove singularity around the wave vector $(\pi,0)$ (reminiscent of those commonly seen in the cuprates) works in favor of the $(\pi,0)/(0,\pi)$ spin fluctuations over $(\pi,\pi)$. In the models adopted above, not only $t_1$, but also other parameters vary with electron doping. To more directly single out the reduction of $t_1$ as the key factor, we have done the following analysis using simplified models. Namely, we start with five orbital models derived from a first principles band calculation performed with $\Delta\alpha = -1$ or +1, both with $15\%$ fluorine doping. Within these models, we vary the electron density in the range of $0.05 \leq x \leq 0.25$. If all the hoppings were fixed (rigid band), the hole Fermi surface would monotonically shrink as $x$ increases. As seen above, however, the Fermi surface around $(\pi,\pi)$ is almost unchanged with electron doping if the variance of the lattice parameters and the O~$\rightarrow$~F substitution effect is taken into account in the first principles calculation. To simulate this effect, in the simplified models, we vary only $t_1$ by hand in conjunction with the electron doping so that the volume of the $(\pi,\pi)$ Fermi surface is the same as that for the original model. For $\Delta\alpha=+1$, where the $(\pi,\pi)$ hole Fermi surface is absent, the energy difference between the chemical potential and the top of the $d_{xy}$ band at $(\pi,\pi)$ is kept to be the same as that in the original model. $\lambda$ calculated this way as a function of $x$ for $\Delta\alpha=\pm 1$ is shown in Fig.~\ref{fig2}~(c), where a trend similar to that in Fig.~\ref{fig2}~(a) is seen; when $t_1$ is significantly larger than $t_2$, $\lambda$ decreases with doping, while when $t_1$ is comparable to or smaller than $t_2$, $\lambda$ increases with doping. In Fig.~\ref{fig5}, we show a schematic figure of the spin fluctuation contribution to $s\pm$ superconductivity. For the $d_{xz}/d_{yz}$ orbital, there is spin fluctuation mediated pairing arising from good nesting in the lightly doped regime, which is suppressed by doping because the nesting is degraded. In the $d_{xy}$ orbital, there can be moderate Fermi surface nesting in the lightly doped regime depending on the absence or presence of the $d_{xy}$ hole Fermi surface around $(\pi, \pi)$. Therefore, for materials with small (i.e., negative) $\Delta\alpha$, the $d_{xy}$ spin fluctuation crosses over from the nesting to the prioritized diagonal motion regime. On the other hand, in materials with large (positive) $\Delta \alpha$, there is no nesting regime in the $d_{xy}$ orbital, so that the $d_{xy}$ spin fluctuation monotonically increases with doping. For materials with small bond angle, the crossover from the nesting to the prioritized diagonal motion regime occurs smoothly because $x_c$ is small. Therefore, the $T_c$ phase diagram consists of a single dome. $x_c$ is large for materials with large bond angle, so that the two regimes are separated, resulting in a double dome structure of the phase diagram. Interestingly, we have also come to realize a relation between the spin fluctuation and the resistivity, which is explained in the Supplemental Material\cite{sup} (Fig.S4). \begin{figure}[t] \centering \includegraphics[width=7cm,clip]{fig5.eps} \caption{Schematic figure of the spin fluctuation contribution to superconductivity. \label{fig5}} \end{figure} To conclude, our study has revealed the importance of the peculiar motion of electrons in the $d_{xy}$ orbital, especially in cases with very high $T_c$. Further tests for the present conclusion can be performed by examining the pressure effect. In Ref.\cite{Iimura}, it was found that applying pressure to LaFeAs(O,H) makes the double dome $T_c$ phase diagram turn into a single dome one. Our preliminary theoretical study on this problem shows that applying pressure enhances the $t_2/t_1$ ratio, and hence has an effect similar to that of replacing La by, say, Ce. Detailed analysis on this problem will be presented elsewhere. Also, it would be interesting to experimentally investigate \textit{Ln}FeAs$_{1-y}$P$_y$O$_{1-x}$H$_x$ other than \textit{Ln}=Sm\cite{Matsuishi} as another test for the present conclusion. A surprisingly interesting feature of the iron-based superconductors is that the prioritized diagonal motion in the $d_{xy}$ orbitals and the nesting within $d_{xy}$ or $d_{xz/yz}$ Fermi surfaces can all be driving forces of the $s\pm $-wave superconductivity. This coherent cooperation among various components is indeed the unparalleled identity of the iron-based superconductors. We thank H. Sakakibara, S. Onari, Y. Yamakawa, and R. Arita for valuable discussions. Numerical calculations were performed at the facilities of the Supercomputer Center, Institute for Solid State Physics, University of Tokyo. This study has been supported by Grants-in-Aid for Scientific Research No.~24340079 and No.~25009605 from the Japan Society for the Promotion of Science. The part of the research at the Tokyo Institute of Technology was supported by the JSPS FIRST Program. \section{Method} Here we describe the details of the adopted method. In the actual materials, the bond angle reduction occurs almost linearly with doping as shown as $\alpha_{\rm La}$ and $\alpha_{\rm Sm}$ in Fig.~\ref{figS1}. In our calculation, we vary the bond angle linearly according to $x$ as $\alpha(x)=-7.48x + 114.36 +\Delta\alpha$, where $\Delta\alpha$ is the amount of parallel shift made with respect to the bond angle variance of the actual La1111. First principles band calculation is performed by using the VASP package\cite{vasp}. We consider hypothetical materials LaFeAsO$_{1-x}$F$_x$, where the As-Fe-As bond angle is determined as mentioned in the main text. Values of all the other lattice parameters are those determined experimentally for LaFeAsO$_{0.92}$H$_{0.08}$. The effect of partially substituting O by F is taken into account through the virtual crystal approximation. Here, we adopt GGA-PBEsol exchange-correlation functional\cite{PBEsol}. The wave functions are expanded by plane waves up to cut-off energy of 550eV and 10$^3$ $k$-point meshes are used. Five orbital tight-binding models are derived from the first principles band calculation exploiting the maximally localized Wannier orbitals\cite{Vanderbilt}. The Wannier90 code is used for generating the Wannier orbitals\cite{Wannier90}. We neglect the hopping in the $z$ direction and concentrate on the two dimensional model for simplicity. We consider two sets of electron-electron interactions, orbital independent and dependent ones. In standard notations, the orbital independent interactions are taken as follows : the intraorbital interaction is $U=1.3$eV, the interorbital $U'=U-2J$, Hund's coupling and the pair hoppings $J=J'=U/6$. The orbital dependent interactions are obtained by multiplying a constant factor (0.55 here) to all of the interactions obtained for LaFeAsO in Ref.~\onlinecite{Miyake}. The results for the orbital independent interactions are shown in the main text, while those for the orbital dependent interactions are shown below. \begin{figure} \centering \includegraphics[width=8cm,clip]{figS1.eps} \caption{Doping dependence of the As-Fe-As bond angle. The dots are the actual experimental data.} \label{figS1} \end{figure} We apply the FLEX\cite{Bickers} method to the obtained five orbital model. In FLEX, bubble and ladder type diagrams are collected to calculate the spin $\chi^S_{ijkl}(\Vec{q},i\omega_n)$ and charge susceptibilities $\chi^C_{ijkl}(\Vec{q},i\omega_n)$, from which the self energy is calculated. Here, $i,j,k,l$ are orbital indices, and the intraorbital susceptibilities displayed in the main text are obtained as $\chi_{iiii}(\Vec{q},i\omega_n=0)$. The renormalized Green function is obtained from the Dyson equation. This process is repeated until a self consistent solution is obtained. As was done in Ref.~\onlinecite{Ikeda}, we subtract the $\omega=0$ component of the self energy for each iterations, which is considered to be taken into account already in the first principles band calculation. The resulting Green function and the susceptibilities are plugged into the linearized Eliashberg equation, whose eigenvalue $\lambda$ reaches unity at $T=T_c$, the superconducting transition temperature. Instead of going down to $T_c$, which is generally a tedious calculation, we obtain $\lambda$ at a fixed temperature ($T=0.005$eV here), and use it as a qualitative measure for $T_c$. In the FLEX calculation, we take $32\times 32$ $k$-point meshes and 4096 Matsubara frequencies. We also perform calculations for $64\times 64$ $k$-point meshes, or 8192 Matsubara frequencies for comparison for some of the parameter sets. \begin{figure*} \centering \includegraphics[width=12cm,clip]{figS2.eps} \caption{Band structure variance against doping for $\Delta \alpha=0^{\circ}$. (a) $x=0.0$, (b) $x=0.15$, (c) $x=0.35$, (d) $x=0.5$. The thickness represents the strength of the $d_{xy}$ orbital character. The $d_{xy}$ portion of the band around $(\pi,\pi)$, $(\pi,0)$, and $(0,0)$ are denoted as (i), (ii), and (iii), respectively.} \label{figS2} \end{figure*} \section{Band structure variation with doping} The band structure varies with doping as shown in Fig.~\ref{figS2}, which is mainly due to the reduction of $t_1$. One effect is that the $d_{xy}$ portion around $(\pi,\pi)$ is pushed up, which cancels out with the increase of the electron number to result in an unchanged Fermi surface. Another variation with doping occurs around the wave vector $(\pi,0)$. For small doping rate (large $t_1$), a saddle point is present in the $d_{xy}$ band around the wave vector $(\pi, 0)/(0, \pi)$, which gives a diverging density of states. Although this van Hove singularity itself is somewhat away from the Fermi level, it does affect the spin susceptibility, which is obtained by summing up contributions including those away from the Fermi level. This kind of band shape is reminiscent of the high $T_c$ cuprates, where the spin fluctuation is enhanced around $(\pi, \pi)$. On the other hand, the saddle point disappears (actually it moves away from $(\pi, 0)$) for small $t_1$, and becomes a local minimum. Such a band shape no longer enhances the $(\pi,\pi)$ spin fluctuation, and therefore has better matching with the Fermi surface configuration with holes and electrons around $(\pi, \pi)$ and $(\pi, 0)/(0, \pi)$, respectively, which itself is in favor of the $(\pi, 0)$ spin fluctuations. Since $x_c$ for $\Delta\alpha=-1$ is smaller than for $\Delta\alpha=+1$, the difference between $\sim(\pi,0)$ and $(\pi,\pi)$ spin fluctuations grows faster with doping for $\Delta\alpha=-1$, as seen in Fig.~4 of main text. The present analysis shows the importance of the prioritized diagonal motion of the electrons for the enhanced $\sim (\pi,0)$ spin fluctuation and thus $s\pm$ superconductivity. \section{Additional analysis regarding the eigenvalue of the Eliashberg equation and the spin fluctuation} \begin{figure} \centering \includegraphics[width=5cm,clip]{figS3.eps} \caption{Eigenvalue of the Eliashberg equation against doping for orbital dependent electron-electron interactions.} \label{figS3} \end{figure} We show in Fig.~\ref{figS3} the doping dependence of $\lambda$ obtained by adopting orbital dependent electron-electron interactions proportional to those evaluated in Ref.~\onlinecite{Miyake} for LaFeAsO. It can be seen that the overall trend is the same as in the case for the orbital independent interactions presented in the main text. In the main text, we have not put much focus on the variance of $\lambda$ against $\Delta \alpha$ (corresponding to the rare earth element dependence of $T_c$) for a fixed doping rate. As seen in Fig.2 of the main text (and also partially Fig.~\ref{figS3} here), $\lambda$ first increases with decreasing $\Delta\alpha$, then tends to saturate, and finally is reduced for smaller $\Delta\alpha$. This is similar to the trend observed experimentally\cite{Lee} as well as the results obtained in a previous theoretical study\cite{Usuiangle}. In Ref.~\onlinecite{Usuiangle}, the reduction of the number of hole Fermi surfaces in the small $\Delta\alpha$ regime has been proposed as the origin of the $T_c$ suppression there, but the more detailed present analysis shows that the $\lambda$ suppression occurs even when the Fermi surface multiplicity is maximized. Within the present study, the maximum $\lambda$ is attained when the $(\pi, \pi)$ $d_{xy}$ band is touching the Fermi level, i.e., when the $d_{xy}$ Fermi surface nesting appears to be ill-conditioned. In addition to $t_2 > t_1$, this condition also has to be fulfilled for the highest $T_c$. This is probably because the finite energy spin fluctuation that is responsible for pairing is enhanced when the top of the $d_{xy}$ hole band is touching the Fermi level. More detailed analysis on this issue is underway. Regarding the doping dependence of the spin fluctuation, it is interesting to note that if we turn up side down the $d_{xy}$ orbital curves in Fig.~\ref{figS4}, they are reminiscent of the $n$ vs. $x$ relation observed experimentally (right panel of Fig.~\ref{figS4})\cite{Iimura}, where $n$ is the exponent of the temperature dependence of the resistivity. This means that the exponent is probing the strength of the spin fluctuation within the $d_{xy}$ orbital. This is natural since the portion of the Fermi surface with light mass determines the transport properties, and the light portion (the portion with large group velocity) is indeed the $d_{xy}$ orbital part of the electron Fermi surface. \begin{figure} \centering \vspace{1.3em} \includegraphics[width=9cm,clip]{figS4.eps} \caption{Left panel : a figure obtained by turning upside down the upper panel of Fig.5 in the main text. Right panel : the experimentally observed doping dependence of the $n$, where $n$ is the exponent of the temperature dependence of the resistivity $\rho\sim T^n$ (taken from Ref.~\onlinecite{Iimura}).} \label{figS4} \end{figure} \section{Re-enhancement of the $\bm{d_{xz}/d_{yz}}$ spin fluctuation} In addition to the re-enhancement of the $d_{xy}$ spin fluctuation, the $d_{xz}/d_{yz}$ spin fluctuation is also re-enhanced in a even more heavily doped regime as shown in Fig.~\ref{figS5}(a). This can again be traced back to the band structure variance with doping. The $d_{xy}$ portion of the band around $(0,0)$ above the Fermi level comes down with doping, and eventually sinks below the $d_{xz}/d_{yz}$ bands, as shown in Fig.~\ref{figS2}~(a)$\sim$ (d). This is once again due to the reduction of $t_1$. A band structure reconstruction occurs, and one of the $d_{xz}/d_{yz}$ bands changes into a concave dispersion as shown schematically in Fig.~\ref{figS5}~(b)\cite{Andersen,Miyake42622,UsuiSUST}. Therefore, the density of states just above the Fermi level (the hole density of states) increases, thereby enhancing the electron-hole interaction between $\sim (0,0)$ and $\sim (\pi,0)$. Reflecting this re-enhancement of the $d_{xz}/d_{yz}$ spin fluctuations, there appears a ``kink'' in the $\lambda$ vs. $x$ plot in the heavily doped regime of $x=0.35\sim 0.45$ (Fig.2 of the main text). It should be noted, however, that this effect may be somewhat obscured in the actual materials due to the three dimensionality of the system, which is neglected here. Namely, this band structure reconstruction actually occurs at $k_z$ that varies with the doping rate\cite{Andersen,Miyake42622,UsuiSUST}. \begin{figure} \centering \includegraphics[width=9cm,clip]{figS5.eps} \caption{(a) Intraorbital spin susceptibilities $\chi_{xy}$ and $\chi_{xz/yz}$ for $x=0.35$, $\Delta\alpha=-1^{\circ}$ (b) schematic figure of the band structure variance around (0,0) with doping. } \label{figS5} \end{figure} \section{Correspondence between real and momentum spaces} In the cuprates, it is known that the nearest neighbor hopping $t_1$ dominates. This is likely to favor nearest neighbor pairing in real space, but there are two forms of nearest neighbor singlet pairing, the $d_{x^2-y^2}$ and the extended $s$-wave pairings. The former and the latter have sign reversing and conserving pair wave functions in real space, which corresponds to the gap forms $\cos(k_x)-\cos(k_y)$ and $\cos(k_x)+\cos(k_y)$, respectively, in momentum space. In the case of the cuprates, the density of states is large around the wave vector $(\pi,0)$ and $(0,\pi)$, and this is in favor of the $d_{x^2-y^2}$-wave pairing, whose gap is large around those wave vectors. On the other hand, systems with $t_2\gg t_1$ can be considered as similar to the cuprates, but with a doubled unit cell, if we neglect $t_1$. In momentum space, this gives the gap of the form $\sin(k_x)\sin(k_y)$, namely the $d_{xy}$-wave pairing, if the density of states is large around the wave vectors $(\pm\pi/2, \pm\pi/2)$, which corresponds to $(\pi,0)/(0,\pi)$ in the halved Brillouin zone (dash dotted line). In real space, $d_{xy}$ pairing is a next nearest neighbor pairing whose wave function changes sign with 90 degrees rotation. Our present study have revealed that high $T_c$ iron-based superconductors tend to have large $t_2$, but there are no Fermi surfaces around $(\pm\pi/2, \pm\pi/2)$, which is unfavorable for $d_{xy}$ pairing. Instead they have Fermi surfaces (or bands near the Fermi level) around $(0,0)$, $(\pi,0)$, $(0,\pi)$, $(\pi,\pi)$, which are the points shifted by $(\pm\pi/2, \pm\pi/2)$ from the wave vectors $(\pm\pi/2, \pm\pi/2)$. The $(\pm\pi/2, \pm\pi/2)$ shift transforms the $d_{xy}$ gap into $\sin(k_x-\pi/2)\sin(k_y-\pi/2) =\cos(k_x)\cos(k_y)$ , i.e., the $s\pm$ form. This gap corresponds to next nearest neighbor pairing in real space, which goes hand in hand with large $t_2$, but in a sign conserving form. This relation is schematically depicted in Fig.~\ref{figS6}. \begin{figure} \centering \vspace{1.3em} \includegraphics[width=9cm,clip]{figS6.eps} \caption{Correspondence of pairing in real (upper) and momentum (lower panels) spaces. The signs in the upper panels represent the sign of the pair wave function. The circled areas in the lower panels are those where the density of states near the Fermi energy is large.} \label{figS6} \end{figure}
\section{Introduction}\label{Intro} \pst Two-component Bose-Einstein condensates with linear inter-component coupling have become an object of considerable interest associated with their non-trivial dynamical properties, such as onset of self-trapping in Rabi oscillations, $\pi$-oscillations \cite{Raghavan}, bifurcations in Rabi oscillations \cite{Oberthaler1}, miscibility-immiscibility transitions induced by Rabi coupling \cite{Malomed,Oberthaler2}, to name a few. Owing to the close analogy with Josephson transitions, these two-component mixtures are often referred to as the Bose-Josephson junctions \cite{Raghavan,Gati}. They are also considered as advantageous tools for storage of quantum information \cite{Byrnes}, or for manufacturing of high-precision atomic clocks \cite{Clocks}. Usually, two ways of composing a Bose-Josephson junction are considered. The first one is based on the creation of the double-well optical potential. Then different components relate to the condensate fractions trapped inside different wells, and the linear inter-component coupling is provided by inter-well tunneling. The second way is the usage of condensates occupying the same trap but belonging to different hyperfine energy levels. The resonant microwave radiation causes transitions between the levels, providing linear inter-component coupling. In both these situations atomic ensembles are tightly confined inside the optical trap, and spatial expansion of the condensate is limited. Elimination of the BEC expansion allows one to describe the inter-component transitions by means of the classical equations of motion corresponding to a pendulum-like Hamiltonian \cite{Raghavan,Oberthaler1}. Meanwhile, it is reasonable to examine the case when the mixture of two linearly-coupled hyperfine states is loaded into the optical lattice and allowed to expand along many potential wells, that is, spatial BEC expansion cannot be eliminated. This system can be regarded as an array of Bose-Josephson junctions \cite{Science} and considered as a promising way for creating large-scale entangled states \cite{Calarco,Bloch}. However, it is well known that interplay between external (i.e. spatial) and internal (corresponding to level transitions) degrees of freedom can essentially complexify atom dynamics even in the absence of interactions between atoms. Earlier studies in the semiclassical regime for the spatial atomic dynamics revealed many intriguing phenomena like chaotic Rabi oscillations \cite{PRA01}, chaotic atomic transport and dynamical fractals \cite{JETP03,PLA03,PU06,PRA07}, occurrence of L\'evy flights and anomalous diffusion \cite{JETPL02,PRE02,JRLR06}, synchronization between spatial and internal dynamics \cite{PRA05}. A detailed theory of chaotic atomic transport of atoms in a laser standing wave has been developed in Refs.~\cite{PRA07,PRA08,EPL08}. Obviously, presence of interactions between atoms should cause additional complications of the underlying physics. In the present work we consider the BEC mixture loaded into the lattice in the presence of interactions between atoms. Our main goal is to describe the main regimes of spatial and internal dynamics of the BEC mixture, and to study how dynamics of Rabi oscillations changes with increasing the interaction strength. The paper is organized as follows. In the next section we represent the model considered and give brief description of the dynamics when tunneling between neighbouring lattice sites is negligible and there is no spatial condensate expansion. Section \ref{Spatial} describes spatial BEC dynamics when tunneling becomes significant. In section \ref{Internal} we examine how spatial dynamics influences inter-component transitions. The results obtained are discussed in Summary. \section{General theory}\label{General} \subsection{Basic equations}\label{Basic} \pst Within the mean-field approximation, the condensate dynamics is governed by the coupled Gross-Pitaevskii equations \begin{equation} \begin{aligned} i\hbar\frac{\partial\Psi_1}{\partial t}&= -\frac{\hbar^2}{2m}\frac{\partial^2\Psi_1}{\partial x^2} +U(x)\Psi_1 + g_1|\Psi_1|^2\Psi_1 + g_{12}|\Psi_2|^2\Psi_1 -\frac{\hbar\Omega}{2}\Psi_2,\\ i\hbar\frac{\partial\Psi_2}{\partial t}&= -\frac{\hbar^2}{2m}\frac{\partial^2\Psi_2}{\partial x^2} +U(x)\Psi_2 + g_2|\Psi_2|^2\Psi_2 + g_{12}|\Psi_1|^2\Psi_2 -\frac{\hbar\Omega}{2}\Psi_1 + \delta \Psi_2, \end{aligned} \label{2GP} \end{equation} where $\Psi_1$ and $\Psi_2$ are macroscopic wave functions corresponding to the first and the second hyperfine states, respectively, $U(x)$ is the lattice potential, $g_1$ and $g_2$ are the nonlinearity parameters corresponding to intra-component interaction, $g_{12}$ is the nonlinearity parameter quantifying the inter-component interaction, $\Omega$ is the Rabi frequency being the rate of the inter-component transitions, and $\delta$ is the transition detuning. Hereafter we shall consider the case of $\delta=0$. In the absence of Rabi coupling, $\Omega=0$, spatial dynamics is controlled by the lattice height and nonlinearity parameters. The lattice height determines the rate of tunneling between neighbouring sites. Strong nonlinearity can significantly reduce the wavepacket expansion due to spatial self-trapping \cite{TrSmerzi,Anker}. In addition, dynamics of two-component condensate mixtures with repulsive interactions (positive nonlinearity parameters) between atoms qualitatively depends on whether this mixture is miscible or not. The condition of immiscibility can be formulated as follows \begin{equation} g_{12}^2>g_1g_2, \label{immiscible} \end{equation} i.~e. inter-component repulsion should prevail over intra-component repulsion. Presence of the Rabi coupling can drive the mixture from miscible to immiscible regime and vice versa \cite{Oberthaler2,Malomed}. In experiments, each of the nonlinearity parameters $g_1$, $g_2$ and $g_{12}$ can be tuned via the Feshbach resonance. As long as different components correspond to the same atoms, one can assume $g_1=g_2\equiv g$. If the lattice is sufficiently deep, one can simplify the problem by means of the tight-binding approximation \cite{Ruostekoski,Kolovsky-Sib,JPB}. In this way, it is assumed that the majority of energy is preserved in the first energy band, and one can expand the wavefunctions $\Psi_1$ and $\Psi_2$ over first-band Wannier states corresponding to the decoupled problem, i.~e. $\Omega=0$. One has to take care that the tight-binding approximation becomes essentially questionable if the criterion (\ref{immiscible}) is satisfied. In the immiscible regime, initially uniform mixture undergoes fragmentation due to spatial separation of different components inside individual lattice sites \cite{Ronen,Shrestha}, that is, the wavepackets occupying individual lattice sites acquire essentially non-symmetric form and cannot be represented as superposition of the first-band Wannier states. Taking this into account, we set $g_{12}=0$ inferring the most unfavorable conditions for demixing. Thus, Eqs. (\ref{2GP}) reduce to the coupled discrete Gross-Pitaevskii equations \begin{equation} \begin{aligned} i\hbar\frac{da_n}{dt}&=-J(a_{n-1} + a_{n+1}) + g|a_n|^2a_n -\frac{\hbar\Omega}{2}b_n,\\ i\hbar\frac{db_n}{dt}&=-J(b_{n-1} + b_{n+1}) + g|b_n|^2b_n -\frac{\hbar\Omega}{2}a_n, \end{aligned} \label{TB} \end{equation} where $a_n$ and $b_n$ are complex-valued amplitudes of the condensate wave function at the lattice site $n$, $J$ is the hopping rate describing tunneling between neighbouring sites. To our knowledge, the first systematic study of such models was presented in \cite{Eilbeck}, where some particular analytical solutions were found. Similar models were considered in \cite{Mazzarella,Maksimov}. Hereafter we shall use the scaling of variables which corresponds to $\hbar=1$. \subsection{Uncoupled lattice sites}\label{Uncoupled} \pst Firstly, let's consider the case of uncoupled lattice sites, $J=0$. It corresponds to large values of the lattice potential height, when tunneling can be neglected. In this case total occupation of each site \begin{equation} \rho_n=|a_n|^2 + |b_n|^2 \label{rho_n} \end{equation} is time-independent. Substituting \begin{equation} a=|a_n|e^{i\alpha_n},\qquad b=|b_n|e^{i\beta_n} \end{equation} into (\ref{TB}), and after some simple algebra, we obtain the system of equations \begin{equation} \begin{aligned} \dot z_n=\sqrt{1-z_n^2}\sin\phi_n,\\ \dot\phi_n=\Lambda_n z_n+\frac{z_n}{\sqrt{1-z_n^2}}\cos\phi_n, \end{aligned} \label{zsys} \end{equation} where the dot denotes differentiation with respect to the rescaled time $t'=\Omega t$, \begin{equation} z_n = \frac{|a_n|^2-|b_n|^2}{\rho_n} \end{equation} is the normalized population imbalance corresponding to the $n$-th lattice site, \begin{equation} \phi_n = \arg a_nb_n^*\equiv\alpha_n-\beta_n, \label{phi_n} \end{equation} $\Lambda_n=g\rho_n/(\hbar\Omega)$. Equations (\ref{zsys}) originate from the classical Hamiltonian \begin{equation} H(z_n,\phi_n) = \frac{\Lambda_n z_n^2}{2} - \sqrt{1-z_n^2}\cos{\phi_n}. \label{Ham} \end{equation} This expression is similar to the Hamiltonian of the nonlinear pendulum. The main difference is the presence of the factor $(1-z_n^2)^{1/2}$. As this factor depends on $z_n$, the phase space structure corresponding to the Hamiltonian (\ref{Ham}) crucially depends on the parameter $\Lambda$. If $\Lambda_n<1$, i.~e. in the case of weak nonlinearity, the phase portrait corresponding to Eqs.~(\ref{zsys}) contains a family of center fixed points located at $z_n=0$, $\phi_n=\pi m$, $m=1,2$. Any trajectory of (\ref{zsys}) is rotation in phase space around one of the fixed points \cite{Oberthaler1}. It corresponds to oscillations of the population imbalance $z_n$ with zero mean. This dynamical regime can be referred to as the Rabi regime \cite{Leggett}. If $\Lambda_n>1$, the fixed point located at $z_n=0$, $\phi_n=\pi$ undergoes the pitchfork bifurcation, that is, it transforms into three fixed points, one saddle and two centers. The saddle fixed point is located at $z_n=0$, $\phi_n=\pi$, i.~e. at the same place as the center fixed point before the bifurcation. The ``new'' center fixed points are located at $z_n=\pm\sqrt{1-\Lambda_n^{-2}}$, $\phi_n=\pi$. Each of them is surrounded by the domain of population imbalance oscillations with nonzero mean, that is, one component dominates over another. This regime can be referred to as the {\it internal self-trapping}. This phenomenon was earlier observed for the cases of interacting \cite{Oberthaler1} and non-interacting \cite{JETP09} atoms. Further growth of the intercation strength $g$ moves the center fixed points towards limiting values of the population imbalance, $z_n=\pm 1$, with increasing of the phase space area corresponding to the internal self-trapping. \section{Spatial dynamics of the mixture in the optical lattice}\label{Spatial} \pst For moderate values of the optical potential height, $U_0\sim 10$ recoil energies or less, tunneling between lattice sites becomes meaningful and has to be taken into account. Tunneling results in spatial spreading of an atomic wavepacket along the lattice, thereby reducing local condensate density. However, if the interaction energy \begin{equation} E_{\text{int}} = g\sum\limits_n (|a_n|^4 + |b_n|^4) \label{Einter} \end{equation} exceeds some threshold value, there can occur spatial self-trapping, when wavepacket depletion becomes very slow due to formation of steady soliton-like states \cite{Anker}. Spatial spreading also depends on the wavepacket form, in particular, on its width and spatial modulation \cite{TrSmerzi,Politi}. All the computations presented in this paper were performed with $J=1$. In order to study the effect of spatial modulation, let's consider the initial condition of the following form: \begin{equation} a_n(t=0) = A\exp\left[-\frac{n^2}{4\sigma^2}\right]\cos{\pi pn},\quad b_n=0, \label{init} \end{equation} where $p\in [0:2]$, $-N\le n\le N$, $\sigma=10$, the factor $A$ is determined by the normalization condition \begin{displaymath} \sum\limits_{n=-N}^{N} |\psi_n|^2 = 1. \end{displaymath} Throughout this paper we perform computations for the lattice with 10001 sites, i.~e. $N=5000$. With this normalization, information about number of condensed atoms and s-wave scattering length is hidden in the nonlinearity parameter $g$. In order to distinguish different regimes of nonlinearity, it is convenient to set a value of the interaction energy $E_{\text{int}}$, and then compute the corresponding value of the nonlinearity parameter by means of the formula \begin{equation} g = \frac{E_{\text{int}}}{\sum\limits_{n=-N}^{N} |a_n(t=0)|^4}. \end{equation} In the absence of Rabi coupling, wavepacket dynamics qualitatively depends on the parameter $p$ which describes phase configuration of the initial state \cite{TrSmerzi,Politi}. The case of $p=0$ corresponds to the same wavepacket phases for all lattice sites, while the case of $p=1$ means the checkboard-like ordering of phases 0 and $\pi$. The case of $p=0.5$ also means the checkboard sequence, but the sites with phases 0 and $\pi$ are separated by unoccupied sites. If the wavepacket width is sufficiently large, $\sigma\sim 10$, the strongest spreading is expected in the absence of spatial modulation of the initial state, i.~e. for $p=0$ or $p=2$, while the strongest spatial self-trapping is anticipated for $p=0.5$ or $p=1.5$. The intermediate case of $p=1$ is the most favorable from the viewpoint of the breather formation \cite{TrSmerzi}. \begin{figure}[!htb] \centerline{ \includegraphics[width=0.4\textwidth]{Gam.eps} \includegraphics[width=0.4\textwidth]{G.eps}} \caption{Left: the map of participation ratio values computed at $t=1000\pi$. Right: the map of relative differences between the paticipation ratios computed with and without Rabi coupling. Color scales for both plots are presented below.} \label{fig-maps} \end{figure} BEC expansion along the lattice can be quantified by position variance \begin{equation} \sigma = \sqrt{\sum\limits_n n^2\rho_n - \left(\sum\limits_n n\rho_n \right)^2}, \end{equation} or by the participation ratio \begin{equation} \Gamma=\frac{1}{\sum\limits_{n} \rho_n^2}, \end{equation} where $\rho_n$ is given by (\ref{rho_n}). The participation ratio is approximately equal to the number of lattice sites which are efficiently occupied. Left panel of figure \ref{fig-maps} represents the map of participation ratio values computed at $t=1000\pi$. Coordinates of the map are the parameter $p$ and the interaction energy $E_{\text{int}}$. There is sharp contrast between ``dark'' regions corresponding to fast condensate depletion due to spatial diffusion, and ``light'' region where the condensate concentration remains sufficiently high. The light region becomes broader with increasing $E_{\text{int}}$, reflecting onset of spatial self-trapping. It involves notable vertical lines corresponding to $p=0.5,1.0$ and $1.5$. The lines at $p=0.5$ and $1.5$ correspond to smaller spreading as compared with the nearest background, while the line at $p=0$ looks a little more dark and corresponds to larger spreading. These results are qualitatively consistent with theoretical predictions of \cite{TrSmerzi}. To examine the difference brought in by the Rabi coupling, it is reasonable to compare the above results with the results obtained without the Rabi coupling. It is made in the right panel of Fig.~\ref{fig-maps}, where we plot the map of the the quantity defined as \begin{equation} \gamma = \frac{\Gamma(\Omega=1,t=1000\pi)}{\Gamma(\Omega=0,t=1000\pi)}. \label{gamma} \end{equation} If the Rabi coupling results in increasing (decreasing) of condensate spreading, then this quantity is larger (smaller) than 1. The most prominent difference is observed near the boundary separating ``dark'' and ``light'' parameter regions in the left panel. Near this boundary, the Rabi coupling slightly extends the parameter area corresponding to fast spatial diffusion. This can be understood as reduction of intra-component interactions due to component interconversion. Indeed, the interaction energy oscillates in the presence of the Rabi coupling, and its time-averaged value becomes smaller. On the other hand, one can see a cloud of dark points in the region corresponding to $p=0$ and large values of $E_{\text{int}}$. It means that, in the regime of strong nonlinearity, Rabi coupling can weaken the spreading of a wavepacket. \begin{figure}[!htb] \centerline{\includegraphics[width=0.48\textwidth]{sigma_wea.eps} \includegraphics[width=0.48\textwidth]{sigma_mod.eps} } \caption{Position variance $\sigma$ as function of time for different values of the interaction energy: a) $E_{\text{int}}=0.1$, b) $E_{\text{int}}=1$. Rabi frequency $\Omega=1$ for all cases. } \label{fig-sigma} \end{figure} \begin{figure}[!htb] \centerline{\includegraphics[width=0.48\textwidth]{Gamma_wea.eps} \includegraphics[width=0.48\textwidth]{Gamma_mod.eps} } \caption{Participation ratio as function of time for different values of the interaction energy: a) $E_{\text{int}}=0.1$, b) $E_{\text{int}}=1$. Rabi frequency $\Omega=1$ for all cases. } \label{fig-pr} \end{figure} Figures \ref{fig-sigma} and \ref{fig-pr} illustrate the process of wavepacket spreading for different values of the interaction energy. In particular, we shall consider the cases of $E_{\text{int}}=0.1$ and $E_{\text{int}}=1$, referring to them as the regime of weak and moderate intra-component interactions, respectively. We consider initial wavepackets with $p=0$, 0.5 and 1. Let's begin with considering Fig.~\ref{fig-sigma} demonstrating time dependence of the position variance. In the regime of weak intra-component interactions, $E_{\text{int}}=0.1$, the variance grows on the ballistic manner (see Fig.~\ref{fig-sigma}a). Notably, rate of the growth for the case of $p=0.5$ is remarkably larger than for $p=0$ and $p=1$. Ballistic expansion is also observed in the case of moderate intra-component interactions, $E_{\text{int}}=1$, but the difference between different values of $p$ becomes not so apparent (see Fig.~\ref{fig-sigma}b). Evolution of participation ratio is presented in Fig.~\ref{fig-pr}. One can see that number of sites occupied by the wavepacket with $p=1$ varies weakly for both regimes of nonlinearity. This can be thought of as a signature of self-trapping and breather formation. The opposite situation is observed in the case $p=0$, when to the most fastest growth of the participation ratio anticipates the strongest tendency to diffusion. In the case of $p=0.5$, fast growth of $\sigma$ in the regimes of weak or moderate intra-component interactions is accompanyied by relatively weak growth of the participation number. This indicates on the formation of ballistically moving breathers. \section{Internal dynamics}\label{Internal} \subsection{Weak intra-component interaction}\label{Weak} \begin{figure}[!htb] \centerline{\includegraphics[width=0.49\textwidth]{rabi_a_wea.eps} \includegraphics[width=0.49\textwidth]{rabi_c_wea.eps} } \caption{ Total population imbalance as function of time for the case of weak intra-component interactios, $E_{\text{int}}=0.1$. a) $p=0$, b) $p=1$. } \label{fig-z-weak} \end{figure} Let's consider how spatial dynamics of the mixture influences inter-level transitions. For this purpose, it is reasonable to examine evolution of the total population imbalance (TPI) \begin{equation} Z=\sum\limits_n \rho_nz_n = \sum\limits_n |a_n|^2 - \sum\limits_n |b_n|^2. \label{Z} \end{equation} Firstly, let's consider the case of weak intra-component interaction. Figure \ref{fig-z-weak} illustrates dependence of TPI on time for $p=0$ and $p=1$. In both these cases, TPI undergoes oscillations which are typical for the Rabi regime. In the case of $p=0$, these oscillations are almost completely coherent and their amplitude is close to the maximal value. It means that linear Rabi coupling provides complete and synchronous component interconversion. The same behavior is observed in the case of the initial state with $p=0.5$. In the case of $p=1$ the situation is qualitatively different, and the amplitude of TPI oscillations varies periodically with time, as is demonstrated in Fig.~\ref{fig-z-weak}b. This difference comes from the difference in spatial behavior. In the case of $p=0$, spatial wavepacket expansion is accompanied by fast growth of the participation ratio. Therefore, local condensate density rapidly decreases and intra-component interaction becomes negligible. This results in decoupling of translational and internal degrees of freedom, that is, all the lattice sites undergo synchronous Rabi oscillations. In the case of $p=1$, the condensate depletion is much weaker (see Fig.~\ref{fig-pr}a) due to formation of two distinct pairs of breathers. Each pair consists of identical breathers disposed symmetrically with respect to the center of the initial state. i.~e., $n=0$. Difference in condensate densities of the pairs results in the difference in frequencies of Rabi oscillations, therefore, TPI varies quasiperiodically like sum of two harmonic functions. \begin{figure}[!htb] \centerline{\includegraphics[width=0.33\textwidth]{rabi_a_mod.eps} \includegraphics[width=0.33\textwidth]{rabi_b_mod.eps} \includegraphics[width=0.33\textwidth]{rabi_c_mod.eps} } \caption{ Population imbalance as function of time for the case of moderate intra-component interactions, $E_{\text{int}}=1$. a) $p=0$, b) $p=0.5$, c) $p=1$. } \label{fig-z-mod} \end{figure} \begin{figure}[!htb] \centerline{\includegraphics[width=0.6\textwidth]{sync2.eps} } \caption{Synchronization of Rabi oscillations corresponding to the sites $n=0$ (dashes), $n=30$ (thick solid) and $n=50$ (thin solid). } \label{fig-sync2} \end{figure} \begin{figure}[!htb] \centerline{ \includegraphics[width=0.5\textwidth]{cell.eps}} \caption{Typical phase portrait describing inter-component transitions inside an individual lattice site for $p=0$. $\phi_n$ is given by (\ref{phi_n}). } \label{fig-cell} \end{figure} \begin{figure}[!htb] \centerline{\includegraphics[width=0.6\textwidth]{sync_length.eps} } \caption{ Number of synchronized sites as function of time. } \label{fig-synclength} \end{figure} \subsection{Moderate intra-component interaction}\label{Moderate} Increasing of nonlinearity in (\ref{TB}) gives rise to anharmonicity in inter-level transitions. Impact of nonlinearity depends on local density of the condensate, therefore, frequencies of Rabi oscillations at different lattice sites don't coincide. The frequency offsets destroy the synchronization of Rabi oscillations. As a consequence, the resulting TPI oscillations become the superposition of many independent oscillations running inside individual lattice sites, and the maximal amplitude of TPI oscillations is significanltly lower than in the case of weak intra-component intercation. However, if the intra-component interaction is not very strong, large fraction of the condensate near the center of the initial state can maintain synchronization, giving rise to a large domain with coherent Rabi dynamics. This regime is realized in the case of the initial state with $p=0$ (see Fig.~\ref{fig-z-mod}a). As it follows from Fig.~\ref{fig-sync2}, there occurs spontaneous synchronization of Rabi oscillations corresponding to different lattice sites. Phase portraits in the $z_n$--$\phi_n$ plane, constructed for individual lattice sites, are very similar to each over and look as is demonstrated in Fig.~\ref{fig-cell}. The marked common feature is the presence of the limit cycle whose phase space location is nearly the same for all synchronized sites. This may appear somewhat surprising. Indeed, limit cycles can occur only in the presence of dissipation while the equations (\ref{TB}) don't contain dissipative terms. However, the role of dissipation can be played by one-way energy transfer between degrees of freedom. In our case, energy from the internal degree of freedom, associated with the Rabi coupling, is absorbed by the external degrees of freedom, namely spatial wavepacket motion and intra-component interaction. Hence, there arises synchronization inherent of non-Hamiltonian systems. Taking into account that the domain of synchronized Rabi oscillations is placed in the center of the lattice, one can define the criterion of synchronization as inequality \begin{equation} |z_n(t) - z_0(t)|\le C,\quad -n'\le n\le n'. \label{sync_crit} \end{equation} Here $C$ is some small constant. We set $C=0.1$ in order to estimate number of synchronized sites \begin{equation} N_{\text{s}}=2n'+1, \end{equation} Time dependence of $N_\text{s}$ is presented in Fig.~\ref{fig-synclength}. It is demonstrated that $N_\text{s}$ undergoes pulsations with growing amplitude. It implies that some lattice sites can exit the synchronized regime and then return into it. This phenomenon indicates on the presence of wave-like excitations inside the domain of synchronization, reminiscent of magnons. Comparison with Fig.~\ref{fig-pr}b allows one to link the growth of pulsations amplitude with increasing of participation ratio, that is, more and more lattice sites become synchronized as the condensate expands along the lattice. In the case of the initial state with $p=0.5$, TPI also undergoes oscillations with decaying amplitude, as is shown in Fig.~\ref{fig-z-mod}b. However, there is no synchronization, and the amplitude decay is associated with dephasing of Rabi oscillations running at different sites. \begin{figure}[!htb] \centerline{ \includegraphics[width=0.4\textwidth]{field-p.eps}} \caption{ Squared modulo of wavefunctions corresponding to the first (solid) and second (dotted) components at $t=500\pi$. } \label{fig-field-mod} \end{figure} Initial state with $p=1$ behaves in a qualitatively different way as compared with $p=0$ and $p=0.5$. As one can see in Fig.~\ref{fig-z-mod}c, TPI undergoes oscillations with non-zero mean, i.~e. the first component of the mixture prevails over the second one. It indicates on the onset of internal self-trapping, the effect discussed in Section~\ref{Uncoupled}. In the absence of coupling between sites, internal self-trapping is linked to the bifurcation of fixed points according to the scenario described in Section \ref{Uncoupled}. However, one should expect that tunneling between neighbouring sites should destroy the phase coherence needed for the onset regular oscillations around displaced fixed points. This means that the onset of the internal self-trapping should occur when tunneling is weak. In this way, it is worth reminding that the initial state with $p=1$ is characterized by the strongest spatial self-trapping, that is, the tunneling is suppressed due to formation of localized states like breathers or solitons. These localized states are presented in Fig.~\ref{fig-field-mod}. It should be noted that they are well separated in space, that is, they don't affect each other. This means that inter-component dynamics of every such state can be fairly described within a single-site approximation corresponding to the Eqs.~(\ref{zsys}). \begin{figure}[!htb] \centerline{ \includegraphics[width=0.48\textwidth]{z_cell.eps} \includegraphics[width=0.48\textwidth]{z_cell4.eps} } \caption{ Population imbalances corresponding to the lattice sites $n=0$ (left figure) and $n=4$ (right). } \label{fig-zcells} \end{figure} Figure \ref{fig-zcells} illustrates population imbalance oscillations corresponding to the center (left figure) and right (right figure) localized states depicted in Fig.~\ref{fig-field-mod}. After short irregular transient regime, population imbalance becomes frozen in both cases, indicating the total dominance of one component. It means that the localized states can be thought of as immiscible solitons. Thus, it turns out that stable internal self-trapping occurs in the presence of spatial self-trapping. \section{Summary} \label{Summary} The present work is devoted to dynamics of two-component BEC with linear inter-component coupling loaded into the optical lattice. The main goal is to describe the interplay between spatial and internal degrees of freedom. It is found that spatial mixture dynamics is readily controlled by the interaction strength and phase configuration of the initial state, i.~e. likewise in the case of one-component condensate \cite{TrSmerzi}. If condensate undergoes fragmentation with occurrence of localized and isolated patterns like solitons or breathers, then the total population imbalance is superposition of incoherent Rabi oscillations with different amplitudes and frequencies. As the intra-component interaction energy exceeds some threshold, then Rabi oscillations corresponding to distinct localized patterns cease, that is, these patterns become immiscible solitons comprising only one component of the mixture. In this case one observes simultaneously spatial and internal self-trapping. It should be mentioned that formation of immiscible solitons is preceded by a short chaotic transient and can be regarded as some kind of self-organization. In the case of the initial state without spatial modulation, Rabi oscillations corresponding to different lattice sites can be synchronized. If the intra-component interaction is weak, the synchronization is total and the condensate undergoes Rabi oscillations as a whole. In the case of moderate intra-component interactions, Rabi oscillations at different sites have different frequencies, therefore total synchronization is destroyed. However, there can arise spontaneous synchronization inside a large lattice domain. It is shown that synchronized oscillations correspond to the same limit cycle in phase space portraits corresponding to individual lattice sites. Number of synchronized states strongly oscillates with time due the presence of wave-like excitations inside the domain of synchronization. On average, this number grows with time due to condensate expansion along the lattice. Detailed study of the particular phenomena described in this paper, like spontaneous synchronization or formation of immiscible solitons, will be the object of our future work. In addition, we intend to go beyond the approximations used in derivation of Eqs.~(\ref{TB}) and verify the results obtained with more realistic approaches. Firstly, one should use continuous mean-field approximation corresponding to the coupled Gross-Pitaevskii equations (\ref{2GP}). This allows one to estimate the role of the inter-band Landau-Zener tunneling which is ignored in the tight-binding approximation. Another important advantage of the continuous mean-field approximation is the ability to avoid limitations on the nonlinearity coefficients $g_{1}$, $g_2$ and $g_{12}$. These limitations are required for validity of the tight-binding approximation and discussed in Section \ref{Basic}. \section*{Acknowledgments} \pst The authors acknowledge the financial support provided within the RFBR project 12-02-31416, and the joint project of the Siberian and Far-Eastern Branches of the Russian Academy of Sciences (project 12-II-SO-07-022).
\section{Introduction} \IEEEPARstart{P}{article}-In-Cell (PIC) codes have been the main tool for plasma physics kinetic simulations for several decades \cite{birdsall,hockney}. In the PIC algorithm, a number of macroparticles (each representing many particles of the physical system) move through a computational grid due to the electromagnetic fields. The latter are self-consistently calculated from the particles on the grid. It is the interplay between the particles and the grid that makes PIC an efficient algorithm: the calculation of the force on the particles scales linearly with the number of particles (assuming that these are many more than the number of grid cells), as opposed to the quadratic scaling typical of grid-free molecular dynamics methods, where the particle force is calculated summing the interaction with every other particle \cite{hockney}. Traditionally, PIC codes use a uniform, Cartesian grid and integrate the particle orbit by using explicit schemes. As such, they need to resolve the shortest length scale and the fastest frequency of the plasma for numerical stability reasons. Since plasmas are inherently multiscale, it follows that PIC methods require a significant amount of computational resources to handle this disparity of scales. (We note that PIC methods with implicit time stepping, such as the implicit moment method PIC \cite{brackbill&forslund} or the fully implicit PIC methods recently developed in Refs. \cite{chacon,markidis}, avoid the stability constraints typical of explicit PIC.) The problems of PIC for multiscale plasma simulations can become worse if one needs to simulate the interaction of a plasma with material objects (for instance spacecraft, dust grains or the walls of laboratory devices), since the objects introduce additional scales in the system. As an example, consider a spacecraft at geosynchronous orbit (an altitude of approximately 36,000 km above the Earth's surface). The typical length scales are the spacecraft characteristic size $L_{spacecraft}\sim 1-10$ m, the plasma Debye length $\lambda_{D}\sim 250$ m, the electron gyroradius $\rho_e\sim 750$ m and the proton gyroradius $\rho_i\sim 30$ km. One quickly realizes that attempting to simulate such systems with a uniform, Cartesian grid explicit PIC in three dimensions, and with grid size set by the spacecraft, is unfeasible even on today's supercomputers. We note that some approaches \cite{roussel98,hutchinson2} assume Boltzmann electrons and therefore only need to resolve the ions scales. These approaches do not have the cell size constraints of the explicit, full PIC method, and are typically applicable when the material object is negatively charged and in absence of potential barriers. Focusing on spatial scales, this discussion points to the importance of introducing some kind of grid adaptivity in the PIC method, in order to handle efficiently (1) complex geometries and/or (2) the interaction of plasmas with objects whose characteristic size is much smaller than typical plasma length scales. Traditionally, PIC with non-uniform/adaptive meshes has followed two paths. One is the use of adaptive mesh refinement (AMR) techniques \cite{vay}. The other is the use of body-fitted grids, namely grids that conform exactly to a given surface \cite{westermann1} and avoid the stair-stepping typical of AMR grids. In this paper, we focus on the body-fitted PIC approach. References \cite{westermann1,westermann2} and \cite{munz} used this approach to model pulsed-power diode devices with complex geometries. In Refs. \cite{westermann1,westermann2} and \cite{munz}, a coordinate transformation (i.e. a map) from the physical space to the logical space, where the grid is uniform and Cartesian, is introduced. The quasi-stationary Maxwell equations are solved with finite differences in logical space, while particle orbits are integrated in physical space. Since the grid is distorted in physical space, a tracking algorithm to locate in which cell a particle resides must be used \cite{westermann3}. Eastwood et al. \cite{eastwood} follow the body-fitted PIC approach for modeling microwave devices. They use a finite element formulation and solve both Maxwell's equations and the particle orbits in logical space. (See also the recent work by Fichtl et al. \cite{fichtl} on an electrostatic PIC code completely designed in logical space with a particle orbit integrator that preserves phase-space area.) Wang et al. \cite{wang}, on the other hand, update Maxwell's equations in physical space but use a hybrid orbit integrator (where the particle position is in logical space and the velocity is in physical space). We also note that in the field of spacecraft-plasma interaction there are several efforts currently being developed that can be classified as body-fitted PIC. These are NASCAP-2k \cite{mandell}, SPIS \cite{roussel} and PTetra \cite{marchand}. While the specific details of the numerical implementations differ (see Refs. \cite{mandell,roussel,marchand}), in general these approaches are in the electrostatic limit and Poisson's equation is solved with an iterative solver (conjugate gradient or GMRES method) with some form of preconditioning. Furthermore, they do not introduce the logical space and all the operations are performed in physical space. Some of these approaches (SPIS and PTetra) use unstructured computational meshes. The other major code, MUSCAT \cite{muranaka}, is not body-fitted: it uses a structured, Cartesian, uniform mesh and Poisson's equation is solved with the Fast Fourier Transform (FFT) algorithm, using parallel domain decomposition on multiple processors. In this paper, we present a fully-kinetic, electrostatic, body-fitted PIC code in general curvilinear geometry called Curvilinear PIC (CPIC). While the general formulation of the method follows earlier works \cite{westermann1,eastwood,wang}, CPIC was designed with flexibility, robustness and performance considerations as the main target. For this reason, its main features are \begin{enumerate} \item The use of structured computational meshes. Modern multi- and many-core computer architectures achieve their best performance when data movement is minimized, and computations can be expressed in a data parallel fashion without indirect memory access patterns. These characteristics give a significant advantage to structured grid approaches that are able to take advantage of known direct access patterns, simpler geometric relationships, and generally lower computational complexity. For instance, the Black Box Multigrid (BoxMG) algorithm used in CPIC takes advantage of both known data access patterns and the bounded complexity of coarse-grid operators to solve general Poisson problems approximately 10 times faster than algebraic multigrid methods (AMG) \cite{maclachlan}. \item The field solver. We use the BoxMG algorithm \cite{dendy,dendy2,moulton} as solver. This algorithm is scalable (namely optimal): the computational cost to converge to the problem solution is linearly proportional to the number of grid cells. This means that, at least theoretically, there is no loss of performance when the problem size increases, as typical of unpreconditioned iterative solvers. In addition, robust variationally-based structured grid methods such as BoxMG \cite{dendy,dendy2} are able to solve problems with strongly discontinuous coefficients on the logically structured (i.e., body fitted) grids needed here. \item The particle mover. We use the hybrid mover of Ref. \cite{wang}. We have compared the performance of the hybrid mover to the (more common) physical space mover on test cases and found that the hybrid mover is typically faster and much more robust. Indeed, it avoids the issue of particle tracking needed by the physical space mover, and it also deals more efficiently with complex geometries, without the need of an extra routine to assess whether a particle crosses the domain boundaries. \end{enumerate} The paper is organized as follows. In Section II, we describe the main algorithmic and implementation aspects of CPIC. In Section III, we present some benchmark tests involving the interaction of a plasma with material boundaries. In Section IV, we draw conclusions. \section{Overview of the curvilinear PIC (CPIC)} We model a collisionless, magnetized plasma described by Vlasov's equations \begin{equation} \frac{\partial f_\alpha}{\partial t}+{\bf v} \cdot \nabla f_\alpha+\frac{q_\alpha}{m_\alpha}\left({\bf E}+{\bf v} \times {\bf B}_0 \right)\cdot \nabla_{v} f_\alpha=0 \label{f} \end{equation} where the subscript $\alpha$ refers to the plasma species, $\alpha=e,\,(i)$ for electrons (ions). In Eq. (\ref{f}), $f_\alpha({\bf x}, {\bf v}, t)$ is the plasma distribution function, ${\bf E}=-\nabla \phi$ is the electric field ($\phi$ is the electrostatic potential), ${\bf B}_0({\bf x})$ is the magnetic field, while $q_\alpha$ and $m_\alpha$ are the charge and the mass of the plasma particles. We focus on the electrostatic limit, therefore Eq. (\ref{f}) is coupled through the electrostatic potential via Poisson's equation \begin{equation} \nabla^2 \phi =-\frac{\sum q_\alpha n_\alpha}{\varepsilon_0}=-\frac{\rho}{\varepsilon_0}, \label{pois} \end{equation} where the plasma densities are given by \begin{equation} n_\alpha=\int f_\alpha d {\bf v}, \label{dens} \end{equation} and the magnetic field ${\bf B}_0$ does not evolve in time. In Eq. (\ref{pois}), $\varepsilon_0$ is the permittivity of vacuum. The PIC method solves Eq. (\ref{f}) numerically by introducing macroparticles, each representing a large number of actual particles of the system, and following their characteristics: \begin{eqnarray} && \frac{d {\bf x}_p}{dt}={\bf v}_p, \label{xp} \\ && m_p \frac{d {\bf v}_p}{dt}=q_p\left[{\bf E}\left({\bf x}_p \right)+{\bf v}_p \times {\bf B}_0\left({\bf x}_p \right) \right], \label{vp} \end{eqnarray} where ${\bf x}_p$ and ${\bf v}_p$ are position and velocity of each particle. Hence, a standard electrostatic PIC cycle, repeated at each time step of the simulation, consists of the following four steps: \begin{enumerate} \item \underline{Particle mover}: particles are moved according to Eqs. (4) and (5), given the electromagnetic fields. \item \underline{Particle-to-Grid (P$\rightarrow$G) interpolation}: the charge carried by the particles is accumulated on the computational grid via interpolation, to obtain the plasma charge density. \item \underline{Field solver}: Poisson's equation is solved on the computational grid, given the charge density. \item \underline{Grid-to-Particle (G$\rightarrow$P) interpolation}: the electromagnetic fields are interpolated to the particle position. \end{enumerate} We now proceed to discuss each of these steps as they are treated in CPIC. \subsection{Curvilinear formulation}\label{curv} First, we introduce a coordinate transformation \begin{equation} {\bf x}={\bf x}\left(\mbox{\boldmath$\xi$} \right) \label{map} \end{equation} between the physical space $X$ with coordinates ${\bf x}=\left(x^1,\,x^2,\,x^3 \right)$, and the logical space $\Xi$ (the unit cube), with coordinates $\mbox{\boldmath$\xi$}=\left(\xi^1,\,\xi^2,\,\xi^3 \right)=\left(\xi,\,\eta,\,\zeta \right)$. Given a computational mesh in physical space corresponding to the simulation domain, Eq. (\ref{map}) maps it to the unit cube in logical space, where the computational mesh is uniform and Cartesian. We define the Jacobi matrix as \cite{liseikin} \begin{equation} j_{\alpha\beta} \left(\mbox{\boldmath$\xi$}\right)=\frac{\partial x^\alpha}{\partial \xi^\beta},\,\,\,\alpha,\,\beta=1,\,2,\,3, \label{jcov} \end{equation} while its inverse is \begin{equation} k^{\alpha\beta} \left({\bf x}\right)=\frac{\partial \xi^\alpha}{\partial x^\beta},\,\,\,\alpha,\,\beta=1,\,2,\,3. \label{jcon} \end{equation} The Jacobian of the transformation is the determinant of the Jacobi matrix, \begin{equation} J\left(\mbox{\boldmath$\xi$}\right)={\rm det}\left[j_{\alpha\beta} \right]. \label{jac} \end{equation} Similarly, the covariant metric tensor is defined as \cite{liseikin} \begin{equation} g_{\alpha\beta}\left(\mbox{\boldmath$\xi$}\right)=\frac{\partial x^\gamma}{\partial \xi^\alpha}\frac{\partial x^\gamma}{\partial \xi^\beta},\,\,\,\alpha,\,\beta=1,\,2,\,3, \label{gcov} \end{equation} while the contravariant metric tensor is \begin{equation} g^{\alpha\beta}\left({\bf x}\right)=\frac{\partial \xi^\alpha}{\partial x^\gamma}\frac{\partial \xi^\beta}{\partial x^\gamma},\,\,\,\alpha,\,\beta=1,\,2,\,3. \label{gcon} \end{equation} Here and everywhere else in the text, repeated indices imply summation (Einstein's notation). Note that the metric tensor matrix is symmetric and that grids are orthogonal if $g_{12}=g_{13}=g_{23}=0$. With these quantities, we can express all the metric elements from physical space to logical space. For instance, Poisson's equation in logical space becomes \begin{equation} \nabla^2 \phi=\frac{1}{\Omega J} \frac{\partial}{\partial \xi^i}\left(\Omega J g^{ij} \frac{\partial \phi}{\partial \xi^j} \right)=-\frac{\rho}{\varepsilon_0}, \label{poiscurv} \end{equation} where we have introduced a geometric factor $\Omega$ that can be used when the geometry of the physical space is non-Cartesian. For instance, for Cartesian geometry we have $\left(x^1=x,\,x^2=y,\,x^3=z\right)$ and $\Omega=1$, while for cylindrical geometry we have $\left(x^1=r,\,x^2=\theta,\,x^3=z\right)$ and $\Omega=r$. From the discussion so far, it is clear that the construction of the map from physical to logical space, Eq. (\ref{map}), is a critical step in CPIC. In the case of extremely simple geometries, like the test examples considered in this paper, one can specify such map analytically. Alternatively, one can use a suitable mesh generator. For cases that are still relatively simple, where the map is characterized by a single or few structured blocks, we normally use methods based on the solution of a set of partial differential equations such as Winslow's method \cite{winslow-unpublished-var-diff} or the optimal distortion method \cite{delzanno08}. On the other hand, in order to capture the full complexity of a spacecraft we plan to use commercially available software for structured meshes, typically developed in the computational fluid dynamics community (see for instance GridPro, https:/gridpro.com). The resulting computational grid is block-structured and, for complex geometries, might involve hundreds of blocks. This will require upgrading CPIC to handle block-structured meshes. However, the fact that each block maps to a logically rectangular Cartesian mesh ensures that the localization of the particles on the mesh of each block and the particle mover remain essentially unchanged, while there is an additional computational cost in assigning each particle to a block. In CPIC this last step is trivial provided that each grid block stores the maximum and minimum value of its logical coordinates (as done in the domain decomposition for the parallelization), since the particle position is in logical space. We also note that in CPIC the metric coefficients (the contravariant metric tensor, the inverse of the Jacobi matrix and the Jacobian) are computed from finite difference approximation given the discrete map (\ref{map}) known at cell vertices (see for instance \cite{delzanno08}). There is no conceptual difference or additional computational cost between a structured or block-structured mesh since the metric coefficients are defined in each block. Furthermore, there are three popular approaches to developing a Poisson solver for block-structured grids. First, one may use a domain decomposition approach, with BoxMG as the sub-domain solver. This shares convergence properties of domain decomposition methods, and for large problems would require overlap to obtain satisfactory scaling. Second, BoxMG could be extended following the approach used in the HYPRE libraries Semi-Structured Grid interface to its parallel Semi-Coarsening Multigrid solver, which manages the coarsening of each block independently, while maintaining the full connectivity of the global problem \cite{hypre}. Finally, a bounding box approach can be used to include the entire global domain in a single BoxMG solve, with internal boundary conditions used at the block interfaces, and this is the approach that we plan to adopt. \subsection{Computational aspects of CPIC} In CPIC quantities are expressed in normalized units: lengths are normalized to the electron Debye length, velocities to the electron thermal velocity, time to the electron plasma frequency, the electrostatic potential to the electron temperature, densities to a reference density and the magnetic field to a reference magnetic field. In the tests presented throughout the paper, we will always use normalized quantities. CPIC is designed in a staggered formulation, where the electrostatic potential and the density are at cell centers (labeled with subscript c) and the electric field is at vertices (v). The staggered formulation and the fact that the assignment function for the interpolation from particles to the grid is one order higher than the one for the interpolation from the grid to the particles (see below) imply that CPIC is an energy-conserving PIC as defined in Ref. \cite{birdsall}. We have developed and successfully tested a two-dimensional (2D) and three-dimensional (3D) version of CPIC. CPIC2D is fully parallelized with domain decomposition and the MPI library to handle the communication among processors. Parallelization of CPIC3D is ongoing. \subsubsection{P$\rightarrow$G interpolation} The accumulation of the particle charge on the grid is given by \begin{equation} n_{\alpha,\,c}=\frac{\sum_p q_p W_{pc} \left(\mbox{\boldmath$\xi$}_p-\mbox{\boldmath$\xi$}_c \right)}{J_c \Delta \xi \Delta \eta \Delta \zeta} \label{pg} \end{equation} where the assignment function is \begin{equation} W_{pc}=b_2\left(\xi_p-\xi_c\right)b_2\left(\eta_p-\eta_c\right)b_2\left(\zeta_p-\zeta_c\right) \label{Wpc} \end{equation} and $b_2$ is the b-spline of order 2: \begin{equation} b_2\left(\xi \right)=\left\lbrace \begin{aligned} &\frac{3}{4}-\xi^2, &|\xi|<\frac{1}{2},\\ &\left(\frac{3}{2}-|\xi|\right)^2, &\frac{1}{2}<|\xi|<\frac{3}{2},\\ & 0, & {\rm otherwise}. \end{aligned} \right. \label{b2} \end{equation} In Eq. (\ref{pg}), $\sum_p$ implies summation over all the particles. \subsubsection{G$\rightarrow$P interpolation} The interpolation of the force field from the computational mesh to the particle position is given by \begin{equation} {\bf E}_p=\sum_v {\bf E}_v W_{vp} \left(\mbox{\boldmath$\xi$}_v-\mbox{\boldmath$\xi$}_p \right) \label{vvp} \end{equation} with the assignment function \begin{equation} W_{vp}=b_1\left(\xi_v-\xi_p\right)b_1\left(\eta_v-\eta_p\right)b_1\left(\zeta_v-\zeta_p\right) \label{Wvp} \end{equation} given as the product of b-splines of order 1 \begin{equation} b_1\left(\xi \right)=\left\lbrace \begin{aligned} &1-|\xi|, &|\xi|<1,\\ & 0, & {\rm otherwise}. \end{aligned} \right. \label{b1} \end{equation} We note that the interpolation operations are performed in logical space and therefore are equivalent to what is done in the standard PIC with uniform grid. The same approach is used in Ref. \cite{lapenta}. \subsubsection{Field solver} We solve Poisson's equation on the logical grid. We use a second-order, conservative discretization and impose boundary conditions via ghost cells. The resulting linear system is solved with a robust variational multigrid method \cite{dendy}. The need for a scalable algorithm motivates the use of a multigrid solver, as it is one of very few truly scalable solvers (it can be shown that the order of complexity of the method is $O(N)$, with $N$ the number of unknowns of the linear system \cite{Briggs}). Moreover, efficient parallel implementations are readily available. Multigrid methods are iterative, and achieve their efficiency through the recursive use of successively coarser discrete problems (i.e., a sequence of coarse-grid discrete operators) in conjunction with smoothing on each level (e.g., a single Gauss-Seidel iteration on each level) to damp the highly oscillatory errors associated with each grid \cite{Briggs}. Early work on multigrid began with structured grids and took a natural but simple geometric viewpoint: coarsening removed every other point in each coordinate direction; the same discretization was applied directly at each grid resolution; and the interpolation and restriction operators were defined based on the grid geometry alone. Unfortunately, this approach was not robust for challenging problems with discontinuous coefficients or severely distorted grids. To address this shortcoming, a robust approach for problems on logically structured grids was developed that only requires the user to provide the fine-grid matrix and the right-hand side \cite{dendy}. This Black Box Multigrid (BoxMG) algorithm uses a variational coarse-grid operator, as it minimizes the error in the range of the interpolation. Moreover, it uses the entries in the matrix (the discrete operator) to define the interpolation, dubbed operator-induced interpolation. This approach ensures that important properties of the fine-scale PDE are well approximated at all resolutions, and leads to a robust scalable solver suitable for grids of any dimension, discontinuous coefficients, and any type of boundary condition (Dirichlet, Neumann, Robin, and periodic) \cite{moulton}. In addition, it is robust for anisotropic problems, provided line-relaxation is used instead of point smoothing. As an illustration of this capability, we use the method of manufactured solutions \cite{roache} on the (normalized) Poisson equation in spherical geometry in two-dimensions (2D). This can be done by assuming that the physical space corresponds to the cylindrical geometry $\left(x^1=r,\,x^2=\theta,\,x^3=z\right)$, with $\partial/\partial \theta=0$, and introducing the following coordinate transformation \begin{eqnarray} && r=\left[r_1+\left(r_2-r_1 \right)\xi \right]\sin \left[\pi\left(1-\eta\right) \right], \label{r} \\ && z=\left[r_1+\left(r_2-r_1 \right)\xi \right]\cos \left[\pi\left(1-\eta\right) \right], \label{z} \end{eqnarray} where $r_1$ and $r_2$ are the inner and outer radii of the physical domain (we choose $r_1=1$ and $r_2=10$ in this example). The geometric factor is set to $\Omega=r$. Next, we seek the following solution \begin{equation} \phi\left(\xi,\,\eta \right)=\sin\left[\left(r_2-r_1\right)^2 \xi \left(1-\xi\right) \right]\cos \left(\pi \eta^2 \right), \label{mms} \end{equation} which is inserted in Poisson's equation to obtain the density. We then solve Poisson's equation with the BoxMG solver on a serial machine. We use line-relaxation since this example is fairly anisotropic in terms of metric tensor coefficients. The results are presented in Table I, in the form of a convergence study changing the number of grid points. The second column of Table \ref{t1} shows the $L_2$ norm of the error of the numerical solution relative to the analytical solution in Eq. (\ref{mms}), while the third column is the inverse of the ratio of the error calculated on a given grid divided by the error from the previous coarser grid. The second-order convergence of our discretization scheme is clearly recovered (i.e. when doubling the resolution, the error ratio is equal to 4 for sufficiently large grids). The fourth column shows the number of iterations needed by the solver to converge, with relative tolerance set to $r_{tol}=10^{-10}$. The iteration count remains practically constant. The fifth and sixth columns show the time to reach convergence and the ratio of such time on a given grid relative to the previous coarser grid. Here we have only plotted the time for the more refined grids in order to identify the asymptotic scaling. As expected, it shows that the algorithm is scalable: doubling the grid resolution in each direction requires about four times more time for convergence. This is also consistent with the iteration count remaining constant as the grid is refined. \begin{table}[!t] \caption{Convergence study of the BoxMG solver changing the number of grid points.} \label{t1} \centering \begin{tabular}{|c|c|c|c|c|c|} \hline Grid & $L_2$ error & Error & Number of & Time [s] & Time \\ & & ratio & iterations & & ratio \\ \hline \hline $16^2$ & $5.3 \cdot 10^{0}$ & & $5$ & & \\ \hline $32^2$ & $2.0 \cdot 10^{-1}$ & $28.6$ & $5$ & & \\ \hline $64^2$ & $3.7 \cdot 10^{-2}$ & $5.1$ & $5$ & & \\ \hline $128^2$ & $8.8 \cdot 10^{-3}$ & $4.1$ & $5$ & $0.06$ & \\ \hline $256^2$ & $2.2 \cdot 10^{-3}$ & $4.0$ & $6$ & $0.27$ & $4.5$ \\ \hline $512^2$ & $5.4 \cdot 10^{-4}$ & $4.0$ & $6$ & $1.2$ & $4.4$ \\ \hline $1024^2$ & $1.3 \cdot 10^{-4}$ & $4.0$ & $5$ & $4.2$ & $3.4$ \\ \hline $2048^2$ & $3.4 \cdot 10^{-5}$ & $4.0$ & $5$ & $16.6$ & $4.0$ \\ \hline $4096^2$ & $8.4 \cdot 10^{-6}$ & $4.0$ & $5$ & $66.4$ & $4.0$ \\ \hline \hline \end{tabular} \end{table} \subsubsection{Particle mover} We initially specialize the discussion to the case of Cartesian physical space. The most common approach for particle mover in PIC codes with non-orthogonal grids is to move particles in physical space via Eqs. (\ref{xp}) and (\ref{vp}). As the particles move through a distorted mesh, a tracking/localization procedure is required in order to locate the particle in a cell, as needed by the interpolation and accumulation routines. (An alternative, used in Ref. \cite{lapenta}, is to move particles in physical space and then invert the map ${\bf x}\left(\mbox{\boldmath$\xi$}\right)$ to obtain the particles' position in logical space. This technique can be very efficient for cases where the map can be inverted analytically.) In addition, the time discretization of Eqs. (\ref{xp}) and (\ref{vp}) is usually explicit, with the leap-frog integrator \cite{birdsall} as the most popular choice. The leap-frog algorithm staggers particles position and velocity by half time step, and is second-order accurate in time. In CPIC, however, we follow a different approach which was proposed in Ref. \cite{wang}: particles retain their physical space velocity but are characterized by their logical space position (and therefore move in logical space). In other words, our mover consists of Eq. (\ref{vp}), while Eq. (\ref{xp}) is replaced by its logical space equivalent (obtained by projecting Eq. (\ref{xp}) on the contravariant base vector): \begin{equation} \frac{d \xi_p^i}{dt}=k^{ij}v_p^j,\,\,\,i=1,\,2,\,3, \label{hybridx} \end{equation} where the inverse of the Jacobi matrix is given by Eq. (\ref{jcon}) and we have used index notation. We refer to Eqs. (\ref{vp}) and (\ref{hybridx}) as the hybrid mover. The hybrid mover offers some clear advantages over a physical space mover, mainly because of its simplicity: \begin{enumerate} \item There is no need of a tracking algorithm to locate the particles: a particle is immediately assigned to a cell through its logical space coordinate. \item While tracking algorithms applied to complex boundaries (multiply connected or with concave boundaries) are typically not robust, this is not the case for the hybrid mover. In fact, a particle trivially leaves the domain if one of its logical coordinates becomes less than zero or greater than one. \end{enumerate} There is, however, an important disadvantage of the hybrid mover relative to its physical space counterpart: a simple leap-frog integrator loses second order accuracy on a non-orthogonal mesh. The simplest way to fix this problem and regain second order accuracy is to use a leap-frog integrator modified with a predictor-corrector approach \cite{wang}. The algorithm is as follows (index $n$ refers to the time level, $\Delta t$ is the time step): \begin{enumerate} \item Perform a full-step update of the particle velocity to obtain ${\bf v}_p^{n+1/2}$: \begin{equation} \frac{{\bf v}_p^{n+1/2}-{\bf v}_p^{n-1/2}}{\Delta t}=\frac{q_p}{m_p}\left[{\bf E}\left(\mbox{\boldmath$\xi$}_p^n \right) +{\bf v}_p^n\times {\bf B}_0 \left(\mbox{\boldmath$\xi$}_p^n \right) \right], \label{step1} \end{equation} with \begin{equation} {\bf v}_p^n=\frac{{\bf v}_p^{n+1/2}+{\bf v}_p^{n-1/2}}{2}. \label{step1a} \end{equation} This step is performed explicitly with the algorithm introduced by Boris \cite{boris} which separates the velocity update due to the electric field from the update due to the magnetic field. \item Predictor: perform a half-step update of the particle logical position to obtain $\mbox{\boldmath$\xi$}_p^\prime$: \begin{equation} \frac{\xi_p^{i\prime}-\xi_p^{i,n}}{\displaystyle{\frac{\Delta t}{2}}}=k^{ij}\left(\mbox{\boldmath$\xi$}_p^n \right)v_p^{j,n+1/2},\,\,\,i=1,\,2,\,3. \label{step2} \end{equation} \item Re-evaluate the metric coefficients at $\mbox{\boldmath$\xi$}_p^\prime$ via interpolation: \begin{equation} k^{ij}\left(\mbox{\boldmath$\xi$}_p^\prime\right)=\sum_v k^{ij}\left(\mbox{\boldmath$\xi$}_v \right)W_{vp} \left(\mbox{\boldmath$\xi$}_v-\mbox{\boldmath$\xi$}_p^\prime \right). \label{step3} \end{equation} \item Corrector: perform a full-step update of the particle logical position with the new metric coefficients, to obtain $\mbox{\boldmath$\xi$}_p^{n+1}$: \begin{equation} \frac{\xi_p^{i,n+1}-\xi_p^{i,n}}{\Delta t}=k^{ij}\left(\mbox{\boldmath$\xi$}_p^\prime \right)v_p^{j,n+1/2},\,\,\,i=1,\,2,\,3. \label{step4} \end{equation} \end{enumerate} With these considerations, an important question is whether the hybrid mover is more or less efficient relative to the physical space mover from a performance standpoint. To answer this question, we have compared the hybrid mover with a physical space mover for the following test. We consider a square domain in 2D (Cartesian geometry), with the following physical-to-logical space map \begin{eqnarray} &&x=\xi+\varepsilon \sin\left(2 \pi \xi\right)\sin\left(2 \pi \eta\right), \nonumber \\ &&y=\eta+\varepsilon \sin\left(2 \pi \xi\right)\sin\left(2 \pi \eta\right). \label{map2} \end{eqnarray} The parameter $\varepsilon$ controls the level of distortion of the grid. We fix the electromagnetic fields on the grid, the mesh size $\left(64 \times 64\right)$, the number of particles per cell $N_{p,\,cell}=100$ and we move the particles for $10^4$ time steps. The boundary conditions on the particles are periodic on all sides. The physical space mover requires a particle tracking and localization technique and we use the method proposed in Ref. \cite{westermann3}. The results of our comparison changing time step are presented in Tables \ref{t2} and \ref{t3} for orthogonal (uniform) $\left(\varepsilon=0\right)$ and non-orthogonal $\left(\varepsilon=0.1\right)$ grids. For the latter, the ratio of the largest to smallest cell size is $4.4$. Tables \ref{t2} and \ref{t3} show the average time per particle per time step to complete the simulation, normalized to the time required by the hybrid mover on the uniform mesh with $\Delta t=0.001$. Let us focus first on Table \ref{t2}. We see that the time required by both movers to complete the simulation increases with the time step. For the hybrid mover, this is because additional operations are performed due to particles crossing the domain boundaries. For the physical space mover, however, there is extra work associated with the tracking/localization algorithm: with larger time steps, particles move further away and there are more particles crossing cell boundaries within the domain. Thus, for realistic time steps used in kinetic simulations, even on a uniform grid the hybrid mover can be faster. Let us now look at the results on the non-uniform grid, Table \ref{t3}. We notice that the average time required by the hybrid mover is comparable to the one on the uniform grid. This result is not surprising since, given the metric coefficients, the hybrid mover treats all the grids equally. It indicates, however, that the hybrid mover is extremely robust. The physical space mover, on the other hand, requires more time relative to its performance on the uniform grid. Again, this is not surprising as now the tracking/localization algorithm has to deal with distorted meshes. On this example, with time step $\Delta t=0.05$ the hybrid mover is about $60\%$ faster. We note that the results presented in Tables \ref{t2} and \ref{t3} involve the time spent by the algorithm to move and localize the particles and the time spent to apply the (periodic) boundary conditions when a particle crosses the domain boundaries. The latter time is the same for the hybrid and physical space movers since the example considered consists of a simple square domain. Therefore the performance loss of the physical space mover is due to the particle tracking/localization only. A further point worth noting is that particle tracking/localization algorithms are notoriously less robust on complex boundaries, and this will make the comparison even more favorable to the hybrid mover. \begin{table} \caption{Comparison of the time (per particle per time step) required by the hybrid mover and the physical space mover on a orthogonal (uniform) grid for various time steps.} \label{t2} \centering \begin{tabular}{|c|c|c|} \hline Time step & Hybrid mover: & Physical space mover: \\ & time & time \\ \hline \hline $0.001$ & $1.00$ & $0.80$ \\ \hline $0.002$ & $1.02$ & $0.86$ \\ \hline $0.004$ & $1.08$ & $1.05$ \\ \hline $0.01$ & $1.17$ & $1.31$ \\ \hline $0.025$ & $1.17$ & $1.31$ \\ \hline $0.05$ & $1.17$ & $1.44$ \\ \hline \hline \end{tabular} \end{table} \begin{table} \caption{Comparison of the time (per particle per time step) required by the hybrid mover and the physical space mover on a non-orthogonal grid $\left(\varepsilon=0.1\right)$ for various time steps.} \label{t3} \centering \begin{tabular}{|c|c|c|} \hline Time step & Hybrid mover: & Physical space mover: \\ & time & time \\ \hline \hline $0.001$ & $1.04$ & $1.04$ \\ \hline $0.002$ & $1.06$ & $1.14$ \\ \hline $0.004$ & $1.10$ & $1.31$ \\ \hline $0.01$ & $1.16$ & $1.71$ \\ \hline $0.025$ & $1.16$ & $1.71$ \\ \hline $0.05$ & $1.17$ & $1.84$ \\ \hline \hline \end{tabular} \end{table} One final comment is in order with regard to the application of CPIC to non-Cartesian geometry. This is useful for instance when one wants to simulate the interaction of a plasma and an object in a spherically symmetric system, and we will show some of these examples in Sec. \ref{tests}. In this case one can run CPIC in 2D cylindrical geometry and impose azimuthal symmetry, thus allowing a much higher resolution than in a fully 3D simulation. From the perspective of the solver, this implies setting $\Omega=r$ and $\partial/\partial \theta=0$ in Poisson's equation. For the mover, on the other hand, one must take into account the contribution of the inertial forces to the particle motion. This can be done with the cylindrical version of the Boris mover \cite{boris}, suitably modified in its hybrid mover version. In addition, for simulations involving a uniform background magnetic field, this algorithm can be improved to resolve the particle gyromotion exactly \cite{patacchini,delzanno_mover}. Its importance resides in the fact that, by describing the gyromotion exactly, one can relax the time step constraints that a PIC code with a standard leap-frog mover would have when simulating a strongly magnetized plasma with $\omega_{ce}\gg\omega_{pe}$ (with $\omega_{ce}$ the electron cyclotron frequency) and deliver a significant computational savings. \section{Tests}\label{tests} We have successfully performed several tests to check the validity of CPIC against known solutions. These tests include Langmuir waves, Landau damping and two-stream instabilities, for cases where the plasma is not in contact with a material wall. Here, we present some additional tests where the plasma is bounded by a material boundary. These tests are performed in 2D (for the fields, while the particles retain three components of the velocity), by exploiting some symmetry of the system. \subsection{Charging and shielding of a conducting planar wall in an unmagnetized plasma} First, we study the charging and shielding of a perfectly conducting planar wall in an unmagnetized plasma consisting of electrons and singly charged ions. We consider a rectangular domain $L_x\times L_y=20 \times 5$ (in units of electron Debye length) in Cartesian geometry and use a uniform mesh: \begin{eqnarray} &&x=L_x \xi, \nonumber \\ &&y=L_y \eta. \label{map3} \end{eqnarray} The wall is located at the left boundary, $\xi=0$, and absorbs plasma. The computational domain is initially empty. At each time step we inject plasma particles from the right boundary $\left(\xi=1 \right)$. The injection fluxes are calculated according to a Maxwellian at rest for the electrons and a drifting Maxwellian (with drift velocity $v_{di}=-2.1$, normalized to the ion thermal speed $v_{thi}=\sqrt{T_i/m_i}$ with $T_i$ the ion temperature) for the ions. The boundary conditions for the field solver are periodic at the bottom $\left(\eta=0 \right)$ and top $\left(\eta=1 \right)$ boundaries, $\phi=0$ at the right boundary, and Gauss' law is applied at the wall. Particles can be absorbed by the wall (namely they are removed from the simulation and their charge is accumulated to the wall), can leave the system at the right boundary, while they re-enter at the bottom (top) if they cross the top (bottom) boundary. Other parameters of the simulation are $\omega_{pe}\Delta t=0.05$ (with $\omega_{pe}$ the electron plasma frequency), the temperature ratio of the injected plasma is $T_e/T_i=1$ and the mass ratio is $m_i/m_e=1836$. Given the plasma fluxes at the right boundary, it is possible to calculate the (one-dimensional) steady-state solution of the system. The theory is based on particle and energy conservation. It requires the solution of a nonlinear Poisson equation, where the plasma densities are obtained analytically as function of the electrostatic potential. We do not report these expressions here for brevity. A similar theory, which does not include the ion drift velocity, can be found in Ref. \cite{schwager}. We refer to this procedure as the analytic solution and use it to benchmark against CPIC. Figure \ref{fig1} shows the time evolution of the wall potential (normalized to the electron temperature) obtained with CPIC for a $128\times128$ mesh. The computational domain is initially empty. As the electrons move towards the wall much faster than the ions, at the beginning there is a strong negative spike in the potential: $\phi_{wall}\left(\omega_{pe}t\right)/T_e\simeq-5.7$. Slowly, the ion distribution function adjusts to its equilibrium value and a steady-state is reached. At equilibrium, the wall potential (averaged over the last sixth of the simulation) is $\phi_{wall}/T_e\simeq-2.11$, in good agreement with the one from the analytic theory $\phi_{wall}^{analytic}/T_e\simeq-2.10$. Figure \ref{fig2} shows the shielding potential at the end of the simulation. As expected, the structure of the potential is one-dimensional (aside from some fluctuations in the region where the potential is very small). A comparison of the equilibrium shielding potential with the analytic solution (not shown) indicates that the screening of the wall by the plasma is captured correctly in the simulation. \begin{figure} \centering \includegraphics[width=3.3in]{proc_fig1-eps-converted-to.pdf} \caption{Test A: time evolution of the wall potential.} \label{fig1} \end{figure} \begin{figure} \centering \includegraphics[width=4in]{proc_fig2-eps-converted-to.pdf} \caption{Test A: equilibrium shielding potential at the end of the simulation.} \label{fig2} \end{figure} \subsection{Charging and shielding of a conducting sphere in an unmagnetized plasma} The next natural test is to study the charging and shielding of a perfectly conducting spherical object in an unmagnetized plasma (see for instance Ref. \cite{laframboise}). We can exploit the curvilinear geometry of CPIC and the fact that the system has spherical symmetry. Thus, we run CPIC in 2D with a physical space represented in cylindrical geometry $\left(x^1=r,\,x^2=\theta,\,x^2=z\right)$, and imposing azimuthal symmetry. The coordinate transformation between physical and logical space is then given by Eqs. (\ref{r}) and (\ref{z}). This corresponds to a physical domain between two spheres of radii $r_1$ and $r_2$. The simulation setup is as follows. The computational box is initially loaded with a Maxwellian plasma at rest. The boundary conditions on the left and right boundaries in logical space (inner and outer spheres, respectively) remain the same as in the previous example. However, we use homogeneous Neumann boundary conditions (zero normal derivative of the potential, $\partial \phi/\partial r=\partial \phi/\partial \xi=0$ since in this case the grid is orthogonal) for the fields at the bottom and top boundaries in logical space (corresponding to the $z$ axis in physical space) and particles are specularly reflected there. At the right boundary, we inject Maxwellian fluxes with no drift velocities. The other simulation parameters are the same as in Test A, except for $\omega_{pe}\Delta t=0.1$. We also set $r_1=1$ and $r_2=10$. The analogue of the equilibrium theory briefly discussed in the previous example can be developed in spherical geometry (see for instance Ref. \cite{kennedy}). It is more complicated due to the fact that the conservation of angular momentum and at least two components of the particle velocity must be included in the calculation. For these reasons, it is common to use an approximated theory, the Orbital-Motion-Limited (OML) theory \cite{langmuir}, to obtain the sphere floating potential. While OML is an approximated theory, previous studies have shown that OML indeed provides a quite accurate result even when the sphere radius is comparable to the plasma Debye length \cite{delzanno1}. Figure \ref{fig3} shows the time evolution of the normalized floating potential obtained for a $128\times 128$ mesh. Unlike the case of Fig. \ref{fig1}, the potential approaches its asymptotic equilibrium value monotonically. This is because the computational domain is loaded with plasma at t=0, and the ions immediately start charging the sphere. One can also see that the asymptotic equilibrium value, $\phi_{sphere}/T_e\simeq -2.54$, matches reasonably well the equilibrium value obtained by the OML theory, $\phi_{OML}/T_e=-2.50$. The fact that the equilibrium value is slightly lower than OML can be attributed to the development of an absorption barrier for the ions, as discussed in Ref. \cite{daugherty}. The equilibrium shielding potential is shown in Fig. \ref{fig4}. The spherical symmetry is evident from the plot. In Fig. \ref{fig5}, we compare the equilibrium shielding potential averaged over $\eta$ with that obtained with a 1D PIC code designed in spherical geometry. The latter was used for instance in Refs. \cite{delzanno1,delzanno2} to study charging and shielding of electron emitting spherical dust grains and is well benchmarked. There is a good agreement between the two, indicating that CPIC captures correctly the screening of the charged sphere by the plasma. Furthermore, we have performed a convergence study changing the number of grid points with the goal of checking the scalability of the solver in a practical situation. The results are presented in Table \ref{t4} (which also shows the average number of particles per cell at the end of the simulation). The number of solver iterations averaged over the entire simulation remains fairly flat. \begin{figure} \centering \includegraphics[width=3.3in]{proc_fig3-eps-converted-to.pdf} \caption{Test B: time evolution of the wall potential.} \label{fig3} \end{figure} \begin{figure} \centering \includegraphics[width=4in]{proc_fig4-eps-converted-to.pdf} \caption{Test B: equilibrium shielding potential at the end of the simulation.} \label{fig4} \end{figure} \begin{figure} \centering \includegraphics[width=3.3in]{proc_fig5-eps-converted-to.pdf} \caption{ Test B: equilibrium shielding potential at the end of the simulation, averaged over $\eta$.} \label{fig5} \end{figure} \begin{table} \caption{Convergence study of the BoxMG solver for Test B.} \label{t4} \centering \begin{tabular}{|c|c|c|c|} \hline Grid & Average number& Average number & Average number\\ & of iterations & of electrons per cell & of ions per cell \\ \hline \hline $32^2$ & $8.0$ & $231$ & $330$ \\ \hline $64^2$ & $8.0$ & $230$ & $276$ \\ \hline $128^2$ & $8.0$ & $230$ & $248$ \\ \hline $256^2$ & $9.0$ & $230$ & $230$ \\ \hline \hline \end{tabular} \end{table} \subsection{Charging and shielding of a conducting sphere in a magnetized plasma} Our last example involves the charging and shielding of a conducting, spherical object in a magnetized plasma. We consider a uniform magnetic field directed along $z$. For this case, since there is no analytic theory that relates the sphere floating potential to the plasma parameters for arbitrary values of the magnetic field, we perform some of the studies of Ref. \cite{marchand} for benchmark. Specifically, we consider two cases. In the first one, the magnetic field has magnitude $B_{z0}\simeq 1.9$ such that the electron thermal gyroradius is $\rho_e/\lambda_{De}=0.533$ and the electron gyrofrequency is $\omega_{ce}/\omega_{pe}\simeq 1.9$. Other parameters and the simulation settings are as in Test B. In particular, the ion thermal gyroradius is $\rho_i/\lambda_{De}\simeq 22.8$ and the ion gyrofrequency is $\omega_{ci}/\omega_{pe}\simeq 0.001$. Thus, in this example, the electron gyroradius is comparable to the sphere radius and we refer to this case as 'weakly magnetized'. In the second example we increase the value of the magnetic field by a factor of $10$, $B_{z0}\simeq 19$. Consistently, the electron thermal gyroradius is $\rho_e/\lambda_{De}=0.0533 \ll r_1/\lambda_{De}$, while the ion thermal gyroradius is $\rho_i/\lambda_{De}\simeq 2.3>r_1/\lambda_{De}$. The electron gyrofrequency is $\omega_{ce}/\omega_{pe}\simeq 18.8$, while for the ions $\omega_{ci}/\omega_{pe}\simeq 0.01$. We refer to this case as 'moderately magnetized', to distinguish from the strongly magnetized case when $\rho_i\ll r_{1}$. We note that with these values of the magnetic field the electron gyromotion corresponds to the shortest length scale and fastest frequency in the system. We have performed simulations with CPIC for the systems just described using a $128 \times 128$ grid with time step $\omega_{pe}\Delta t=0.05$. For the moderately magnetized case we note the important fact that this choice of the time step underresolves the electron gyromotion ($\omega_{ce} \Delta t \approx 1$). A PIC code with a standard leap-frog particle mover would be very inaccurate and possibly numerically unstable for these parameters. As we have argued above, the reason why we can afford such large time step is that for a uniform magnetic field CPIC uses the cyclotronic particle integrator \cite{patacchini,delzanno_mover} and only needs to resolve accurately the dynamics associated with the electric field. For the weakly magnetized case, the floating potential averaged over the last sixth of the simulation is $\phi_{sphere}/T_e\simeq -2.51$, slightly lower than the one obtained for the unmagnetized plasma. In addition, the shielding potential remains symmetric (not shown). For the moderately magnetized case, the floating potential averaged over the last sixth of the simulation is $\phi_{sphere}/T_e\simeq -2.43$, further lowered relative to the weakly magnetized case. Also, the shielding potential is not symmetric anymore (not shown). In order to interpret the results just discussed, we have studied the trajectories of test particles moving in the equilibrium configuration obtained from CPIC. We have injected $500,000$ particles for each species, uniformly located on the outer boundary (in the region defined by $\cos\left[\pi\left(1-\eta\right) \right]\ge0.8$) and with injection velocities distributed according to a Maxwellian, and have followed them until they either hit the sphere or leave the system. Figure \ref{fig6} shows the number of particles that were able to hit the sphere in terms of their initial angular distribution on the outer boundary $\cos(\delta)$, with \begin{equation} \delta=\pi \left(1-\eta\right). \label{delta} \end{equation} The parameter $\delta$ is the colatitude in the spherical coordinate system with axis along the magnetic field: $\cos \delta=0$ corresponds to the equatorial plane, while $\cos \delta =\pm 1$ is the sphere axis. The plot in Fig. \ref{fig6} is obtained for the weakly magnetized case ($B_{z0}\simeq1.9$). One can clearly see that the ions (bottom panel) can reach the sphere isotropically from the outer boundary. On the other hand, the majority of electrons (top panel) is collected in a flux tube of characteristic radius $r_{flux}\simeq r_1+\rho_e$ directed along the magnetic field, for which $\cos(\delta)\simeq 0.98$. However, it is interesting to note that some electrons are collected in a larger flux tube. For instance, $\cos(\delta)=0.9$ corresponds to a radius $r\simeq 4.4$. These are electrons that have higher initial velocities, towards the tail of the Maxwellian distribution function, and therefore have a large gyroradius. \begin{figure} \centering \includegraphics[width=3in]{fig6-eps-converted-to.pdf} \caption{Test C, $B_{z0}\simeq 1.9$: Outer boundary angular distribution, $\cos(\delta)=\cos[\pi(1-\eta)]$, of the test particles that hit the sphere. Initially $500,000$ test particles for each species were injected from the outer boundary with Maxwellian distribution functions. The test particles were distributed uniformly on the boundary in the region defined by $\cos\left[\pi\left(1-\eta\right) \right]\ge0.8$.} \label{fig6} \end{figure} The diagnostic just described is complemented with the distribution of particle collection on the surface of the sphere. This is shown in Fig. \ref{fig7} with data obtained directly from CPIC over the last tenth of each simulation in Tests B and C. The top panels are for the unmagnetized case, the middle panels for the weakly magnetized case and the bottom panels for the moderately magnetized case. As expected, when the plasma is unmagnetized, both electrons and ions are collected isotropically on the sphere. When the plasma is weakly magnetized the ions are still collected isotropically, since $\rho_i\gg r_1$. The electrons start to show a small depletion around $\cos(\delta)=0$. Interestingly, even though the electrons are collected primarily in a flux tube centered around the sphere with characteristic radius of a few electron thermal gyroradii, they can still hit the sphere quite isotropically since their gyroradius is comparable to the sphere radius. As a consequence, the shielding potential remains symmetric. For the moderately magnetized case ($B_{z0}\simeq 19$), $\rho_e \ll r_1$ and the electron motion is severely constrained by the magnetic field. Therefore, the electrons do not hit the sphere isotropically, and there is a large depletion in the number of electrons around $\cos(\delta)\approx 0$. The ions, on the other hand, still hit the sphere isotropically since $\rho_i>r_1$. As a final comment, we note that these results are consistent with the findings of Ref. \cite{marchand}. For weakly to moderately magnetized plasmas, the electron collection is restricted by the magnetic field and the resulting electron current is lower relative to the unmagnetized case. Consequently, the floating potential on the sphere becomes less negative. Quantitatively, Ref. \cite{marchand} gives $\phi_{sphere}/T_e=-2.41$ ($B_{z0}=0$), $\phi_{sphere}/T_e=-2.37$ ($B_{z0}=1.9$), and $\phi_{sphere}/T_e=-2.21$ ($B_{z0}=19$), which is in reasonable agreement (within $\sim 10\%$) with our results. \begin{figure} \centering \includegraphics[width=3.5in]{fig7-eps-converted-to.pdf} \caption{Test C: Angular distribution, $\cos(\delta)=\cos[\pi(1-\eta)]$, of particles collected by the sphere over the last tenth of each simulation performed for Tests B and C.} \label{fig7} \end{figure} \section{Conclusion} We have described CPIC: a fully-kinetic, electrostatic, body-fitted PIC code in curvilinear geometry for structured grids. CPIC was designed with the goal of flexibility (namely the ability to handle different geometries, complex computational domains, and the interaction of plasmas with complex objects with characteristic size smaller than the characteristic plasma length scales), robustness and performance. It introduces a coordinate transformation from the physical space to the logical space, where the grid is uniform and Cartesian. In CPIC, most operations are performed in logical space. Its main features are (1) the use of structured grids, (2) a scalable and robust field solver based on the Black Box multigrid algorithm, and (3) an hybrid particle mover, where the particles' position is updated in logical space while the particles' velocity is updated in physical space. In our tests, the hybrid mover has proven more efficient and robust than the conventional physical space mover. We have presented some successful benchmark tests involving the interaction of a plasma with material boundaries, in Cartesian and spherical geometries, with and without magnetic field. \section*{Acknowledgment} The authors wish to thank Luis Chacon, Chris Fichtl, John M. Finn, Rao Garimella, Richard Marchand, and Xianzhu Tang. This research is supported by the LDRD program of the Los Alamos National Laboratory and was performed under the auspices of the NNSA of the U.S. DOE by LANL, operated by LANS LLC under Contract No. DE-AC52-06NA25396. {\bibliographystyle{ieeetr}
\section{Introduction} Spectra of sub-Laplace operators on sub-Riemannian manifolds are intensely studied. Especially interesting is the case where the distribution defining the sub-Riemannian structure is bracket generating, what we shall assume in the following. In this case the sub-Laplacian is known to be hypo-elliptic \cite{H}. Many explicit calculations of the spectrum have been done in the situation where the underlying manifold is a compact Lie group or a quotient of a Lie group by a discrete cocompact subgroup, see, for example, \cite{BF1,BFI,BF2,P}. In \cite{BF1,BFI,P} the authors study spectral properties of sub-Laplace operators on nilpotent groups of step two and on compact quotients by discrete subgroups. They determine the heat kernels of these operators. This allows an explicit determination of the spectrum of the sub-Laplacian, which is discrete in this situation. The aim of this paper is to study spectra of sub-Riemannian analogs of the classical Dirac operator. In the definition of the sub-Dirac operator the following difficulty occurs: In contrast to the Riemannian case where we have the Levi-Civita connection as a preferred connection, in general, there is no connection canonically associated with a sub-Riemannian structure. Only in special geometric situations a canonical connection exists. Hence the following definition of the sub-Dirac operator depends on the choice of a connection. Let $(M,\cH,g)$ be a sub-Riemannian manifold, $\dim {\cal H}=d$. Suppose that $\nabla$ is a metric connection on $\cH$. Moreover, assume that $\cH$ is oriented and that the bundle of orthonormal frames of $\cH$ admits a reduction to $\Spin(d)$. Such a reduction will be called a spin structure of $\cH$. Then we can associate a spinor bundle $S$ with this spin structure. Moreover, using the connection $\nabla$ we can define a sub-Riemannian Dirac operator, which acts on sections in $S$. Only few results for sub-Riemannian analogs of the Dirac operator are known. For the case of a sub-Riemannian manifold of contact type such an operator was introduced and studied by Petit \cite{Pe}, who called this operator Kohn-Dirac operator. More exactly, this was done for a $\Spin^c$-structure. Studying the sub-Riemannian Dirac operator the following natural questions arise: Which structure does its spectrum have? How does the spectrum depend on the sub-Riemannian geometry of the manifold and on the spin structure of $\cH$? How do the sub-Dirac operator and its spectrum depend on the chosen connection? In general, i.e., for arbitrary metric connections in $\cH$, the sub-Dirac is not symmetric. We will characterize the symmetry of this operator by a simple condition on the connection. Here we focus on nilmanifolds. More precisely, we study sub-Dirac operators on manifolds of the form $M=\GG$ where $G=\RR^n\rtimes_A\RR$ is a semi-direct product defined by a one-parameter subgroup $A(t)$ of unipotent matrices in $\GL(n,\RR)$ and $\Gamma$ is the subgroup $\ZZ^n\rtimes_A\ZZ$. These manifolds $M$ can be interpreted as a suspension of the diffeomorphism of the torus $\RR^n/\ZZ^n$ induced by $A(1)$. This is also the starting point of \cite{J} where the spectrum of the Laplacian on left-invariant differential forms on $M$ is considered. Our sub-Dirac operators will be associated with sub-Riemannian structures $(\dot{\cH},\dot{g})$ on $\GG$ coming from a left-invariant and bracket-generating distribution $(\cH,g)$ on $G$. We choose a metric connection in $H$ such that $D$ is symmetric. Our approach is to give an explicit decomposition of the regular representation of $G$. Roughly speaking, it turns out that the sub-Dirac operator is an orthogonal sum of elliptic operators on the real line, each having a discrete spectrum. This shows that $D$ on $\GG$ has pure point spectrum. We apply our results to compute the spectrum of $D$ explicitly for two classes of two-step nilmanifolds of the above form. First we consider three-dimensional Heisenberg manifolds. Secondly, we study a class of five-dimensional two-step nilpotent nilmanifolds with a three-dimensional distribution. The latter example shows that the spectrum of the sub-Dirac operator is not necessarily a discrete subset of $\RR$ and that its eigenvalues may have infinite multiplicity, contrary to the results for the spectrum of the sub-Laplacian on compact $2$-step nilmanifolds. Finally, we discuss a three-step nilpotent example of dimension four with a two-dimen\-sional distribution. In this case the spectrum can be expressed in terms of the spectra of the family of operators $P_c=\del_t^2+(t^2+c)^2\pm 2t$, $c\in\RR$. In all three examples, the multiplicities of the eigenvalues of $D$ can be read off from the coadjoint orbit picture. \\[2ex] {\bf Acknowledgements} We would like to thank Paul-Andi Nagy for several useful discussions. \section{Sub-Riemannian Dirac operators} \subsection{Definition of sub-Dirac operators} Let $M$ be a smooth manifold and let $\cH\subset TM$ be a smooth distribution, where $\dim\cH_x=d$ for all $x\in M$. Let $\Gamma(\cH)$ denote the space of smooth sections of $\cH$. We assume that $\cH$ is bracket-generating. That means, that for each $x\in M$ there is a $J\in\NN$ such that the sequence $$\Gamma_0:=\Gamma(\cH),\quad \Gamma_{j+1}:=\Gamma_j+[\Gamma_0,\Gamma_j]$$ satisfies $\{X(x)\mid X\in \Gamma_{J}\}=T_xM$. If $g$ is a Riemannian metric on $\cH$, then the pair $(\cH,g)$ is called a sub-Riemannian structure on $M$ and $(M,\cH,g)$ is called a sub-Riemannian manifold. Let $\nabla:\Gamma(\cH)\otimes\Gamma(\cH)\rightarrow\Gamma(\cH)$ be a metric connection on $\cH$. Note that here we consider only derivations by vector fields in $\cH$. Suppose that $\cH$ is oriented and that it admits a spin structure, i.e., that there is a $\Spin(d)$-reduction $P_{\Spin}(\cH)$ of the principal $\SO(d)$-bundle $P_{\SO}(\cH)$ of oriented orthonormal frames of $(\cH,g)$. We consider the complex representation of $\Spin(d)$ which is obtained by restriction of (one of) the complex irreducible representation(s) of the Clifford algebra ${\cal C}l(d):=\Cl(\RR^d)$. We will call it spinor representation and denote it by $\Delta_d$. The associated bundle $P_{\Spin}(\cH)\times_{\Spin(d)}\Delta_d$ is called spinor bundle $S$ of $(\cH,g)$. The space of smooth sections in $S$ is denoted by $\Gamma(S)$. The connection $\nabla$ defines a connection $\nabla^S:\Gamma(\cH)\times \Gamma(S)\rightarrow \Gamma(S)$ in the following way. Let $s_1,\dots, s_d$ be a local orthonormal frame of $\cH$ and consider the local connection forms $\omega_{ij}=g(\nabla s_i,s_j)$. Then we define $$\nabla^S_X\ph:=X(\ph)+{\textstyle\frac12}\sum_{i<j} \omega_{ji}(X)\;s_i\cdot s_j\cdot\ph,$$ where `$\,\cdot\,$' denotes the Clifford multiplication. Now we can define a sub-Riemannian Dirac operator, or sub-Dirac operator for short, by \begin{equation}\label{ED} D=\sum_i s_i\cdot \nabla^S_{s_i}:\Gamma(S)\longrightarrow \Gamma(S), \end{equation} where again $s_1,\dots, s_d$ is a local orthonormal frame of $\cH$. Note, that the definition of $D$ depends on the choice of the connection $\nabla$ on $\cH$ and that, in general, this choice is far from being canonical in contrast to the Riemannian case, where we have the Levi-Civita connection as a preferred connection. A large class of metric connections in $\cH$ can be obtained in the following way. Suppose we are given a further distribution ${\cal V}\subset TM$ such that $TM=\cH \oplus{\cal V}$. Then this decomposition of $TM$ gives us a projection $\proj:TM\rightarrow \cH$ and we can define a connection $\nabla$ by the Koszul formula \begin{eqnarray}\label{EK} 2g(\nabla_X Y,Z)&=& X(g(Y,Z))+Y(g(X,Z))-Z(g(X,Y))\nonumber \\ &&+g(\proj[X,Y],Z)-g(\proj[X,Z],Y)-g(\proj [Y,Z],X), \end{eqnarray} where $X,Y,Z\in\Gamma(\cH)$. In this case $\nabla$ is uniquely determined by the vanishing of $\nabla_XY-\nabla_YX-\proj[X,Y]$. \subsection{Symmetry of the sub-Dirac operator}\label{S2.2} A sub-Riemannian manifold is said to be regular if for each $j=1,\dots,J$ the dimension of $\{X(x)\mid X\in \Gamma_{j}\}$ does not depend on the point $x\in M$. Let $(M,\cH,g)$ be an orientable regular sub-Riemannian manifold. Then $(M,\cH,g)$ admits an intrinsic volume form $\omega_0$ on $M$, see \cite{M}, Section 10.5 and \cite{ABGR}. Let $(M,\cH,g)$ be an oriented and regular sub-Riemannian manifold. Consider any volume form $\omega$ on $M$, e.g., the intrinsic one. Define the divergence of a vector field $X$ on $M$ by ${\cal L}_X\omega=(\Div X)\cdot \omega$. Let $\nabla$ be a metric connection on $\cH$. Suppose that $\cH$ admits a spin structure and define $D:\Gamma(S)\rightarrow \Gamma(S)$ as above. Let $\ip$ be a hermitian inner product on $\Delta_d$ for which the Clifford multiplication is antisymmetric. This inner product is unique up to scale. It induces a hermitian inner product on $S$, which together with $\omega$ gives an $L^2$-inner product $\rip$ on the space $\Gamma_0(S)$ of sections in $S$ with compact support. It is easy to find examples of three-dimensional Heisenberg manifolds with two-dimensional distribution $\cH$ and metric connection for which $D$ is not symmetric, see Section~\ref{S4.2}. The following lemma states that the sub-Dirac operator is symmetric if and only if the divergence defined by the sub-Riemannian structure coincides with the divergence given by the connection, compare also \cite{FS} for the Riemannian case. \begin{lm}\label{lm:sym_of_D} Under the above conditions, $D$ is symmetric if and only if \begin{equation}\label{Ediv}\Div X = \sum_{i=1}^d \la \nabla_{s_i}X,s_i\ra \end{equation} holds for one (and therefore for every) local orthonormal basis $s_1,\dots, s_d$ of $\cH$. If, in addition, ${\cal V}$ is a complement of $\cH$ in $TM$ and $\nabla$ is defined as in {\rm (\ref{EK})}, then {\rm (\ref{Ediv})} is equivalent to the following condition. For one (and therefore for all) sets $\{\xi_1,\ldots,\xi_{l}\}$, $l=\dim M-k$, of local sections of ${\cal V}$ that satisfy $\omega(s_1,\ldots,s_d,\xi_1,\ldots,\xi_{l})=1$ the equation \[\eta_1([X,\xi_1])+\dots+\eta_l([X,\xi_l])=0\] holds for all $X\in\Gamma(\cH)$, where $\eta_1,\ldots,\eta_{l}\in\Gamma(T^\ast M)$ are defined to be zero on $\cH$ and dual to $\xi_1,\ldots,\xi_{l}$. In particular, if $\codim\,\cH=1$, then $D$ is symmetric if and only if $[\Gamma(\cH),\xi_1]\subset\Gamma(\cH)$. \end{lm} \proof Consider sections $\ph,\psi\in\Gamma_0(S)$ and define $f:\cH\rightarrow \CC$ by \begin{equation}\label{E1a}f(w):= \langle \ph,w \cdot \psi\rangle. \end{equation} Moreover, define $u\in \Gamma_0(\cH\otimes \CC)$ by \begin{equation}\label{E1} g^{\Bbb C}(u,w)=f(w) \end{equation} for all $w\in\Gamma(\cH)$, where $g^{\Bbb C}$ denotes the the complex bilinear extension of $g$. Choose a local orthonormal frame $s_1,\dots, s_d$ of $\cH$. Then $$\langle D\ph,\psi \ra-\la\ph,D\psi\ra =\sum_{i=1}^d\Big( f(\nabla_{s_i}s_i)-s_i(f(s_i))\Big) =\sum_{i=1}^d g^{\Bbb C}(\nabla_{s_i}u,s_i),$$ thus $$(D\ph,\psi)-(\ph,D\psi)=\int_M \left(\sum_{i=1}^d g^{\Bbb C}(\nabla_{s_i}u,s_i)\right)\omega =\int_M \left(\sum_{i=1}^d g^{\Bbb C}(\nabla_{s_i}u,s_i)-\Div(u)\right)\omega.$$ In particular, (\ref{Ediv}) is sufficient for the symmetry of $D$. On the other hand, any section $u_1\in\Gamma_0(\cH)$ is the real part of a section $u\in \Gamma_0(\cH\otimes \CC)$ that satisfies~(\ref{E1a}) and (\ref{E1}) for some $\ph,\psi\in\Gamma_0(S)$. Indeed, choose $\psi$ such that $\la \psi(x),\psi(x)\ra=1$ for all $x\in {\rm supp}\,u_1$ and put $\ph:= u_1\cdot\psi$. Define $u$ by (\ref{E1a}) and (\ref{E1}). Then $$g(u_1,w)=\Re \la u_1\cdot\psi,w\cdot\psi\ra=\Re \la \ph,w\cdot\psi\ra=\Re f(w)$$ for all $w\in\Gamma(\cH)$, hence $u_1=\Re u$. Consequently, the symmetry of $D$ implies $$\int_M \left(\sum_{i=1}^d g(\nabla_{s_i}u,s_i)-\Div(u)\right)\omega=0$$ for all $u\in\Gamma_0(\cH)$. Since the integrand is $C^\infty_0(M)$-linear in $u$, Equation (\ref{Ediv}) follows. The second part of the lemma now follows from \begin{eqnarray*} &&\;\Div(u)\ =\ ({\cal L}_u\omega)(s_1,\dots,s_d,\xi_1,\dots,\xi_l)\\ &=& -\sum_{i=1}^d\omega(s_1,\ldots,[u,s_i],\dots, s_d,\xi_1,\dots,\xi_l) -\sum_{j=1}^l\omega(s_1,\dots,s_d,\xi_1,\ldots,[u,\xi_j],\dots,\xi_l)\\ &=& -\sum_{i=1}^d g(\proj [u,s_i],s_i)-\sum_{j=1}^l \eta_j([u,\xi_j]) \ =\ \sum_{i=1}^d g(\nabla_{s_i}u,s_i)-\sum_{j=1}^l \eta_j([u,\xi_j]),\\ \end{eqnarray*} where the last equality is a consequence of Equation~(\ref{EK}).\qed \subsection{Sub-Dirac operators on Lie groups and compact quotients} Let $G$ be a simply connected Lie group and $\Gamma\subset G$ a uniform discrete subgroup. Let $\cH\subset TG$ be a left-invariant distribution and $ g$ a left-invariant Riemannian metric on $\cH$. Obviously, $\cH$ is spanned by orthonormal left-invariant vector fields $s_1,\dots, s_d$. In particular, the frame bundle $P_{\SO}(\cH)$ is a trivial bundle and the unique spin structure of $\cH$ equals $P_{\Spin}(\cH)=G\times \Spin(d)$. The pair $(\cH, g)$ induces a sub-Riemannian structure on $\Gamma\setminus G$, which we will denote by $(\dot\cH,\dot g)$. The frame bundle $P_{\SO}(\dot\cH)$ can be identified with $$P_{\SO}(\dot\cH)=G\times_\Gamma \SO(d),$$ where $\Gamma$ acts by left multiplication on $G$ and trivially on $\SO(d)$. There is a one-to-one correspondence between homomorphisms $\eps:\Gamma\rightarrow\ZZ_2=\{0,1\}$ and spin structures of $\dot\cH$ given by $$\eps\longmapsto P_{\Spin,\eps}(\dot\cH)=G\times_\Gamma \Spin(d),$$ where $\gamma\in \Gamma$ acts by multiplication by $e^{i\pi\eps(\gamma)}$ on $\Spin(d)$. Spinor fields are sections of the associated spinor bundle $P_{\Spin,\eps}(\dot\cH)\times_{\Spin(d)}\Delta_d\cong{G\times_\Gamma}\Delta_d$ or, equivalently, maps $\psi:G\rightarrow \Delta_d$ that satisfy $\psi(\gamma g)=e^{i\pi\eps(\gamma)}\psi(g)$ for all $\gamma\in\Gamma$, $g\in G$. The intrinsic volume form $\omega_0$ on $\Gamma\setminus G$ introduced in \cite{ABGR} and discussed in Section~\ref{S2.2} is left-invariant. Now let $\nabla$ be a left-invariant metric connection on $\dot\cH$. Let $s_1,\dots, s_d$ be an orthonormal basis of $\dot\cH$ consisting of left-invariant vector fields. As for the symmetry of the Dirac operator discussed in Lemma~\ref{lm:sym_of_D}, note that Equation~(\ref{Ediv}) is equivalent to \begin{equation} 0=\sum_{i=1}^d\,g(\nabla_{s_i}s_j,s_i) \end{equation} for all $j=1,\dots,d$. Indeed, since the intrinsic volume form is left-invariant and $G$ must be unimodular the divergence of any left-invariant vector field vanishes. \section{Sub-Riemannian structures on $\Gamma\setminus (\RRu ^n\rtimes_A\RRu)$} \subsection{The standard model}\label{S3.1} Let $A(t)=\exp(tB)$ be a one-parameter subgroup of $\GL(n,\RR)$. We consider the simply-connected solvable Lie group $\RR^n\rtimes_A\RR$ with group law \[(x,s)\,(y,t)=(x+A(s)y\,,\,s+t)\;.\] In particular, $(0,t)\,(x,0)\,(0,t)^{-1}=(A(t)x\,,\,0)$. In addition, we assume $A(1)\in\SL(n,\ZZ)$ so that the set $\ZZ^n\times\ZZ$ becomes a uniform discrete subgroup. The pair $(\RR^n\rtimes_A\RR,\ZZ^n\rtimes_A\ZZ)$ serves as a standard model in the following sense. \begin{lm}\label{lm:model} Let $G$ be an exponential Lie group admitting a connected abelian normal subgroup $N$ of codimension one. Let $\Gamma$ be a uniform discrete subgroup of $G$ such that $\Gamma\cap N$ is uniform in $N$. Then there exists a one-parameter subgroup $A$ of~$\GL(n,\RR)$, $n=\dim N$, with $A(1)\in\SL(n,\ZZ)$, and an isomorphism~$\Phi$ of $\RR^n\rtimes_A\RR$ onto $G$ mapping $\ZZ^n\rtimes_A\ZZ$ onto $\Gamma$. \end{lm} \begin{proof} We fix generators $v_1,\ldots,v_n$ of the lattice $\Gamma\cap N$ of the vector group $N$ and consider the linear isomorphism $M$ of $\RR^n$ onto $N$ given by $M(e_j)=v_j$. On the other hand, the assumption on $\Gamma$ implies that $\Gamma N$ is closed in $G$ and that $\Gamma N/N$ is a discrete subgroup of $G/N$. Hence there exists $b\in\fg$ with $\exp(b)\in\Gamma$ and such that $\exp(b)N$ is a generator of $\Gamma N/N$. Put $A(t)x=M^{-1}(\exp(tb)M(x)\exp(tb)^{-1})$. Now it follows that $\Phi(x,t)=M(x)\exp(tb)$ is an isomorphism of $\RR^n\rtimes_A\RR$ onto $G$ with $\Phi(\ZZ^n\times\ZZ)=\Gamma$. In particular, $\ZZ^n\rtimes_A\ZZ$ is a subgroup of $\RR^n\rtimes_A\RR$. This means that $\ZZ^n$ is $A(l)$-invariant for all $l\in\ZZ$ so that $A(l)\in\SL(n,\ZZ)$. \qed \end{proof} The condition $A(1)\in\SL(n,\ZZ)$ implies $B\in\fsl(n,\RR)$ and $A(t)\in\SL(n,\RR)$ for all $t\in\RR$. This reflects the fact that locally compact groups admitting a uniform discrete subgroup are unimodular, compare Theorem~7.1.7 of~\cite{W}. The Lie algebra of $G:=\RR^n\rtimes_A\RR$ is isomorphic to $\fg=\RR^n\rtimes_B\RR$, and $B=\ad(b)\,|\,\fn$, where $b=(0,1)$ and $\fn=\RR^n\times\{0\}$. Note that $G$ is exponential if and only if $B$ has no purely imaginary eigenvalues, compare Theorem~1 of~\cite{LL}. It is evident that \[\pi(x,t)=\left( \begin{array}{cccc} & & & x_1 \\ & A(t) & & \vdots \\ & & & x_n \\ 0 & \ldots & 0 & 1 \end{array}\right)\] defines a representation, which is faithful provided that $G$ is exponential and not abelian. \begin{ex}\label{ex:3_dim_Heis} {\rm Fix $r\in\ZZ_+$ and set $B=\left(\begin{array}{cc} 0 & r\\ 0 & 0\end{array}\right)$ so that $A(t)=\exp(tB)=I+tB$. Since $A(l)\in\SL(2,\ZZ)$ for $l\in\ZZ$, $\Gamma=\ZZ^2\rtimes_A\ZZ$ is a subgroup of $G=\RR^2\rtimes_A\RR$. On the other hand, \[\pi(x_1,x_2,t)=\left(\begin{array}{ccc} 1 & rt & x_1 \\ 0 & 1 & x_2 \\ 0 & 0 & 1\end{array}\right)\] gives an isomorphism from $G$ onto the three-dimensional Heisenberg group $H(1)$ in its standard realisation as a group of matrices mapping $\Gamma$ onto \[\Gamma_r=\left\{\;\left(\begin{array}{ccc} 1 & rl & k_1 \\ 0 & 1 & k_2 \\ 0 & 0 & 1 \end{array}\right):l,k_1,k_2\in\ZZ\;\right\}\;.\] In particular, the above construction yields all uniform discrete subgroups of $H(1)$ and hence all three-dimensional Heisenberg manifolds, compare Section~2 of~\cite{GW}.} \end{ex} Heisenberg manifolds and certain generalisations of them will be discussed in Section~\ref{S4.2} and~\ref{S4.3} in greater detail. For the subgroup $\Gamma:=\ZZ^n\rtimes_A\ZZ$, the spin structures of any distribution of $T(\GG)$ induced by a left-invariant distribution of $T(G)$ are determined as follows. \begin{lm}\label{lm:eps} A map $\eps:\Gamma\to\ZZ_2$ is a homomorphism if and only if $\eps(k,l)=\eps'(k)+\dot\eps(l)$ for some homomorphism $\dot{\eps}:\ZZ\to\ZZ_2$ and a homomorphism $\eps':\ZZ^n\to\ZZ_2$ satisfying \begin{equation}\label{EsumA} \sum_\mu \eps'(e_\mu)(A(1)-I)_{\mu\nu}\,\in 2\ZZ\end{equation} for all $\nu$. \end{lm} \begin{proof} Any homomorphisms $\eps:\Gamma\to\ZZ_2$ defines homomorphisms $\dot{\eps}:\ZZ\to\ZZ_2$, $\dot{\eps}(l)=\eps(0,l)$, and $\eps':\ZZ^n\to\ZZ_2$, $\eps'(k)=\eps(k,0)$, where $\eps'$ satisfies \[\eps'(A(l)k)=\eps(A(l)k,0)=\eps(\,(0,l)(k,0)(0,l)^{-1}\,)=\eps(k,0)=\eps'(k)\] for all $l\in\ZZ$ and $k\in\ZZ^n$. The latter condition reduces to $\eps'(A(1)k)=\eps'(k)$ for all $k$, and hence to $\eps'(e_\nu)=\eps'(A(1)e_\nu)=\sum_\mu A_{\mu\nu}(1)\,\eps'(e_\mu)$ in $\ZZ_2$ for all $\nu$, which is equivalent to (\ref{EsumA}). It is easy to check that the converse holds also true.\qed \end{proof} Since $\Gamma$ is uniform and discrete, there exists a unique normalised right $G$-invariant Radon measure $\mu$ on~$\GG$ which can be obtained as follows: Let $I_n=[0,1]^n$ denote the unit cube of $\RR^n$. Then \begin{equation}\label{eq:inv_measure} \int_{\GG}\ph\;d\mu=\int_0^1\Big(\int_{I_n}\;\ph(x,s)\;dx\Big)\;ds \end{equation} for all $\ph\in C(\GG)$. This can be proved using that the Haar measure of~$G$ equals the Lebesgue measure of~$\RR^{n+1}$ and that $F=[0,1)^{n+1}$ is a fundamental set for~$\Gamma$ on~$G$. \subsection{Decomposition of the right-regular representation} Let $A(t)$ be one-parameter subgroup of $\GL(n,\RR)$ such that $G=\RR^n\rtimes_A\RR$ is exponential and $\Gamma=\ZZ^n\rtimes_A\ZZ$ is a subgroup of $G$. Let $\eps:\Gamma\to\ZZ_2$ be a group homomorphism. Our aim is to decompose the right regular representation of $G$ on $L^2(G,\eps)$. Let $C(G,\eps)$ denote the space of all continuous $\CC$-valued functions $\ph$ on $G$ satisfying $\ph(gy)=e^{i\pi\eps(g)}\,\ph(y)$ for all $g\in\Gamma$ and $y\in G$, and $L^2(G,\eps)$ the completion of $C(G,\eps)$ with respect to the Hilbert norm \[|\,\ph\,|_{L_2}^2=\int_{\GG}|\,\ph\,|^2\;d\mu\;.\] Now right translation $\rho(x)\ph\;(y)=\ph(yx)$ gives rise to a unitary representation of $G$ on~$L^2(G,\eps)$. This is precisely the definition of the induced representation $\rho=\ind_\Gamma^G e^{i\pi\eps}$ of the unitary character $e^{i\pi\eps}$ of~$\Gamma$. By Theorem~7.2.5 of \cite{W}, $\rho$ can be written as a countable orthogonal sum of irreducible subrepresentations with finite multiplicity. We will give such a decomposition explicitly, generalizing the results of $\cite{AB}$ for the three-dimensional Heisenberg group, which motivated this article. To this end, we consider partial Fourier transformation with respect to the first n variables: If $\ph\in C(G,\eps)$, then $\ph(k+A(l)x,l+t)=e^{i\pi\eps(k,l)}\,\ph(x,t)$ for all $(k,l)\in\Gamma$. In particular, $\ph(2k+x,t)=e^{i\pi\eps(2k,0)}\,\ph(x,t)=\ph(x,t)$ which shows that $x\mapsto\ph(x,t)$ is $2\ZZ^n$-invariant. For such functions it is natural to consider \[\widehat\ph(\xi,t)=\int_{I_n} \ph(2x,t)\,e^{-2\pi i\langle\xi,x\ra}\,dx\] for $\xi\in\ZZ^n$. Clearly $\ph$ is uniquely determined by its Fourier coefficient functions. It follows from~(\ref{eq:inv_measure}) that the restriction of $\ph\in L^2(G,\eps)$ to $2I_n\times[0,1]$ is $L^2$- and hence $L^1$-integrable with respect to the Lebesgue measure. In particular, the integral in the definition of $(F_\xi\cdot\ph)(t):=\widehat\ph(\xi,t)$ makes sense for almost all $t$. \begin{pr}\label{pr:Plancherel} For $\ph\in L^2(G,\eps)$ there holds the Plancherel formula $$|\,\ph\,|^2_{L^2}=\sum_{\xi\in{\Bbb Z}^n}\;\int_0^1|\widehat \ph(\xi,t)|^2\;dt\;.$$ \end{pr} \begin{proof} By the Plancherel theorem for $L^2$-functions on the torus, we obtain $$|\,\ph\,|^2_{L^2}=\int_0^1\Big(\int_{I_n}\;|\,\ph(x,t)\,|^2\;dx\Big)\;dt =\int_0^1\sum_{\xi \in{\Bbb Z}^n}\;|\widehat \ph(\xi,t)|^2\;dt\;,$$ where summation and integration can be interchanged. \qed \end{proof} As a corollary we get $F_\xi\cdot\ph\in L^2(\,[0,1]\,)$. Moreover, the inequality $|\,F_\xi\cdot\ph\,|_{L^2}\le |\,\ph\,|_{L_2}$ shows that $F_\xi:L^2(G,\eps)\to L^2(\,[0,1]\,)$ is a continuous linear operator. Taking into account \[\int_{I_n}g(Mx)\;dx=\int_{I_n}g(x)\;dx\] for integrable $\ZZ^n$-invariant functions $g$ on $\RR^n$ and $M\in\SL(n,\ZZ)$, one can prove that the $\eps$-equivariance of $\ph$ entails the following conditions on its Fourier transform. \begin{lm}\label{lm:eps_equiv} For $\ph\in L^2(G,\eps)$ and $(k,l)\in\Gamma$ it holds \begin{eqnarray*} e^{i\pi\eps(k,l)}\;\widehat\ph(\xi,t)=e^{\pi i\la\,A(-l)^\top\xi\,,\,k\,\ra}\; \widehat\ph(\,A(-l)^\top\xi\,,\,l+t\,)\;, \end{eqnarray*} where $A(l)^\top$ denotes the transpose of the operator $A(l)$ with respect to the standard inner product on $\RR^n$. \end{lm} The one-parameter subgroup $A$ represents the adjoint representation of the subgroup $\RR\cong\{0\}\times\RR$ of $G$ on the Lie algebra $\RR^n$ of the normal subgroup $\RR^n\times\{0\}$. Identifying the linear dual of this Lie algebra with $\RR^n$ by means of the standard inner product, we see that $t\mapsto A(t)^\top$ is the coadjoint representation. The preceding lemma reveals the importance of this group action in the present context. Any $\xi\in\RR^n$ has a $\ZZ$-orbit $\theta=\{A(l)^\top\xi:l\in\ZZ\}$ and an $\RR$-orbit $\omega=\{A(t)^\top\xi:t\in\RR\}$. Let $\ph\in L^2(G,\eps)$ and $\xi\in\ZZ^n$. The equality $\widehat{\ph}(\,A(l)^\top\xi,t)=e^{i\pi\eps(0,-l)}\,\widehat{\ph}(\xi,l+t)$ shows that $\widehat{\ph}(\xi,\cdot)$ determines $\widehat{\ph}(\eta,\cdot)$ for all $\eta\in\theta$. In particular, $\supp\,\widehat{\ph}:=\{\xi\in\ZZ^n: \widehat{\ph}(\xi,\cdot)\neq 0\}$ is a $\ZZ$-invariant subset. \begin{lm}\label{lm:Sigma} The set \[\Sigma_{\eps'}:=\{\,\xi\in\ZZ^n:\xi_\nu\in 2\ZZ+\eps'(e_\nu) \mbox{ \textup{for all} }1\le\nu\le n\,\}\] is $\ZZ$-invariant and contains $\supp\,\widehat{\ph}$ for every $\ph\in L^2(G,\eps)$. \end{lm} \begin{proof} Let $\xi\in\Sigma_{\eps'}$ and $l\in\ZZ$. Since $\eps'(A(l)^\top e_\nu)=\eps'(e_\nu)$ and $\la\xi,k\ra\in 2\ZZ+\eps'(k)$ for all $k\in\ZZ^n$, it follows $\la A(l)^\top\xi,e_\nu\ra=\la\xi,A(l)e_\nu\ra\in 2\ZZ+\eps'(e_\nu)$ and $A(l)^\top\xi\in\Sigma_{\eps'}$. This proves $\Sigma_{\eps'}$ to be $\ZZ$-invariant. Let $\ph\in L^2(G,\eps)$ and $\xi\not\in\Sigma_{\eps'}$. Then there is $1\le\nu\le n$ such that $\xi_\nu\not\in 2\ZZ+\eps'(e_\nu)$ and hence $ e^{\pi i\eps'(e_\nu)}\neq e^{\pi i\la\xi,e_\nu\ra}$. On the other hand, by Lemma~\ref{lm:eps_equiv} we have $e^{i\pi\eps(k,0)}\,\widehat{\ph}(\xi,t) =e^{\pi i\la\xi,k\ra}\,\widehat{\ph}(\xi,t)$ for all $k$. This implies $\widehat{\ph}(\xi,\cdot)=0$. \qed \end{proof} Note that $\xi\in\Sigma_{\eps'}$ implies $e^{i\pi\eps(k,0)}=e^{i\pi\la\xi,k\ra}$ for all $k\in\ZZ^n$. Let $\ZZ\setminus\Sigma_{\eps'}$ denote the set of all $\ZZ$-orbits in $\Sigma_{\eps'}$. For $\theta\in\ZZ\setminus\Sigma_{\eps'}$, it follows that \[U_\theta\,:=\,\bigcap_{\xi\not\in\theta}\ker F_\xi\,=\,\{\ph\in L^2(G,\eps): \supp\,\widehat{\ph}\subset\theta\}\,\neq\,0\] is a closed subspace of $L^2(G,\eps)$. It will be shown below that $U_\theta$ is $\rho(G)$-invariant. \begin{pr}\label{pr:L2_decomp} These subspaces form an orthogonal decomposition \[L^2(G,\eps)=\mathop{\overline{\bigoplus}}_{\theta\in{\Bbb Z}\setminus\Sigma_{\eps'}} U_\theta.\] \end{pr} \begin{proof} Let $\theta_1,\theta_2\in\ZZ\setminus\Sigma_{\eps'}$ be distinct orbits. By Proposition~\ref{pr:Plancherel} and the polarisation identity, we obtain \[\la\,\ph_1,\ph_2\,\ra_{L^2}\;=\;\sum_{\xi\in\Bbb Z^n}\,\int_0^1\, \widehat\ph_1(\xi,t)\,\overline{\widehat\ph_2(\xi,t)}\;dt\;=\;0\] for $\ph_1\in U_{\theta_1}$ and $\ph_2\in U_{\theta_2}$. Hence $U_{\theta_1}$ and $U_{\theta_2}$ are orthogonal. It remains to be shown that the direct sum of the $U_\theta$ is dense. Since $C^\infty(G,\eps)$ is dense in $L^2(G,\eps)$, it suffices to prove that every smooth $\eps$-equivariant function can be approximated by a finite sum of functions in the $U_\theta$. By the decay of the Fourier transform of $\ph\in C^\infty(G,\eps)$, it follows that \[\ph_\theta(x,t)=\sum_{\xi\in\theta}\;\widehat\ph(\xi,t)\,e^{\pi i\la \xi,x\ra}\] is a smooth function in $U_\theta$ and that $\ph=\sum_{\theta\in{\Bbb Z}\setminus\Sigma_{\eps'}}\,\ph_\theta$ converges uniformly on~$\RR^n\times [0,1]$. In particular, $|\,\ph-\sum_{\theta\in J}\ph_\theta\,|_{L^2}\to 0$ for $J\subset\ZZ\setminus\Sigma_{\eps'}$ finite and increasing. (This also proves that $U_\theta\cap C^\infty(G,\eps)$ is dense in $U_\theta$.) \qed \end{proof} The right regular representation $\rho(x,s)\ph\;(y,t)=\ph(y+A(t)x,t+s)$ on $L^2(G,\eps)$ is compatible with partial Fourier transform in the sense that \begin{equation}\label{eq:rho_xi} (\rho(x,s)\ph)\,^{\widehat{}}\ (\xi,t)=e^{\pi i\la A(t)^\top\xi,x\ra}\; \widehat\ph(\xi,t+s)\;. \end{equation} In particular, $\widehat\ph(\xi,\cdot)=0$ implies $(\rho(x,s)\ph)\,^{\widehat{}}\ (\xi,\cdot)=0$ for all $(x,s)\in G$ which proves $U_\theta$ to be $\rho(G)$-invariant. We define $\rho_\theta=\rho\,|\,U_\theta$. For any $\RR$-orbit $\omega\subset\RR^n$, we consider the stabilizer $\dot{H}_\omega=\{t\in\RR:A(t)^\top\xi=\xi\}$ whose definition does not depend on the choice of the point $\xi\in\omega$. Since $t\mapsto A(t)^\top$ is the coadjoint representation of an exponential Lie group, we know that the closed subgroup $\dot{H}_\omega$ is connected, see p.\ 49 of~\cite{LL}. Thus there are only two possibilities, either $\dot{H}_\omega=\RR$ or $\dot{H}_{\omega}=\{0\}$. Suppose that $\theta\subset\Sigma_{\eps'}$ is a $\ZZ$-orbit such that $\dot{H}_\omega=\RR$ where $\omega$ is the unique $\RR$-orbit containing~$\theta$. Then $\omega=\theta=\{\xi\}$ is a fixed point. We claim that $(T_\xi\cdot\ph)(t)=\widehat{\ph}(\xi,t)$ gives a unitary isomorphism of $U_{\theta}$ onto $L^2(\RR,\dot{\eps})$. First \[\widehat{\ph}(\xi,l+t)=\widehat{\ph}(\,A(-l)^\top\xi,l+t\,) =e^{i\pi\dot{\eps}(l)}\,\widehat{\ph}(\xi,t)\] by Lemma~\ref{lm:eps_equiv} which shows $T_\xi\cdot\ph\in L^2(\RR,\dot{\eps})$ for $\ph\in U_{\theta}$. Clearly $T_\xi$ is linear and \[|\,T_\xi\cdot\ph\,|_{L^2}^2=\int_0^1|\,\widehat{\ph}(\xi,t)\,|^2\;dt =|\,\ph\,|_{L^2}^2\] by Proposition~\ref{pr:Plancherel}. If $\psi\in C(\RR,\dot{\eps})$, then $\ph(x,t)=\psi(t)\,e^{\pi i\la\xi,x\ra}$ is in $C(G,\eps)$ by Lemma~\ref{lm:Sigma}, and $T_\xi\cdot\ph=\psi$. Thus $T_\xi$ is onto. From Equation~(\ref{eq:rho_xi}) it follows that the representation $\rho_\xi(x,s)=T_\xi\,\rho_\theta(x,s)\,T^\ast_\xi$ on $L^2(\RR,\dot{\eps})$ is given by \[\rho_\xi(x,s)\psi\,(t)=e^{\pi i\la \xi,x\ra}\;\psi(t+s)\;.\] Define $\eps^\sharp(t)=e^{\pi i\dot{\eps}(1)t}$ for $t\in\RR$. Then $\eps^\sharp$ is a unitary character of $\RR$ extending $\dot{\eps}$. Any extension has the form $t\mapsto\eps^{\sharp}(t)\,e^{2\pi imt}$ for some $m\in\ZZ$. By means of the unitary isomorphism $(U\cdot\psi)(t)=\eps^\sharp(-t)\,\psi(t)$ of $L^2(\RR,\eps)$ onto $L^2(\TT)$, we see that $\rho_\xi$ is unitarily equivalent to \begin{equation}\label{eq:rho_xi_for_fixed_points} \tilde{\rho}_\xi(x,s)\psi\;(t)=e^{\pi i\la\xi,x\ra}\;\eps^\sharp(s)\; \psi(t+s) \end{equation} on $L^2(\TT)$. For $m\in\ZZ$ we consider the unitary character of $G$ given by \[\chi_{\eps^{\sharp},\omega,m}(x,s)=e^{\pi i\la\xi,x\ra}\;e^{2\pi ims}\; \eps^\sharp(s).\] Finally, using the Fourier transformation and the Plancherel theorem for $L^2$-functions on the torus, we conclude that $\rho_\theta$ is unitarily equivalent to an orthogonal sum \[\rho_\theta\;\cong\;\bigoplus_{m\in{\Bbb Z}}\chi_{\eps^{\sharp},\omega,m}\] of $1$-dimensional subrepresentations. Note that up to isomorphism this decomposition does not depend on the choice of the extension $\eps^\sharp$. Now we suppose that $\theta\subset\Sigma_{\eps'}$ is a $\ZZ$-orbit such that $\dot{H}_\omega=\{0\}$. This means that $\theta$ is an infinite set and that the unique $\RR$-orbit $\omega$ containing $\theta$ is not relatively compact. In this case we have \begin{lm}\label{lm:rho_eta} For every $\eta\in\omega$ there exists a unitary isomorphism $T_\eta$ of~$U_\theta$ onto~$L^2(\RR)$ which intertwines $\rho_\theta$ and \begin{equation}\label{eq:rho_eta} \rho_\eta(x,s)\psi\;(t)=e^{\pi i\la A(t)^\top\eta,x\ra}\;\psi(t+s)\;. \end{equation} \end{lm} \begin{proof} Let $\xi\in\theta$ and $r\in\RR$ such that $\eta=A(r)^\top\xi$. We claim that $(T_\eta\ph)(t)=\widehat\ph(\xi,t+r)$ is a unitary isomorphism satisfying our needs: First of all, Proposition~\ref{pr:Plancherel} implies \begin{eqnarray*} |\,T_\eta\ph\,|_{L^2}^2 &=& \int_{-\infty}^{+\infty}\,|\,\widehat{\ph}(\xi,t+r)\,|^2\;dt =\int_{-\infty}^{+\infty}\,|\,\widehat{\ph}(\xi,t)\,|^2\;dt\\ &=& \sum_{l\in\Bbb Z}\;\int_0^1\,|\widehat{\ph}(\xi,l+t)\,|^2\;dt = \sum_{l\in\Bbb Z}\;\int_0^1\,|\widehat{\ph}(A(l)^\top\xi,t)\,|^2\;dt=|\,\ph\,|_{L^2}^2 \end{eqnarray*} which shows that $T_\eta\ph\in L^2(\RR)$ is well-defined and that $T_\eta$ is isometric. If $\psi\in C_0(\RR)$, then the sum \[\ph(x,t)=\sum_{l\in\Bbb Z}e^{-i\pi\dot{\eps}(l)}\;\psi(l+t-r)\; e^{\pi i\la A(l)^\top\xi,x\ra}\] is locally finite in $t$ and $\ph\in C(G,\eps)$ by Lemma~\ref{lm:Sigma}. Using $A(l)^\top\xi\neq\xi$ for $l\neq 0$, we conclude $T_\eta\ph=\psi$. This proves $T_\eta$ to be surjective. Finally, we observe that (\ref{eq:rho_eta}) is a consequence of (\ref{eq:rho_xi}). \qed \end{proof} Let $X_{\eps'}^\infty$ denote the set of all $\RR$-orbits which intersect the subset $\Sigma_{\eps'}$ of $\ZZ^n$ and which are not relatively compact. Let $X_{\eps'}^0$ be the set of all orbits of the form $\omega=\{\,\xi\,\}$ for a fixed point $\xi\in\Sigma_{\eps'}$. If $\omega\in X_{\eps'}^\infty$, $\eta\in\omega$ is arbitrary, and $\theta_1,\theta_2$ are $\ZZ$-orbits contained in $\omega\cap\Sigma_{\eps'}$, then $\rho_{\theta_1}\cong\rho_\eta\cong\rho_{\theta_2}$ by Lemma~\ref{lm:rho_eta}. This implies \[\bigoplus_{\theta\in{\Bbb Z}\setminus(\omega\cap\Sigma_{\eps'})}\rho_\theta\; \cong\;m_{\eps,\omega}\,\rho_{\omega}\] where $m_{\eps,\omega}$ is the number of $\ZZ$-orbits contained in $\omega\cap\Sigma_{\eps'}$, which is apriori known to be finite, and $\rho_\omega$ is the common unitary equivalence class of the representations $\rho_\theta$ for $\theta\in\ZZ\setminus(\omega\cap\Sigma_{\eps'})$. Summing up the preceding conclusions, we obtain \begin{theo}\label{theo:decomp} Let $A(t)$ be one-parameter group of $\GL(n,\RR)$ with $A(1)\in\SL(n,\ZZ)$ and such that $G=\RR^n\rtimes_A\RR$ is exponential. Let $\eps:\Gamma\to\ZZ_2$ be a homomorphism. Then the right regular representation $\rho$ of $G$ in~$L^2(G,\epsilon)$ decomposes as follows: \[\rho\;\cong\;\bigoplus_{\theta\in{\Bbb Z}\setminus\Sigma_{\eps'}}\rho_\theta\; \cong\;\left(\,\bigoplus_{\omega\in X_{\eps'}^0}\bigoplus_{m\in{\Bbb Z}} \chi_{\dot{\eps},\omega,m}\,\right)\;\bigoplus\;\left(\, \bigoplus_{\omega\in X_{\eps'}^\infty}m_{\eps,\omega}\;\rho_\omega\,\right)\] where the multiplicities $m_{\epsilon,\omega}=\#\ZZ\setminus(\omega\cap\Sigma_{\epsilon'})$ are finite, the \[\chi_{\dot{\eps},\omega,m}(x,s)=e^{\pi i\la\xi,x\ra}\;e^{\pi i(\,2m+\dot{\eps}(1)\,)s}\] are characters of $G$, and the $\rho_\omega$ are irreducible on $L^2(\RR)$. For every $\eta\in\omega$, \[\rho_\eta(x,s)\psi\;(t)=e^{\pi i\la A(t)^\top\eta,x\ra}\;\psi(t+s)\] is a representative for the unitary equivalence class of $\rho_\omega$. Moreover, the representations $\{\,\chi_{\dot{\eps},\omega,m}:\omega\in X_{\eps'}^0, m\in\ZZ\} \cup\{\,\rho_\omega:\omega\in X_{\eps'}^\infty\,\}$ are mutually inequivalent. \end{theo} \begin{proof} It remains to verify the last assertion and the irreducibility of $\rho_\omega$. Clearly characters are unitarily equivalent if and only if they are equal, and not unitarily equivalent to a representation on~$L^2(\RR)$. Let $C^\ast(N)$ be the enveloping $C^\ast$-algebra of the group algebra $L^1(N)$ of $N=\RR^n\times\{0\}$. Recall that $C^\ast(N)$ is isomorphic to $C_\infty(\widehat{N})$ via Fourier transformation. The above formula for $\rho_\eta$ shows that the $C^\ast$-kernel of the integrated form of $\rho_\omega\,|\,N$ consists of all $g\in C^\ast(N)$ whose Fourier transform vanishes on $\omega$. Since the $\RR$-orbits are locally closed, it follows that $\rho_{\omega_1}$ and $\rho_{\omega_2}$ are inequivalent whenever $\omega_1\neq\omega_2$. -- If $U$ is a closed $\rho_\eta(G)$-invariant subspace of $L^2(\RR)$, then $U$ is invariant under translations and multiplication by bounded continuous functions. Thus it follows $U=\{0\}$ or $U=L^2(\RR)$. \qed \end{proof} \begin{lm}\label{lm:range_of_T_eta} Suppose that $G=\RR^n\rtimes_A\RR$ is a nilpotent Lie group. Let $\omega\in X_{\eps'}^\infty$ and $\theta\in\ZZ\setminus(\omega\cap\Sigma_{\eps'})$. For $\eta\in\omega$ let $T_\eta$ denote the unitary isomorphism of $U_\theta$ onto $L^2(\RR)$ defined in the proof of Lemma~\ref{lm:rho_eta}. Then for every Schwartz function $\psi\in\cS(\RR)$ there exists a (unique) $\ph\in U_\theta\cap C^\infty(G,\eps)$ such that $T_\eta\ph=\psi$. \end{lm} \begin{proof} Given $\psi\in\cS(\RR)$ we consider \[\ph(x,t)=\sum_{l\in{\Bbb Z}}e^{-i\pi\dot{\eps}(l)}\;\psi(l+t-r)\; e^{\pi i\la A(l)^\top\xi,x\ra}\] whose formal derivatives are \[(\partial_t^\alpha\partial_x^\beta\ph)(x,t)=\sum_{l\in{\Bbb Z}}e^{-i\pi\dot{\eps}(l)}\; (\pi i)^{|\beta|}\;(A(l)^\top\xi)^\beta\;(\partial_t^\alpha\psi)(l+t-r)\; e^{\pi i\la A(l)^\top\xi,x\ra}\;.\] Since $B$ is nilpotent, the expression $A(l)^\top\xi=\exp(lB)^\top\xi$ is polynomial in $l$. Hence for each multi-index $\beta$ there exist constants $N\in\NN$ and $C_0>0$ such that \[|\,(A(l)^\top\xi)^\beta\,|\le C_0\,(\,1+l^2\,)^N\;.\] for all $l\in\ZZ$. On the other hand, since $\psi\in\cS(\RR)$, for each $\alpha$ there are $C_1,C_2>0$ such that \[|\,(\partial_t^\alpha\psi)(l+t-r)\,|\le C_1\,|\,1+(l+t-r)^2\,|^{-(N+1)}\le C_2\,|\,1+l^2\,|^{-(N+1)}\] for all $l$, and for $t$ ranging over a compact subset $K$ of $\RR$. This implies that the above series converge absolutely and uniformly on $\RR^n\times K$ so that $\ph\in C^\infty(G,\eps)$ is well-defined. Clearly $\ph\in U_\theta$ and $T_\eta\ph=\psi$. \qed \end{proof} \subsection{Sub-Dirac operators with discrete spectrum} Let $(H,\la\cdot,\cdot\ra)$ be a real vector space with inner product and $(\Delta,\la\cdot,\cdot\ra)$ a complex vector space with a hermitian inner product. Suppose that $\Delta$ carries a $\Cl(H)$-module structure such that $\la x\cdot v,w\ra=-\la v,x\cdot w\ra$ for all $x\in H\subset\Cl(H)$ and $v,w\in \Delta$. Let $s_1,\ldots,s_d$ be an orthonormal basis of $H$ and $a\in H$ a non-zero multiple of $s_d$. Furthermore, let $\Omega:\RR\to\Span\{s_1, \ldots,s_{d-1}\}$ be a non-constant polynomial function. We consider the operator \[P=a\,\del_t+i\,\Omega(t)\] on the domain $\cS(\RR,\Delta)$. Here $a,\Omega(t)\in\Cl(H)$ are understood as operators acting by pointwise multiplication. Clearly $P$ is symmetric with respect to the $L^2$-inner product and densely defined in the Hilbert space $L^2(\RR,\Delta)$. Thus $P$ is closable. The closure $\bar{P}$ of $P$ is a symmetric operator. Let $P^\ast$ denote the adjoint of $P$. On its domain \[\dom(P^\ast)=\{\,\psi\in L^2(\RR,\Delta): \ph\mapsto\la P\ph,\psi\ra_{L^2} \mbox{ is continuous w.r.t.\ to the }L^2\mbox{-norm}\,\}\] we consider the norm $|\cdot|_P$ given by $|\,\psi\,|_P^2=|\,\psi\,|_{L^2}^2 +|\,P^\ast\psi\,|_{L^2}^2$. Our aim is to prove the following result. \begin{pr}\label{pr:discrete_spectrum} The operator $P$ is essentially self-adjoint and its closure $\bar{P}$ has discrete spectrum. \end{pr} \begin{proof} We can assume $|a|=1$ what will simplify the estimates below. To prove the first assertion, we imitate the proof of the essential selfadjointness of the Dirac operator, compare Theorem~5.7 of~\cite{LM} and Proposition~1.3.5 of~\cite{G}. As a basic fact we know that it suffices to verify $\ker(P^\ast\pm iI)=\{0\}$. Moreover, since $\bar{P}$ is symmetric, it is enough to show that $\ker(P^\ast\pm iI)\subset\dom(\bar{P})$. To begin with, we note that, if $f\in\cS(\RR)$ and $\psi\in\dom(P^\ast)$, then $f\psi\in\dom(P^\ast)$ and \[P^\ast(f\psi)=f(P^\ast\psi)+(\del_tf)a\cdot\psi\;.\] Let $\psi\in\ker(P^\ast\pm iI)$. If $\hat{P}$ denotes the extension of $P$ to tempered distributions, then we get, as $P$ is symmetric, $(\hat{P}\pm iI)\psi=P^\ast\psi\pm i\psi=0$. Since the principal symbol $p(\xi)=\xi\,a$ of $\hat{P}\pm iI$ is invertible for $\xi\neq 0$, the regularity theorem for elliptic differential operators implies that $\psi$ is a smooth function. Choose $h\in C_0^\infty(\RR)$ satisfying $0\le h\le 1$ and $h(0)=1$, and put $h_k(t)=h(t/k)$ for $k\ge 1$. By definition $h_k\to 1$ and $h_k\psi\in C_0^\infty(\RR,\Delta)\subset\dom(P)$. Since \[\Big|\,(\del_th_k)a\cdot\psi\,\Big|_{L^2}\le|\,\del_th_k\,|_\infty\,|\,a\,|\; |\,\psi\,|_{L^2}\le\frac{1}{k}\,|\,\del_th\,|_\infty\,|\,\psi\,|_{L^2}\;,\] it follows that \begin{eqnarray*} |\,\psi-h_k\psi\,|_P^2 &=& |\psi-h_k\psi|_{L^2}^2 +|P^\ast\psi-P^\ast(h_k\psi)|_{L^2}^2\\ &\le& |\psi-h_k\psi|_{L^2}^2+\left(\;|P^\ast\psi-h_k(P^\ast\psi)|_{L^2} +|\,(\del_th_k)a\cdot\psi\,|_{L^2}\;\right)^2 \end{eqnarray*} converges to $0$ for $k\to\infty$ by dominated convergence. Hence $\psi\in\dom(\bar{P})$. This establishes the essential selfadjointness of $P$. To prove that the spectrum of $\bar{P}$ is discrete, we need the following two lemmata. \begin{lm}\label{lm:cp_emb} Let $(\Delta,\la\cdot,\cdot\ra)$ be a $\Cl(H)$-module as above and $\Omega:\RR\to H$ a continuous function satisfying $|\,\Omega(t)\,|\to\infty$ for $|\,t\,|\to\infty$. Then \[X:=\{\,\ph\in L^2(\RR,\Delta):\del_t\ph\in L^2(\RR,\Delta)\mbox{ and } \Omega\cdot\ph\in L^2(\RR,\Delta)\,\}\] becomes a Hilbert space when endowed with the norm $||\,\ph\,||^2=|\ph|_{L^2}^2 +|\del_t\ph|_{L^2}^2+|\Omega\cdot\ph|_{L^2}^2$, and the inclusion $X\to L^2(\RR,\Delta)$ is a compact operator. \end{lm} \begin{proof} The first assertion is obvious. Let $(\ph_m)$ be a bounded sequence in $X$. We prove that $(\ph_m)$ has a subsequence which is Cauchy in $L^2(\RR,\Delta)$. For every $n\in\NN\setminus\{0\}$ there exists $r_n>0$ such that $|\Omega(t)|\ge n$ for $|t|\ge r_n$. We can assume $r_n<r_{n+1}$ and $r_n\to\infty$ for $n\to\infty$. Using $|\Omega(t)\cdot\ph(t)|=|\Omega(t)|\,|\ph(t)|\ge n\,|\ph(t)|$, we obtain \[\int_{|t|\ge r_n}|\ph(t)|^2\,dt\le\frac{1}{n^2}\int_{|t|\ge r_n} |\Omega(t)\cdot\ph(t)|^2\,dt\le\frac{1}{n^2}\,|\,\Omega\cdot\ph\,|_{L^2}^2\] for $\ph\in X$. For the moment, we fix the parameter $n$. Let $\chi\in C_0^\infty(\RR)$ be such that $0\le\chi\le 1$ and $\chi(t)=1$ for $|t|\le r_n$. By Rellich's theorem, applied to the bounded sequence~$\chi\ph_m$ in~$H^1(\RR,\Delta)$, we conclude that there exists a subsequence $(\ph_{m_{n,k}})$ such that \[\int_{-r_n}^{r_n}|\ph_{m_{n,k}}(t)-\ph_{m_{n,l}}(t)|^2\,dt\to 0\] for $k,l\to\infty$. Proceeding by induction, we establish $\{m_{n+1,k}:k\in\NN\} \subset\{m_{n,k}:k\in\NN\}$ for all $n$. We define $m_k=m_{k,k}$. Now it is easy to see that $(\ph_{m_k})$ is Cauchy w.r.t.\ the $L^2$-norm.\qed \end{proof} \begin{lm}\label{lm:dom_of_Dast} The domain of $\bar{P}$ is contained in $X$ and the inclusion $\dom(\bar{P})\to X$ is continuous. \end{lm} \begin{proof} Since $\dom(P)=\cS(\RR,\Delta)$ is contained in $X$ and dense in $\dom(\bar{P})$ w.r.t.\ the norm $|-|_P$, it suffices to prove that there exists $K>0$ such that $||\,\ph\,||\le K\,|\,\ph\,|_{P}$ for all $\ph\in\cS(\RR,\Delta)$. Using $a\cdot(\del_t\ph)=\del_t(a\cdot\ph)$ and $\Omega a=-a\Omega$ in $\Cl(H)$, we compute \[\la\,a\cdot(\del_t\ph),i\Omega\cdot\ph\,\ra_{L^2} =-\la\,a\cdot\ph,i\del_t(\Omega\cdot\ph)\,\ra_{L^2}\] and \[\la\,i\Omega\cdot\ph,a\cdot(\del_t\ph)\,\ra_{L^2} =\la\,a\cdot\ph,i\Omega\cdot(\del_t\ph)\,\ra_{L^2}\;.\] As $\del_t(\Omega\cdot\ph)=(\del_t\Omega)\cdot\ph+\Omega\cdot(\del_t\ph)$, it follows \[|\,P\ph\,|_{L^2}^2 = |a\cdot(\del_t\ph)+i\Omega\cdot\ph|_{L^2}^2 =|\,\del_t\ph\,|_{L^2}^2-\la\,a\cdot\ph,i(\del_t\Omega)\cdot\ph\,\ra_{L^2} +|\,\Omega\cdot\ph\,|_{L^2}^2\] for all Schwartz functions. Since $\Omega$ is a polynomial function, there exists $r>0$ such that $|\,(\del_t\Omega)(t)\,|\le|\,\Omega(t)\,|$ for all $|t|\ge r$. Fix $C>1$ such that $|\,(\del_t\Omega)(t)\,|\le C$ for $|t|\le r$. From \begin{eqnarray*} |\,\la\,a\cdot\ph,i(\del_t\Omega)\cdot\ph\,\ra_{L^2}\,| &\le& \int_{-r}^r|\,(\del_t\Omega)(t)|\,|\ph(t)|^2\;dt +\int_{|t|\ge r}|(\del_t\Omega)(t)|\,|\ph(t)|^2\;dt\\ &\le& C\int_{-r}^r|\ph(t)|^2\;dt+\int_{|t|\ge r}|\ph(t)|\,|\Omega(t)\cdot\ph(t)|\;dt\\ &\le& C\,|\,\ph\,|_{L^2}^2+|\,\ph\,|_{L^2}\,|\,\Omega\cdot\ph\,|_{L^2} \end{eqnarray*} it then follows \begin{eqnarray*} |\,\ph\,|_P^2 &=& |\,\ph\,|_{L^2}^2+|\,P\ph\,|_{L^2}^2\ge \frac{1}{2C}\left(\,C\,|\,\ph\,|_{L^2}^2+\frac{1}{4}\,|\,\ph\,|_{L^2}^2 +|\,P\ph\,|_{L^2}^2\,\right)\\ &\ge& \frac{1}{2C}\left(\,\frac{1}{4}\,|\,\ph\,|_{L^2}^2+|\,\del_t\ph\,|_{L^2}^2 +|\,\Omega\cdot\ph\,|_{L^2}^2-|\,\ph\,|_{L^2}\,|\Omega\cdot\ph\,|_{L^2}\,\right) \ge\frac{1}{2C}\,|\,\del_t\ph\,|_{L^2}^2 \end{eqnarray*} for $\ph\in\cS(\RR,\Delta)$. Moreover, $i\Omega\cdot\ph=P\ph-a\cdot(\del_t\ph)$ gives \[|\,\Omega\cdot\ph\,|_{L^2}\le |\,P\ph\,|_{L^2}+|\,\del_t\ph\,|_{L^2} \le(1+\sqrt{2C})\,|\,\ph\,|_P\;.\] Altogether, we obtain \[||\,\ph\,||^2 = |\,\ph\,|_{L^2}^2+|\,\del_t\ph\,|_{L^2}^2+|\,\Omega\cdot\ph\,|_{L^2}^2 \le \left(\,1+2C+(1+\sqrt{2C})^2\,\right)|\,\ph\,|_P^2\] proving the lemma.\qed \end{proof} Now we can prove the second assertion of Proposition~\ref{pr:discrete_spectrum}. As $\sigma(\bar{P})\subset\RR$, we know that $\bar{P}-iI$ is bijective. Put $R:=(\bar{P}-iI)^{-1}$. Since $\bar{P}-iI$ is continuous w.r.t.\ the complete norm $|\,\cdot\,|_P$ on $\dom(\bar{P})$ and the $L^2$-norm, the open-mapping theorem implies that $R:L^2(\RR,\Delta)\to\dom(\bar{P})$ is continuous. Moreover, since the inclusion $\dom(\bar{P})\to X\to L^2(\RR,\Delta)$ is compact by Lemma~\ref{lm:cp_emb} and~\ref{lm:dom_of_Dast}, it follows that $R$ is a compact normal operator on $L^2(\RR,\Delta)$ with $\ker R=\{0\}$. By the spectral theorem there exists an orthonormal basis $\{\ph_n:n\in\NN\}$ of $L^2(\RR,\Delta)$ with $R\ph_n=\mu_n\ph_n$ for suitable $0\neq\mu_n\in\CC$. This implies $\bar{P}\ph_n=(\lambda+\frac{1}{\mu_n})\ph_n$ for all $n$. If $\{\mu_n:n\in\NN\}$ happens to be an infinite set, then $\mu_n\to 0$ and hence $|\lambda+\frac{1}{\mu_n}|\to\infty$ for $n\to\infty$. This proves the spectrum of $\bar{P}$ to be discrete. \qed \end{proof} Now we resume the assumptions of Section~\ref{S3.1}: Let $A(t)=\exp(tB)$ be a one-parameter group of $\GL(n,\RR)$ with $A(1)\in\SL(n,\ZZ)$. Suppose that $B$ does not possess any purely imaginary eigenvalues. Let $\fg$ denote the Lie algebra of~$G$, and $\fn$ the Lie algebra of~$N=\RR^n\times\{0\}$. As before we identify $\fg$ with the tangent space at the identity element $e=(0,0)$ of $G$ and denote by $b$ the $(n+1)$th basis vector of $\fg\cong\RR^n\rtimes_B\RR$. Let $\cH$ be a left-invariant oriented distribution on $G$ such that $b\in\cH_e$. Suppose that $\cH$ carries a left-invariant Riemannian metric $g$ such that $b$ is orthogonal to $\cH_e\cap\fn$ w.r.t.\ the inner product $\ip:=g_e$ on $\cH_e$. We assume that $\cH$ is bracket-generating. With $C^0(\cH_e)=\cH_e$ and $C^k(\cH_e)=[\cH_e,C^{k-1}(\cH_e)]$ for $k\ge 1$, this means $\cH=\sum_{k=0}^n C^k(\cH_e)$. In particular, it follows \begin{equation}\label{eq:weak_bg} [\fg,\fg]=\sum_{k=1}^n C^k(\cH_e)\;. \end{equation} The latter condition is crucial for the proof of Theorem~\ref{theo:pure_point_spec} but we do not claim that (\ref{eq:weak_bg}) has significance for left-invariant distributions $\cH$ on Lie groups $G$ which do not have the form $G=\RR^n\rtimes_A\RR$. Let $\nabla$ be a left-invariant metric connection on $\cH$ satisfying condition~(\ref{Ediv}) of Lemma~\ref{lm:sym_of_D}, which guarantees the symmetry of the sub-Dirac operator. Here the divergence in~(\ref{Ediv}) is defined w.r.t.\ the left-invariant volume form corresponding to the Haar measure of~$G$. If $\dot{\cH}$ is the distribution on~$\GG$ defined by~$\cH$, then the Riemannian metric on $\dot{\cH}$ induced by $g$ is denoted by $\dot{g}$, and the connection on $\dot{\cH}$ by $\dot{\nabla}$. Let $\eps:\Gamma\to\ZZ_2$ be a homomorphism defining a spin structure $P_{\Spin,\eps}(\dot{\cH})\cong G\times_\Gamma\Spin(d)$ of $(\dot{\cH},\dot{g})$, where $d=\dim\cH$. We fix a positively oriented orthonormal basis $s_1,\ldots,s_d$ of $\cH_e$ with $s_1,\ldots,s_{d-1}\in\cH_e\cap\fn$ and such that $s_d$ is a positive multiple of~$b$. Denoting the corresponding left-invariant vector fields again by $s_1,\ldots,s_d$, the sub-Dirac operator, as acting on smooth sections of the spinor bundle $S(\dot{\cH},\eps)$, is given by $D=\sum_i s_i\cdot\dot{\nabla}^S_{s_i}$. Let $\Delta$ be the representation space of the complex spinor representation of~$\Spin(\cH_e)$. Identifying $\Gamma(S(\dot{\cH},\eps))$ with $C^\infty(G,\eps)\otimes\Delta$, we see that $D$ is given by \begin{equation}\label{eq:expl_form_of_D} D=\sum_i d\rho(s_i)\otimes s_i+\frac{1}{4}\sum_{i,j,k}\Gamma_{ij}^k\; I\otimes s_is_js_k=:P+W \end{equation} where the $s_i$'s in the second factor of the tensor products are understood as operators on~$\Delta$. Furthermore, the constants $\Gamma_{ij}^k=g(\nabla_{s_i}s_j,s_k)$ are the Christoffel symbols of~$\nabla$ w.r.t.\ the orthonormal frame $s_1,\ldots,s_d$ of $\cH$, and $d\rho$ is the derivative of the right regular representation $\rho$ of~$G$ on~$L^2(G,\eps)$. By~(\ref{Ediv}) the second sum in~(\ref{eq:expl_form_of_D}) reduces to a sum over all pairwise distinct indices $i,j,k$. \begin{theo}\label{theo:pure_point_spec} If, in addition to the above assumptions, $G=\RR^n\rtimes_A\RR$ is nilpotent, then the closure of the operator $D$ on $\GG$ has a pure point spectrum. \end{theo} \begin{proof} By Proposition~\ref{pr:L2_decomp} we know that $L^2(G,\eps)\otimes\Delta$ is a direct sum of the orthogonal subspaces $\{U_\theta\otimes\Delta:\theta\in\ZZ\setminus \Sigma_{\eps'}\}$ which are invariant under the action of~$\rho(G)\otimes\Cl(\cH_e)$. Let $U_\theta^\infty:=U_\theta\cap C^\infty(G,\eps)$. Then $U_\theta^\infty\otimes\Delta$ is $D$-invariant. To prove that $\bar{D}$ has a pure point spectrum, it suffices to prove that the closure of~$D_\theta:=D\,|\,U_\theta^\infty\otimes\Delta$ has a pure point spectrum for all $\theta$. If $\theta=\{\xi\}$ is a fixed point, then, according to Equation~(\ref{eq:rho_xi_for_fixed_points}), the subspace $U_\theta$ is an orthogonal sum of one-dimensional $\rho(G)$-invariant subspaces of~$U_\theta^\infty$. Thus $U_\theta\otimes\Delta$ is an orthogonal sum of two-dimensional $D_\theta$-invariant subspaces of $\dom(D_\theta)$ which shows that $D_\theta$ has a pure point spectrum. Thus we are left with the case where $\theta$ is an infinite set and $U_\theta$ is isomorphic to~$L^2(\RR)$. Fix $\xi\in\theta$. Lemma~\ref{lm:rho_eta} implies that there exists a unitary isomorphism $T_\xi:U_\theta\to L^2(\RR)$ such that $\rho_\xi:=T_\xi\,\rho_\theta\,T_\xi^\ast$ is given by \begin{equation}\label{eq:rho_xi_again} \rho_\xi(x,s)\psi\;(t)=e^{\pi i\la A(t)^\top\xi,x\ra}\;\psi(t+s). \end{equation} By Lemma~\ref{lm:range_of_T_eta} we may define $D_\xi=(T_\xi\otimes I)\,D_\theta\, (T_\xi^\ast\otimes I)\,|\,\cS(\RR,\Delta)$. Since $d\rho_\xi(s_j)=\pi i\,\la A(t)^\top\xi,s_j\ra$ for $1\le j\le d-1$ and $d\rho_\xi(s_d)=|\,b\,|\,\del_t$, it follows that the operator $P_\xi:=\sum_{j=1}^d d\rho_\xi(s_j)\otimes s_j$ on $\cS(\RR,\Delta)$ has the form \begin{equation}\label{eq:D_xi} P_\xi=a\del_t+i\Omega_\xi(t) \end{equation} with $a=|\,b\,|^{-1}s_d$ and $\Omega_\xi(t)=\pi\,\sum_{j=1}^{d-1}\,\la A(t)^\top\xi,s_j\ra\,s_j\in\cH_e\subset\Cl(\cH_e)$. Note that $\la\,A(t)^\top\xi,s_j\,\ra=\la\xi,\exp(tB)s_j\,\ra$ is a polynomial in $t$ because $B$ is nilpotent. Since \[ [\fg,\fg]=\sum_{k=1}^nB^k(\cH_e\cap\fn) =\sum_{k=1}^n\sum_{j=1}^{d-1}\,\RR\cdot(B^ks_j)\] by~(\ref{eq:weak_bg}) and $\la\,\xi,[\fg,\fg]\,\ra\neq 0$ for non-fixed points, it follows that at least one of the components of $\Omega_\xi$ is not constant. Thus Proposition~\ref{pr:discrete_spectrum} implies that $\bar{P}_\xi$ has discrete spectrum. In other words, the essential spectrum of $\bar{P}_\xi$ is empty. The operator $W_\xi:=\frac{1}{4}\sum_{i,j,k}\Gamma_{ij}^k\,s_is_js_k$ is bounded on $L^2(\RR,\Delta)$. In particular, $W_\xi$ is relatively $\bar{P}_{\xi}$-compact in the sense that $\dom(\bar{P}_\xi)\subset\dom(W_\xi)$ and $W_\xi(\bar{P}_\xi-iI)^{-1}$ is compact. By the Kato-Rellich theorem we know that $\bar{D}_\xi=\bar{P}_\xi+W_\xi$ is selfadjoint. Moreover, Weyl's theorem which asserts the stability of the essential spectrum under relatively compact perturbations, and for which we refer to Theorem~14.6 of~\cite{HS}, implies that the essential spectrum of $\bar{D}_\xi=\bar{P}_\xi+W_\xi$ is empty. This shows that $\bar{D}_\xi$ and hence $\bar{D}_\theta$ have discrete spectrum. The proof of the theorem is complete.\qed \end{proof} \begin{re} In general, the eigenvalues of the sub-Dirac operator $D$ do not have finite multiplicity and the spectrum of $D$ is not a discrete subset of $\RR$. A relevant example is given in Section~\ref{S4.3}. \end{re} \subsection{Two- and three-dimensional distributions} In this subsection we will compute the spectra of the operators $D_\theta$ arising in the proof of Theorem~\ref{theo:pure_point_spec} provided that $G=\RR^n\rtimes_A\RR$ is 2-step nilpotent and $\dim\cH=2$ or~$3$. The explicit formulas that will be given below in the non-fixed point case are a consequence of the following result. \begin{pr}\label{pr:spectrum_of_D} Let $\alpha,\beta\in\RR$ and $\omega(t)=a\omega_1t+\omega_0$ where $a>0$, $\omega_0,\omega_1\in\CC$ and $|\omega_1|=1$. Then the spectrum of the operator \begin{equation}\label{eq:concrete_D} D=\alpha I+\beta\,\left(\begin{array}{cc} i\del_t & \bar{\omega} \\ \omega & -i\del_t \end{array}\right) \end{equation} on $\cS(\RR,\CC^2)$ is discrete. More precisely, $\sigma(D)=\{\,\lambda_0\,\}\cup \{\,\lambda_{k}^{\pm}:k\in\NN\setminus\{0\}\,\}$, where \[\lambda_0:=\alpha+\beta\Im(\omega_0\bar{\omega}_1)\quad{and}\quad \lambda_k^{\pm}:=\alpha\pm\beta\left(2ak+\Im(\omega_0\bar{\omega}_1)^2\right)^{1/2}.\] If $\lambda_0$ and the $\lambda_k^\pm$ are pairwise distinct, then all eigenvalues are simple. \end{pr} \begin{proof} We can assume $\alpha=0$ and $\beta=1$. Instead of $D$, we consider the operator $S:=Q^*DQ$, where $Q$ is the unitary matrix \[Q=\frac{1}{\sqrt{2}}\;\left(\begin{array}{cc} 1 & -i\bar{\omega}_1 \\ -i\omega_1 & 1 \end{array}\right),\] which diagonalizes $D^2$ and does not depend on $t$. Obviously the spectra of $D$ and $S$ coincide. We have \begin{eqnarray*} S &=& \left(\begin{array}{cc} \Im(\omega_1\bar{\omega}) & \bar{\omega}_1 \left(\del_t+\Re(\omega_1\bar{\omega})\,\right)\; \\ \;\omega_1\left(-\del_t+ \Re(\omega_1\bar{\omega})\,\right) & -\Im(\omega_1\bar{\omega})\end{array}\right)\\[0.4cm] &=& \left(\begin{array}{cc} -\Im(\omega_0\bar{\omega}_1) & \bar{\omega}_1\left(\,\del_t+a\,t+\Re(\omega_0\bar{\omega}_1)\,\right)\; \\ \;\omega_1\left(\,-\del_t+a\,t+\Re(\omega_0\bar{\omega}_1)\right) & \Im(\omega_0\bar{\omega}_1)\end{array}\right). \end{eqnarray*} To detect $S$-invariant subspaces, we start with the orthonormal basis $\{h_k:k\in\NN\}$ of~$L^2(\RR)$ given by the (normalized) Hermite functions \[h_k(x)=(\,2^k\,\pi^{1/2}\,k!\,)^{-1/2}\,H_k(x)\,e^{-x^2/2}.\] Here $H_k(x)=(-1)^k\,e^{x^2}\del_x^k[e^{-x^2}]$ is the $k$th Hermite polynomial. Put $b=2a\Re(\omega_0\bar\omega_1)$. Using the unitary isomorphism \[(U\cdot w)(t):=a^{1/4}\,w(\,a^{1/2}(t+\frac{b}{2a^2})\,)\] of $L^2(\RR)$, we then define $u_k=U\cdot h_k$. Recall that the creation operator $\Lambda_+=-\del_x+x$ and the annihilation operator $\Lambda_-=\del_x+x$ satisfy $\Lambda_+(h_k)=\sqrt{2(k+1)}\,h_{k+1}$ and $\Lambda_-( h_k)=\sqrt{2k}\,h_{k-1}$. Since \[U\Lambda_\pm U^\ast=a^{-1/2}\,\left(\mp \del_t+a\,t+\frac{b}{2a}\right),\] we thus obtain \begin{eqnarray*} \left(-\del_t+a\,t+\frac{b}{2a}\right)u_k &=& \sqrt{2a(k+1)}\;u_{k+1}\mbox{ for }k\ge 0,\\ \left(\del_t+a\,t+\frac{b}{2a}\right)u_k &=& \sqrt{2ak}\;u_{k-1}\mbox{ for }k\ge 1,\\ \left(\del_t+a\,t+\frac{b}{2a}\right)u_0 &=& 0. \end{eqnarray*} This shows \begin{eqnarray*} S\cdot\left(\begin{array}{c} 0 \\ u_0 \end{array}\right) &=& \left(\begin{array}{c} \bar{\omega}_1\left(\del_t+a\,t +\Re(\omega_0\bar{\omega}_1)\right)u_0 \\ \Im(\omega_0\bar{\omega}_1)\,u_0 \end{array}\right) \ =\ \Im(\omega_0\bar{\omega}_1) \left(\begin{array}{c} 0 \\ u_0 \end{array}\right),\\[1ex] S\cdot\left(\begin{array}{c} u_{k-1} \\ 0 \end{array}\right) &=& \left(\begin{array}{cc} -\Im(\omega_0\bar{\omega}_1)\,u_{k-1}\\ \omega_1\left(-\del_t+a\,t+\Re(\omega_0\bar{\omega}_1)\right)u_{k-1}\end{array}\right) \ =\ \left(\begin{array}{c} -\Im(\omega_0\bar{\omega}_1)\;u_{k-1} \\ \sqrt{2ak}\; \omega_1\;u_{k}\end{array}\right),\\[1ex] S\cdot\left(\begin{array}{c} 0 \\ u_k \end{array}\right) &=& \left(\begin{array}{c} \bar{\omega}_1\left(\del_t+a\,t+\Re(\omega_0\bar{\omega}_1) \right)u_k \\ \Im(\omega_0\bar{\omega}_1)\,u_k \end{array}\right)\ =\ \left(\begin{array}{c} \sqrt{2ak}\;\bar{\omega}_1\;u_{k-1}\\ \Im(\omega_0\bar{\omega}_1)\;u_k \end{array}\right). \end{eqnarray*} In particular, the subspaces \[V_0:=\CC\left(\begin{array}{c} 0 \\ u_0\end{array}\right)\quad\mbox{and}\quad V_k:=\Span\left\{\ph_k:=\left(\begin{array}{c} u_{k-1} \\ 0\end{array}\right), \psi_k:=\left(\begin{array}{c} 0 \\ u_k\end{array}\right)\right\},\ k\ge1,\] are $S$-invariant, and the restriction of $S$ to $V_k$, $k\ge 1$, is given by the matrix $$\left(\begin{array}{cc} -\Im(\omega_0\bar \omega_1) & \sqrt{2ak}\,\bar\omega_1 \\ \sqrt{2ak}\,\omega_1 & \Im(\omega_0\bar\omega_1) \end{array}\right)$$ with respect to the basis $\ph_k,\psi_k$. Since $L^2(\RR,\CC^2)$ is the direct sum of the $V_k$, $k\ge0$, the assertion follows.\qed \end{proof} Assume that $G=\RR^n\rtimes_A\RR$ is $2$-step nilpotent. Let $(\cH,g,\nabla)$ be as in the preceding subsection with $2\le\dim\cH\le 3$. First we will determine the spectrum of $D_\theta:=D\,|\,U_\theta^\infty\otimes\Delta$ when $\theta$ does not consist of a single point. Suppose that $\dim\cH=2$. Let $s_1\in\cH_e\cap\fn$ and $s_2\in\cH_e$ be a positive multiple of $b$ such that $s_1,s_2$ is a positively oriented orthonormal basis of~$\cH_e$. In particular, $s_2=|\,b\,|^{-1}\,b$. By (\ref{Ediv}) we have $\Gamma_{11}^1+\Gamma_{21}^2=0$ and $\Gamma_{12}^1+\Gamma_{22}^2=0$ which implies that all Christoffel symbols of $\nabla$ vanish. Up to isomorphism, there exists only one simple $\Cl(\cH_e)$-module. Let $\Delta=\CC^2$ be the one such that $s_1$ and~$s_2$, represented as operators on $\Delta$, are given by \[s_1=\left(\begin{array}{cc} 0 & -1 \\ 1 & 0 \end{array}\right)\quad\mbox{and} \quad s_2=\left(\begin{array}{cc} i & 0 \\ 0 & -i \end{array}\right)\;.\] Let $\theta\in\ZZ\setminus\Sigma_{\eps'}$ and $\xi\in\theta$ be a non-fixed point. If $T_\xi:U_\theta\to L^2(\RR)$ is a unitary isomorphim as in the proof of Theorem~\ref{theo:pure_point_spec} and $D_\xi=(T_\xi\otimes I)D_\theta(T_\xi^\ast\otimes I)\,|\,\cS(\RR,\Delta)$, then we know by~(\ref{eq:D_xi}) that $D_\xi$ has the form \[D_\xi=|b|^{-1}\left(\begin{array}{cc} i\del_t & \bar{\omega}_\xi \\ \omega_\xi & -i\del_t\end{array}\right) \] with $\omega_\xi(t)=\pi i\,|\,b\,|\,\la\,A(t)^\top\xi,s_1\,\ra$. Since $B^2=0$, we have $A(t)=I+tB$ so that \[\omega_\xi(t)=\pi i\,|\,b\,|\left(\,\la\,B^\top\xi,s_1\,\ra\,t +\la\,\xi,s_1\,\ra\,\right)\] is a non-constant affine-linear function. Thus Proposition~\ref{pr:spectrum_of_D} implies that $\bar{D}_\xi$ has discrete spectrum. Moreover, the eigenvalues of $D_\xi$ can be computed as follows: Put $\alpha=0$, $\beta=|\,b\,|^{-1}$, $a=\pi\,|\,b\,|\,|\,\la B^\top\xi,s_1\ra\,|$, $\omega_1=i\sgn\la B^\top\xi,s_1\ra$ and $\omega_0=\pi i\,\la\xi,s_1\ra$. Note that $\Im(\omega_0\bar{\omega}_1)=0$. Hence it follows that \begin{equation}\label{eq:lambda_for_2dim_H} \lambda_0(\xi)=0\qquad\mbox{and}\qquad\lambda_k^\pm(\xi) =\pm\,\left(\,2\pi\,|\,b\,|^{-1}\,|\la\,B^\top\xi,s_1\,\ra|\,k\right)^{1/2} \end{equation} with $k\in\NN\setminus\{0\}$ are the eigenvalues of $D_\xi$. This completes the case $\dim\cH=2$.\\\\ Suppose that $\dim\cH=3$. Choose $s_1,s_2\in\cH_e\cap\fn$ and $s_3=|\,b\,|^{-1}b$ such that $s_1,s_2,s_3$ becomes a positively oriented orthonormal basis of $\cH_e$. Up to isomorphism, there exist two simple $\Cl(\cH_e)$-modules. Let $\Delta=\CC^2$ be the one given by the representation \[s_1=\left(\begin{array}{cc} 0 & i \\ i & 0 \end{array}\right)\quad,\quad s_2=\left(\begin{array}{cc} 0 & -1 \\ 1 & 0 \end{array}\right)\quad,\quad s_3=\left(\begin{array}{cc} i & 0 \\ 0 & -i \end{array}\right).\] Note that $s_1s_2=s_3$ and $s_is_j+s_js_i=-2\delta_{ij}$ for all $1\le i,j\le 3$, as operators on $\Delta$. Using this and that $\nabla$ is metric, we conclude that the second sum in~(\ref{eq:expl_form_of_D}) simplifies to $W=-\frac{1}{2}(\Gamma_{12}^3+\Gamma_{23}^1+\Gamma_{31}^2)\,I\otimes I$. Let $\theta\in\ZZ\setminus\Sigma_{\eps'}$ and $\xi\in\theta$ be a non-fixed point. If $T_\xi$ is a unitary isomorphism of $U_\theta$ onto $L^2(\RR)$ such that $\rho_\xi=T_\xi\,\rho_\theta\,T_\xi^\ast$ is given by Equation~(\ref{eq:rho_xi}), then the restriction $D_\xi$ of $(T_\xi\otimes I)\,D_\theta\,(T_\xi^\ast\otimes I)$, when realized in $L^2(\RR,\Delta)$, to Schwartz functions has the form \[D_\xi=-\frac{1}{2}(\Gamma_{12}^3+\Gamma_{23}^1+\Gamma_{31}^2)\,I+ |b|^{-1}\left(\begin{array}{cc} i\del_t & \bar{\omega}_\xi \\ \omega_\xi & -i\del_t\end{array}\right) \] where $\omega_\xi$, as $G$ is $2$-step nilpotent, is a non-constant affine linear function given by \begin{align*} \begin{split} \omega_\xi(t) &= \pi i\,|\,b\,|\left(\,i\la\,A(t)^\top\xi,s_1\,\ra\, +\la\,A(t)^\top\xi,s_2\,\ra\,\right)\\ &= -\pi\,|\,b\,|\left(\,\big(\la\,B^\top\xi,s_1\,\ra\,-i\la\,B^\top\xi,s_2\,\ra\,\big)\,t +\la\,\xi,s_1\,\ra\,-i\la\,\xi,s_2,\ra\,\right)\,. \end{split} \end{align*} Hence Proposition~\ref{pr:spectrum_of_D} implies that the eigenvalues of $D_\xi$ are \begin{equation}\label{eq:lambda_0_for_3dim_H} \lambda_0(\xi)=-\frac{1}{2}(\Gamma_{12}^3+\Gamma_{23}^1+\Gamma_{31}^2) -\pi\;\frac{\la B^\top\xi,s_1\ra\la\xi,s_2\ra-\la\xi,s_1\ra\la B^\top\xi,s_2\ra} {\left(\,\la B^\top\xi,s_1\ra^2+\la B^\top\xi,s_2\ra^2\,\right)^{1/2}} \end{equation} and \begin{multline}\label{eq:lambda_k_for_3dim_H} \lambda_k^{\pm}(\xi) = -\frac{1}{2}(\Gamma_{12}^3+\Gamma_{23}^1 +\Gamma_{31}^2)\pm\bigg(\,2\pi k\,|\,b\,|^{-1}\left(\,\la B^\top\xi, s_1\ra^2+\la B^\top\xi,s_2\ra^2\,\right)^{1/2}\\ +\pi^2\;\frac{\left(\,\la B^\top\xi,s_1\ra\la\xi,s_2\ra -\la\xi,s_1\ra\la B^\top\xi,s_2\ra\,\right)^2}{\la B^\top\xi,s_1\ra^2 +\la B^\top\xi,s_2\ra^2}\;\bigg)^{1/2}. \end{multline} In~(\ref{eq:lambda_for_2dim_H}) - (\ref{eq:lambda_k_for_3dim_H}) we rediscover the fact that the eigenvalues $\lambda_k(\xi)$ do not depend on the choice of the point $\xi$ on the orbit. More precisely, since $B^2=0$ and $A(t)B=B$, it follows, in accordance with Lemma~\ref{lm:rho_eta}, that $\lambda_k^\pm(A(t)^\top\xi)=\lambda_k^\pm(\xi)$ for all $t\in\RR$ and $\xi\in\RR^n\setminus[\fg,\fg]^\bot$. This completes the case $\dim\cH=3$.\\\\ Finally we compute the spectrum of $D_\theta$ when $\theta=\{\xi\}$ is a fixed point. For this purpose, we can drop the assumption that $G$ is nilpotent. By~(\ref{eq:rho_xi_for_fixed_points}) we know that $U_\theta$ is an orthogonal sum of 1-dimensional subspaces $\{U_{\theta,k}:k\in\ZZ\}$ of~$U_\theta\cap C^\infty(G,\eps)$ on which $(x,s)\in G$ acts by multiplication with \[\chi_{\eps^\sharp,\theta,k}(x,s)=e^{\pi i\la\xi,x\ra} e^{\pi i(2k+\dot{\eps}(1))s}\;.\] Suppose that $\dim\cH=2$. Let $s_1,s_2$ and $\Delta$ be as above. In this case the sub-Dirac operator reads $D=d\rho(s_1)\otimes s_1+d\rho(s_2)\otimes s_2$. Since $d\chi_{\eps^\sharp,\theta,k}(s_1)=\pi i\,\la\,\xi,s_1\,\ra$ and $d\chi_{\eps^\sharp,\theta,k}(s_2)=\pi i\,|\,b\,|^{-1}\,(2k+\dot{\eps}(1))$, it follows that $D_{\theta,k}:=D_\theta\,|\,U_{\theta,k}\otimes\CC^2$ is unitarily equivalent to \begin{equation}\label{eq:D_theta_k} D_{\xi,k}:=\alpha I+\beta\left(\begin{array}{cc} 2k+\dot{\eps}(1) & \bar{\omega}_\xi \\ \omega_\xi & -2k-\dot{\eps}(1) \end{array} \right) \end{equation} on $\CC^2$, where $\alpha=0$, $\beta=-\pi\,|\,b\,|^{-1}$ and $\omega_\xi=-i\,|\,b\,|\,\la\,\xi,s_1\,\ra$ are constants. Obviously, $D_{\xi,k}$ admits the eigenvalues \begin{equation}\label{eq:mu_k_for_2dim_H} \mu_k^\pm(\xi)=\pm\,\pi\,\left(\,|\,b\,|^{-2}(2k+\dot{\eps}(1))^2 +\la\xi,s_1\ra^2\,\right)^{1/2}\;,\;k\in\ZZ\,. \end{equation} Suppose that $\dim\cH=3$. Let $s_1,s_2,s_3$ and $\Delta$ be as above. In this case $D=P+W$ where $W=\alpha\,I\otimes I$ and $\alpha=-\frac{1}{2}(\Gamma_{12}^3 +\Gamma_{23}^1+\Gamma_{31}^2)$. Hence $D_{\theta,k}$ is unitarily equivalent to $D_{\xi,k}$ as in~(\ref{eq:D_theta_k}) with $\beta=-\pi\,|\,b\,|^{-1}$ and $\omega_{\xi}=|\,b\,|\,(\,\la\,\xi,s_1\,\ra-i\la\,\xi,s_2\,\ra\,)$. Thus $D_{\xi,k}$ has the eigenvalues \begin{equation}\label{eq:mu_k_for_3dim_H} \mu_k^\pm(\xi)=-\frac{1}{2}(\Gamma_{12}^3+\Gamma_{23}^1+\Gamma_{31}^2) \,\pm\,\pi\,\left(\,|\,b\,|^{-2}(2k+\dot{\eps}(1))^2+\la\xi,s_1\ra^2 +\la\xi,s_2\ra^2\,\right)^{1/2}. \end{equation} This shows that $D_\theta$ has discrete spectrum in the fixed point case. \section{Examples of spectra of sub-Dirac operators} \subsection{A preliminary remark} To compute the spectrum of the sub-Dirac operator $D$, it remains, by the results in the preceding section for the spectra of the $D_\theta$, to determine a set of representatives for the set of all $\ZZ$-orbits contained in $\Sigma_{\eps'}$. More precisely, in view of Theorem~\ref{theo:decomp}, we carry out the following steps: \begin{enumerate} \item Describe all homomorphisms $\eps:\Gamma\to\ZZ_2$. \item Find a set of representatives $\cR_{\eps'}$ for all $\RR$-orbits $\omega$ intersecting $\Sigma_{\eps'}$. \item Compute the number of $\ZZ$-orbits contained in $\omega\cap\Sigma_{\eps'}$. \item Determine the spectrum of $D_\xi$ for some $\xi\in\omega$. \end{enumerate} This requires a detailed knowledge of the orbit picture of the coadjoint representation. In the following examples, the eigenvalues of the sub-Dirac operator including their multiplicities will be determined completely. \subsection{Three-dimensional Heisenberg manifolds}\label{S4.2} As we will see next, the results of this section comprise Theorem~3.1 of~{\rm \cite{AB}} concerning the spectrum of the Dirac operator on three-dimensional Heisenberg manifolds. Let $G=\RR^2\rtimes_A\RR$ be the Heisenberg group as discussed in Example~\ref{ex:3_dim_Heis} and $\Gamma=\ZZ^2\rtimes_A\ZZ$. Then $\fg=\Span\{e_1,e_2,b\}$ with $[b,e_2]=re_1$. For positive real numbers $d$ and~$T$ we consider the orientation and the Riemannian metric $g$ on $\cH:=TG$ such that $s_1=\frac{1}{T}e_1$, $s_2=-de_2$ and $s_3=\frac{d}{r}b$ becomes a positively oriented orthonormal frame. The constants are chosen in accordance with~\cite{AB}, where the collapse of Heisenberg manifolds $M(r,d,T)$ for $T\to 0$ is studied. Let $\nabla$ be the Levi-Civita connection of~$g$. In particular, $\nabla$ satisfies~(\ref{Ediv}) and $-\Gamma_{12}^3=\Gamma_{23}^1=\Gamma_{31}^2=\frac{d^2T}{2}$. Let $\eps':\ZZ^2\to\ZZ_2$ be a homomorphism. We abbreviate $\eps'(e_\mu)$ to $\eps_\mu$. Then (\ref{EsumA}) is satisfied if and only if $\eps_1r$ is even. It is easy to see that the disjoint union $\cR_{\eps'}$ of \[\cR_{\eps'}^{(1)} = \{\,\xi\in\Sigma_{\eps'}:\xi_1=0\,\}\quad\mbox{and}\quad \cR_{\eps'}^{(2)} = \{\,\xi\in\Sigma_{\eps'}:\xi_1\neq 0\mbox{ and }\xi_2=\eps_2\,\}\] is a set of representatives for the set of all $\RR$-orbits intersecting $\Sigma_{\eps'}$. The set $\cR_{\eps'}^{(1)}$ consists of all fixed points in $\Sigma_{\eps'}$. If $\omega$ is the $\RR$-orbit represented by $(\xi_1,\eps_2)\in\cR_{\eps'}^{(2)}$, then $\omega\cap\Sigma_{\eps'}$ contains $|\xi_1r|/2$ distinct $\ZZ$-orbits. We compute $|\,b\,|=\frac{r^2}{d^2}$, $\la\xi,s_1\ra=\frac{1}{T}\,\xi_1$, $\la\xi,s_2\ra=-d\xi_2$, $\la B^\top\xi,s_1\ra=0$ and $\la B^\top\xi,s_2\ra=-dr\,\xi_1$. Inserting this into~(\ref{eq:mu_k_for_3dim_H}) gives \begin{equation}\label{1*} \mu_k^\pm(0,\xi_2)=-\frac{d^2T}{4}\pm\pi\left(\,(2k+\dot\eps(1))^2\frac{d^2}{r^2} +d^2\xi_2^2\,\right)^{1/2}. \end{equation} Similarly, for $\xi\in\Sigma_{\eps'}$ with $\xi_1\neq 0$, Equations~(\ref{eq:lambda_0_for_3dim_H}) and~(\ref{eq:lambda_k_for_3dim_H}) yield \begin{equation}\label{2*} \lambda_0(\xi_1,\eps_2)=-\frac{d^2T}{4}-\frac{\pi}{T}\,|\xi_1| \end{equation} and \begin{equation}\label{3*} \lambda_k^{\pm}(\xi_1,\eps_2)=-\frac {d^2T}{4}\pm\left(\,2\pi d^2k|\xi_1| +\frac{\pi^2}{T^2}\,\xi_1^2\right)^{1/2}\,. \end{equation} We will use the following notation for the description of the spectrum of the Dirac operator $D$ on $\GG$. We define the spectral multiplicity function $$m(D):\RR\to\NN\cup\{|\NN|\},\quad m(D)(\lambda)=\dim\ker(D-\lambda I).$$ Moreover, $\delta=\delta(\lambda)$ denotes the function that takes the value $1$ in $\lambda$ and that is zero on $\RR\setminus\{\lambda\}$. Suppose we are given a spin structure corresponding to a homomorphism $\eps:\ZZ^2\rtimes_A\ZZ\rightarrow \ZZ_2$ with $\eps_1=0$. Summation of (\ref{1*}), (\ref{2*}) and (\ref{3*}) over $\xi_\nu\in 2\ZZ+\eps_\nu$ gives $$m(D)=m_1^+(D)+m_1^-(D)+m_2^+(D)+m_2^0(D)+m_2^-(D),$$ where \begin{eqnarray*} m_1^\pm(D) &=& \sum_{k\in\Bbb Z}\sum_{l\in\Bbb Z}\,\delta\left(\,-\frac{d^2T}{4}\pm \frac{\pi d}{r}\left(\,(2k+\dot\eps(1))^2+r^2(2l+\eps_2)^2\,\right)^{1/2}\,\right)\,,\\ m_2^0(D) &=& \sum_{l=1}^\infty\,2rl\,\delta\left(\,-\frac{d^2T}{4}-\frac{2\pi l}{T} \,\right)\,,\\ m_2^\pm(D) &=& \sum_{l=1}^\infty\,2rl\,\sum_{k=1}^\infty\delta\left(\,-\frac{d^2T}{4} \pm\left(\,4\pi d^2kl+\frac{4\pi^2l^2}{T^2}\,\right)^{1/2}\,\right)\;. \end{eqnarray*} Now assume $\eps_1=1$, which is only possible if $r$ is even. Then $\cR_{\eps'}^{(1)}=\emptyset$ and we obtain $$m(D)=m_2^+(D)+m_2^0(D)+m_2^-(D),$$ where now \begin{eqnarray*} m_2^0(D) &=& \sum_{l=0}^\infty\,(2l+1)r\,\delta\left(\,-\frac{d^2T}{4}-\frac{\pi(2l+1)}{T} \,\right)\,,\\ m_2^\pm(D) &=& \sum_{l=0}^\infty\,(2l+1)r\,\sum_{k=1}^\infty\delta\left(\,-\frac{d^2T}{4} \pm\left(\,2\pi d^2k(2l+1)+\frac{\pi^2(2l+1)^2}{T^2}\,\right)^{1/2}\,\right)\;. \end{eqnarray*} Now let us turn to the sub-Riemannian case and suppose $\cH=\Span\{s_2,s_3\}$, where again $s_2$ and $s_3$ are orthonormal. Let $\nabla$ be defined by (\ref{EK}) for a leftinvariant complement $\cV:=\RR\cdot u$, $u\in\fg$ of $\cH$. Since we wish to get a symmetric sub-Dirac operator, the only possible choice is $\cV:=\RR\cdot s_1$. Indeed, otherwise $[\Gamma(\cH),u]\not\subset\Gamma(\cH)]$, thus $D$ is not symmetric by Lemma~\ref{lm:sym_of_D}. We proceed as above, now using Equations~(\ref{eq:mu_k_for_2dim_H}) and~(\ref{eq:lambda_for_2dim_H}). For a spin structure that corresponds to a homomorphism $\eps:\ZZ^2\rtimes_A\ZZ\rightarrow \ZZ_2$ with $\eps_1=0$ we obtain $$m(D)=|\NN|\cdot \delta(0) +m_1^+(D)+m_1^-(D)+m_2^+(D)+m_2^-(D),$$ where \begin{eqnarray*} m_1^\pm(D) &=& \sum_{k\in\Bbb Z}\sum_{l\in\Bbb Z}\,\delta\left(\,\pm \frac{\pi d}{r}\left(\,(2k+\dot\eps(1))^2+r^2(2l+\eps_2)^2\,\right)^{1/2}\,\right)\,,\\ m_2^\pm(D) &=& \sum_{l=1}^\infty\,2rl\,\sum_{k=1}^\infty\delta\left(\, \pm\left(\,4\pi d^2kl\,\right)^{1/2}\,\right)\;. \end{eqnarray*} If $\eps_1=1$, then \begin{eqnarray*} \lefteqn {m(D)=|\NN|\cdot \delta(0)\ + }\\&&\sum_{l=0}^\infty\,(2l+1)r\,\sum_{k=1}^\infty\left(\delta\big(\, (\,2\pi d^2k(2l+1))^{1/2}\,\big)+\delta\big( -(\,2\pi d^2k(2l+1))^{1/2}\big)\,\right) . \end{eqnarray*} \subsection{A five-dimensional two-step nilpotent example} \label{S4.3} We start by considering 2-step nilpotent Lie groups which are isomorphic to a standard model $G=\RR^{2p}\rtimes_A \RR$ as described in Lemma~\ref{lm:standard_2_step_nilp} and therefore generalise the example from the preceding subsection. We will describe the orbits of $\RR$ and $\ZZ$ acting on $\RR^{2p}$ by $A^\top$. Then we will specialise to $\dim G=5$ for the computation of the spectrum of the sub-Dirac operator on $\GG$. \begin{lm}\label{lm:standard_2_step_nilp} Let $G$ be a simply connected Lie group satisfying $[G,G]=Z(G)$ and admitting a connected abelian normal subgroup $N$ of codimension~$1$. Let $\Gamma$ be a uniform discrete subgroup of $G$ such that $\Gamma\cap N$ is uniform in~$N$. Then there exist $p\ge 1$, a one-parameter subgroup of $\GL(2p,\RR)$ of the form \[A(t)=\left( \begin{array}{cc} I & tR \\ 0 & I \end{array} \right)\] with $R=\diag(r_1,\ldots,r_p)$ and positive integers $r_\nu$ such that $r_{\nu+1}\,|\,r_\nu$ for $\nu=1,\ldots,p-1$, and an isomorphism $\Phi$ of $G$ onto $\RR^{2p}\rtimes_A\RR$ such that $\Phi(\Gamma)=\ZZ^{2p}\times_A\ZZ$. \end{lm} \begin{proof} Put $p=\dim Z(G)=\frac{1}{2}\dim N$. Since $\Gamma\cap Z(G)$ is uniform in $Z(G)$, we find generators $v_1,\ldots,v_{2p}$ of $\Gamma\cap N$ such that $\Gamma\cap Z(G)=\ZZ v_1+\ldots+\ZZ v_p$. As in the proof of Lemma~\ref{lm:model}, we consider the linear isomorphism $M:\RR^{2p}\to N$ given by $M(e_\nu)=v_\nu$, choose $b\in\fg$ such that $\exp(b)\in\Gamma$ and $\exp(b)N$ generates $\Gamma N/N$, and define $A_0(t)\in\GL(2p,\RR)$ such that $\Phi_0(x,t)=M(x)\exp(tb)$ becomes an isomorphism of~$G$ onto~$\RR^{2p}\rtimes_{A_0}\RR$. Since $\Phi_0(Z(G))=\RR^p\times\{0\}\times\{0\}$, we have \[A_0(t)=\left(\begin{array}{cc} I & tR_0 \\ 0 & I\end{array}\right)\] with $R_0\in\GL(p,\ZZ)$. Recall that $R_0$ can be brought into Smith normal form, i.e., there exist $Q_1,Q_2\in\GL(p,\ZZ)$ such that $R:=Q_1R_0Q_2^{-1}= \diag(r_1,\ldots,r_p)$ is diagonal with positive integers $r_\nu$ such that $r_{\nu+1}\,|\,r_\nu$. Clearly $\Psi(x',x'',t):=(Q_1x',Q_2x'',t)$ gives an isomorphism of~$\RR^{2p}\rtimes_{A_0}\RR$ onto~$\RR^{2p}\rtimes_{A}\RR$ with $A(t)(x',x'')=(x'+tRx'',x'')$. Finally, it follows that the assertion of the lemma holds for $\Phi:=\Psi\Phi_0$. \qed \end{proof} Let $G=\RR^{2p}\rtimes_A\RR$ be as in Lemma~\ref{lm:standard_2_step_nilp} with uniform discrete subgroup $\Gamma=\ZZ^{2p}\rtimes_A\ZZ$. In particular, $A(t)e_\nu=e_\nu$ and $A(t)e_{p+\nu}=e_{p+\nu}+r_\nu t\,e_\nu$ for all $1\le\nu\le p$. Let $\dot{\eps}:\ZZ\to\ZZ_2$ and $\eps':\ZZ^{2p}\to\ZZ_2$ be homomorphisms. As before, we put $\eps_\nu=\eps'(e_\nu)$ for $\nu=1,\dots,2p$. By Lemma~\ref{lm:eps} it follows that $\eps(k,l):=\eps'(k)+\dot{\eps}(l)$ is a homomorphism of~$\Gamma$ if and only if $r_\nu\,\eps_\nu\in 2\ZZ$ for $1\le\nu\le p$. Note that the latter condition implies $\eps_\nu=0$ whenever $r_\nu$ is odd. Next we will describe the coadjoint orbits. First of all, \begin{equation}\label{eq:coadj_rep} \la A(t)^\top\xi,e_\nu\ra=\xi_\nu\quad\text{and}\quad \la A(t)^\top\xi,e_{p+\nu}\ra=\xi_{p+\nu}+r_\nu\xi_\nu\,t \end{equation} for $1\le\nu\le p$. To formulate the subsequent result, a little more notation is needed. If $\xi\in\ZZ^{2p}$, then $\bar{\xi}\in\ZZ^p$ denotes the projection of $\xi$ onto the first $p$ variables. For $\eta\in\ZZ^p$, the subset $\{\,\xi\in\ZZ^{2p}:\bar{\xi}=\eta\,\}$ is $\ZZ$-invariant. In particular, $\{\,\xi:\bar{\xi}=0\,\}$ is the set of all points remaining fixed under the coadjoint action. Put $J_\eta=\{\,\nu:\eta_\nu\neq 0\,\}$. For $\eta\neq 0$, let $d_\eta>0$ be the greatest common divisor of the integers $|r_1\,\eta_1|,\ldots,|r_p\eta_p|$. We choose $j_\eta=\min J_\eta$ and set $q_\eta=|r_{j_\eta}\eta_{j_\eta}|\,/\,d_\eta$. Let $\bar{\Sigma}_{\eps'}$ be the image of $\Sigma_{\eps'}$ under projection. If $\eta\in\bar{\Sigma}_{\eps'}\setminus\{0\}$, then $d_\eta$ is even because $\eta_{\nu}$ is even whenever $r_\nu$ is odd. Furthermore, we define $\cR_{\eps'\!,0}=\{\,\xi\in\Sigma_{\eps'}:\bar{\xi}=0\,\}$ and $\cR_{\eps'\!,\eta}=\{\,\xi\in\Sigma_{\eps'}:\bar{\xi}=\eta$ and $0\le \xi_{p+j_\eta}\le 2q_\eta-1\,\}$ for $\eta\in\bar{\Sigma}_{\eps'}$ non-zero. Note that $\cR_{\eps'\!,0}$ is empty if $\eps_\nu=1$ for some $1\le\nu\le p$. \begin{lm}\label{lm:set_of_rep} In this situation, the following holds true: \begin{enumerate}[label=(\roman*)] \item The disjoint union $\cR_{\eps'}:=\bigcup_{\eta\in\bar{\Sigma}_{\eps'}} \cR_{\eps'\!,\eta}$ is a set of representatives for the set of all $\RR$-orbits intersecting $\Sigma_{\eps'}$. \item Let $\omega$ be an $\RR$-orbit which intersects $\Sigma_{\eps'}$. Then $\eta:=\bar\xi$ does not depend on the choice of $\xi\in\omega\cap\Sigma_{\eps'}$. If $\omega$ is not a fixed point, then $\omega\cap\Sigma_{\eps'}$ consists of~$d_{\eta}/2$ distinct $\ZZ$-orbits. \end{enumerate} \end{lm} \begin{proof} Let $\xi\in\Sigma_{\eps'}$ such that $\bar{\xi}\neq 0$. By (\ref{eq:coadj_rep}) we know that $A(t)^\top\xi\in\Sigma_{\eps'}$ if and only if $r_\nu\xi_{\nu}t\in 2\ZZ$ for all $\nu\in J_{\bar{\xi}}$. This proves \begin{equation}\label{eq*}\{\,t\in\RR:A(t)^\top\xi\in\Sigma_{\eps'}\,\} \;=\;\bigcap_{\nu\in J_{\bar{\xi}}}\frac{2}{|r_\nu\xi_{\nu}|}\,\ZZ\; =\;\frac{2}{d_{\bar{\xi}}}\,\ZZ \end{equation} To prove \textit{(i)}, let $\omega$ be an $\RR$-orbit and $\xi\in\omega\cap\Sigma_{\eps'}$. Clearly $\eta:=\bar{\xi}$ does not depend on the choice of $\xi$. We can assume $\bar{\xi}\neq 0$. Define $d_\eta$ and $j=j_\eta$ as above. Since $\la A(t)^\top\xi,e_{p+j}\ra=e_{p+j}+r_j\xi_{j}\,t$, it follows from (\ref{eq*}) that there exists $t\in\frac{2}{d_\eta}\ZZ$ such that $A(t)^\top\xi\in\Sigma_{\eps'}$ and $0\le\la A(t)^\top\xi,e_{p+j}\ra\le 2q_\eta-1$. This proves $A(t)^\top\xi\in\omega\cap\cR_{\eps'\!,\eta}$ because $\overline{A(t)^\top\xi}=\bar{\xi}$. We claim that $\omega\cap\cR_{\eps'\!,\eta}$ consists of a single point: If $\xi,\xi^\ast\in\omega\cap\cR_{\eps'\!,\eta}$, then, again by~(\ref{eq*}), there exists a $t\in\frac{2}{d_\eta}\,\ZZ$ such that $\xi^\ast=A(t)^\top\xi$. In particular $\xi^\ast_{p+j}=\xi_{p+j}+r_j\xi_{j}t$. Since $0\le\xi_{p+j},\xi^\ast_{p+j}\le 2q_\eta-1$, it follows $t=0$ and hence $\xi^\ast=\xi$. This proves $\cR_{\eps'}$ to be a set of representatives. Let $\xi\in\omega\cap\Sigma_{\eps'}$ be an arbitary non-fixed point. Then $f:\RR\to\omega$, $f(t)=A(t)^\top\xi$, is bijective and $\RR$-equivariant. By~(\ref{eq*}) it holds $f^{-1}(\omega\cap\Sigma_{\eps'})=\frac{2}{d_{\eta}}\,\ZZ$. Since $d_{\eta}/2$ is an integer, it follows \[\#\;\ZZ\setminus\omega\cap\Sigma_{\eps'}=\#\;\ZZ\setminus\frac{2}{d_{\eta}} \ZZ=\frac{d_{\eta}}{2}\;.\] More precisely, the points $\{A(\frac{2k}{d_{\eta}})^\top\xi: 0\le k<\frac{d_{\eta}}{2}\}$ are representatives for the set of all $\ZZ$-orbits in $\omega\cap\Sigma_{\eps'}$. \qed \end{proof} We point out that the choice of the set $\cR_{\eps'}$ is in no way canonical. For example, any choice of indices $j_\eta\in J_\eta$ leads to a set of representatives.\\\\ Now let us restrict ourselves to $p=2$. Then canonical basis $e_1,\ldots,e_4,b$ of the Lie algebra $\fg\cong\RR^{2p}\rtimes_B\RR$ of $G$ satisfies the relations $[b,e_3]=r_1e_1$ and $[b,e_4]=r_2e_2$. Put $s_1=e_3$, $s_2=e_4$ and $s_3=b$. As before, the corresponding left-invariant vector fields are denoted by the same symbol. The left-invariant distribution $\cH:=\Span\{s_1,s_2,s_3\}$ is given the orientation and Riemannian metric $g$ such that $s_1,s_2,s_3$ is a positively oriented, orthonormal frame. In particular, $|\,b\,|=1$. Note that $\cH$ is bracket-generating. \begin{re} In general, when $\cH$ is a left-invariant 3-dimensional distribution on a Lie group $G$, the affine space of all left-invariant metric connections in $\cH$ satisfying~(\ref{Ediv}) has dimension~6. However, in the present example, the left-invariant connections which are defined by a left-invariant projection~$\proj$ onto~$\cH$ and the Koszul formula~(\ref{EK}) and which satisfy~(\ref{Ediv}) form a $3$-dimensional space. \end{re} Let $\nabla$ be a left-invariant metric connection in $\cH$ satisfying~(\ref{Ediv}). For example, we could take the connection given by projection onto $\cH$ along~$\cV:=\Span\{e_1,e_2\}$, which, according to~(\ref{EK}), satisfies $\Gamma_{ij}^k=0$ for all $i,j,k$ because $[\fg,\fg]\subset\cV$. Let $\eps:\Gamma\to\ZZ_2$ be a homomorphism giving a spin structure of $\dot{\cH}$. By Lemma~\ref{lm:sym_of_D} the sub-Dirac operator $D$ defined by $(\cH,g,\nabla,\eps)$ is symmetric. We compute its spectrum. To this end, we note that the coadjoint representation is given by \[B^\top\xi=\left(\begin{array}{c} 0 \\ 0 \\ r_1\xi_1 \\ r_2\xi_2 \end{array}\right)\qquad\mbox{and}\qquad A(t)^\top\xi= \left(\begin{array}{c} \xi_1 \\ \xi_2 \\ \xi_3+r_1\xi_1t \\ \xi_4+r_2\xi_2t \end{array}\right)\;.\] In particular, we get $\la\xi,s_\nu\ra=\la\xi,e_{2+\nu}\ra=\xi_{2+\nu}$ and $\la B^\top\xi,s_\nu\ra=r_\nu\xi_{\nu}$ for $\nu=1,2$. Put $\alpha=-\frac{1}{2}\,(\,\Gamma_{12}^3+\Gamma_{23}^1+\Gamma_{31}^2\,)$. By (\ref{eq:mu_k_for_3dim_H}), (\ref{eq:lambda_0_for_3dim_H}) and (\ref{eq:lambda_k_for_3dim_H}), the eigenvalues of $D_\xi$ are of the form \[\mu_k^\pm(\xi)=\alpha\pm\pi\left(\,(2k+\dot{\eps}(1))^2+\xi_3^2+\xi_4^2\, \right)^{1/2}\] for fixed points, and \[\lambda_0(\xi)=\alpha-\pi\,\frac{r_1\xi_1\xi_4-r_2\xi_2\xi_3} {(\,r_1^2\xi_1^2+r_2^2\xi_2^2\,)^{1/2}}\] or \[\lambda_k^\pm(\xi)=\alpha\pm\left(\,2\pi k(r_1^2\xi_1^2+r_2^2\xi_2^2)^{1/2}+\pi^2\, \frac{(\,r_1\xi_1\xi_4-r_2\xi_2\xi_3\,)^2}{r_1^2\xi_1^2+r_2^2\xi_2^2}\,\right)^{1/2}\] else. We want to decompose the set $\cR_{\eps'}$ of representatives into a disjoint union of sets that we can describe explicitly. To this end, consider $\eta=\bar\xi \in \bar{\Sigma}_{\eps'}$ and assume $\eta\not=0$. If $\eta_1\not=0$ and $\eta_2=0$, then $j_\eta=1$, $d_\eta=|r_1\eta_1|$ and $q_\eta=1$. Similarly, if $\eta_1=0$ and $\eta_2\not=0$, then $j_\eta=2$, $d_\eta=|r_2\eta_2|$ and $q_\eta=1$. For $\eta_1\eta_2\not=0$ we get $j_\eta=1$, and obtain $d_\eta=\gcd(|\,r_1\eta_1|,|r_2\eta_2|\,)$ and $q_\eta=|r_1\eta_1|\,/\,d_\eta$. This leads to a decomposition of~$\cR_{\eps'}$ into the following subsets: \begin{eqnarray*} \cR_{\eps'}^{(1)} &=& \cR_{\eps',0}\,,\\ \cR_{\eps'}^{(2)} &=& \{\,\xi\in\Sigma_{\eps'}:\xi_1\neq 0,\,\xi_2=0,\, \xi_3=\eps_3\,\}\,,\\ \cR_{\eps'}^{(3)} &=& \{\,\xi\in\Sigma_{\eps'}:\xi_1=0,\, \xi_2\neq 0,\, \xi_4=\eps_4\,\}\,,\\ \cR_{\eps'}^{(4)} &=& \{\,\xi\in\Sigma_{\eps'}:\xi_1\neq 0,\,\,\xi_2\neq 0,\, 0\le \xi_3\le 2q_{(\xi_1,\xi_2)}-1\,\}\,. \end{eqnarray*} We have $\cR_{\eps'}^{(1)}=\emptyset$ if $\eps_1=1$ or $\eps_2=1$, $\cR_{\eps'}^{(2)}=\emptyset$ if $\eps_2=1$, and $\cR_{\eps'}^{(3)}=\emptyset$ if $\eps_1=1$. Recall that for $\nu=1,2$ the case $\eps_\nu=1$ can occur only if $r_\nu$ is even. The spectrum of $D$ depends on the spin structure given by $\eps$. It holds $m(D)=\sum_{i=1}^4m_i$ where $m_i=\sum_{\xi\in\cR_{\eps'}^{(i)}}m(D_\xi)$. Note that $m_i=0$ if $\cR_{\eps'}^{(i)}=\emptyset$. Otherwise, $m_i$ is given as follows, where the sums are meant to be taken over $\xi_\nu\in 2\ZZ+\eps_\nu$ and $\xi_5\in 2\ZZ+\dot{\eps}(1)$. \begin{eqnarray*} m_1 &=& \sum_{\xi_3,\xi_4,\xi_5}\,\delta(\,\alpha+\pi(\,\xi_3^2+\xi_4^2 +\xi_5^2\,)^{1/2}\,)+\delta(\,\alpha-\pi(\,\xi_3^2+\xi_4^2+\xi_5^2\,)^{1/2}\,)\\ m_2 &=& \sum_{\xi_1\neq 0}\frac{|r_1\xi_1|}{2}\sum_{\xi_4} \bigg(\,\delta(\,\alpha-\pi\sgn(r_1\xi_1)\xi_4\,)\\ & & \hspace{0.5cm}+\sum_{k=1}^\infty\,\left(\,\delta(\,\alpha+(2\pi k|r_1\xi_1| +\pi^2\xi_4^2)^{1/2}\,)+\delta(\,\alpha-(2\pi k|r_1\xi_1|+\pi^2\xi_4^2)^{1/2}\,) \,\right)\,\bigg)\\ m_3 &=&\sum_{\xi_2\neq 0}\frac{|r_2\xi_2|}{2}\sum_{\xi_3}\bigg(\,\delta(\, \alpha+\pi\sgn(r_2\xi_2)\xi_3\,)\\ & & \hspace{0.5cm}+\sum_{k=1}^\infty\,\left(\,\delta(\,\alpha+(2\pi k|r_2\xi_2| +\pi^2\xi_3^2)^{1/2}\,)+\delta(\,\alpha-(2\pi k|r_2\xi_2| +\pi^2\xi_3^2)^{1/2}\,)\,\right)\,\bigg)\\ m_4 &=& \sum_{\xi_1\neq 0}\sum_{\xi_2\neq 0} \frac{\gcd(\,|r_1\xi_1|,|r_2\xi_2|\,)}{2}\\ & & \hspace{1cm}\sum_{0\le\xi_3\le 2q_{(\xi_1,\xi_2)}-1}\sum_{\xi_4} \left(\,\delta(\,\lambda_0(\xi)\,)+\sum_{k=1}^\infty\, \left(\,\delta(\,\lambda^{+}_k(\xi)\,)+\delta(\,\lambda^{-}_k(\xi)\,)\, \right)\,\right) \end{eqnarray*} In particular, if $\eps_\nu=0$ for $\nu=1$ or $2$, then the numbers $\{\alpha+(2k+\eps_{2+\nu})\pi :k\in\ZZ\}$ are eigenvalues of $D$ and each of them has infinite multiplicity. In this example, the spectrum of $D$ is a non-discrete subset of $\RR$, no matter which homomorphism $\eps:\Gamma\to\ZZ_2$ defining the underlying spin structure is chosen. Indeed, $\alpha^\ast:=\alpha+\pi\sgn(r_2)\eps_3$ is an accumulation point of~$\sigma(D)$. To see this, we consider the sequence $\xi_n\in\cR_{\eps'}^{(4)}$ given by $\xi_{n1}=2+\eps_1$, $\xi_{n2}=2n+\eps_2$, $\xi_{n3}=\eps_3$ and $\xi_{n4}=\sgn(r_1r_2)(2+\eps_4)$. Then $\lambda_0(\xi_n)\neq\alpha^\ast$ and $\lambda_0(\xi_n)\to\alpha^\ast$ for $n\to+\infty$. \subsection{A three-step nilpotent example} Let $r_1,r_2\in \ZZ\setminus \{0\}$ be such that $r_1r_2$ is even. Define a Lie algebra structure on $\fg:=\Span\{e_1,e_2,e_3,b\}$ such that $\fn:=\Span\{e_1,e_2,e_3\}$ is an abelian ideal and $[b,X]=B(X)$ for $X\in\fn$, where $B:\fn\rightarrow\fn$ is given by $$ B=\left( \begin{array}{ccc} 0&r_1&0\\ 0&0&r_2\\ 0&0&0 \end{array}\right)$$ with respect to the basis $e_1,e_2,e_3$ of $\fn$. Let $G$ be the simply-connected Lie group with Lie algebra $\fg$. Then $G=\RR^3\rtimes_A\RR$ with $$A(t)=\exp tB =\left( \begin{array}{ccc} 1&tr_1&t^2r_1r_2/2\\ 0&1&tr_2\\ 0&0&1 \end{array}\right).$$ Since $A(1)$ is in $SL(2,\ZZ)$, the subset $\Gamma:=\ZZ^3\rtimes_A\ZZ$ a uniform discrete subgroup of~$G$. Let $(\cH,g)$ be the oriented sub-Riemannian structure having $s_1:=e_3$, $s_2:=b$ as a positively oriented orthonormal frame. Then $\cH$ is bracket generating. The spin structures of $\dot{\cH}$ correspond to homomorphisms $\eps:\Gamma \rightarrow\ZZ_2$. As above we write $\eps(k,l)=\eps'(k)\cdot\dot{\eps}(l)$, where $\dot{\eps}:\ZZ\rightarrow \ZZ_2$ is an arbitrary homomorphism and $\eps':\ZZ^3\rightarrow \ZZ_2$ is a homomorphism satisfying~(\ref{EsumA}), which, in this example, means that $r_1\eps_1$ and $r_1r_2\eps_1/2+r_2\eps_2$ are both even. More precisely, this shows: If $r_1$ and $r_2$ are both even, then $\eps_1$ and $\eps_2$ are arbitrary. If $r_1$ is odd and $r_2$ is even, then $\eps_1=0$ and $\eps_2$ is arbitrary. Now suppose that $r_2$ is odd. If, in addition, $r_1$ is odd, then $\eps_1=\eps_2=0$. If $r_1$ is even but not divisble by $4$, then either $\eps_1=\eps_2=0$ or $\eps_1=\eps_2=1$. Finally, if $r_1$ is divisible by $4$, then $\eps_2=0$. Clearly ${\cal V}:=\Span\{e_1,e_2\}$ is a complement of $\cH$ in the tangent bundle $TG$. Using the projection onto $\cH$ along ${\cal V}$, we define a left-invariant connection $\nabla$ in $\cH$ by the Koszul formula (\ref{Ediv}). Since $\proj [s_1,s_2]=0$, all Christoffel symbols $\Gamma_{ij}^k$ vanish. In particular, the sub-Dirac operator is symmetric and equals $$D=s_1\cdot \partial_{s_1}+s_2\cdot \partial_{s_2}\,,$$ where we use the simple $\Cl(\cH_e)$-module structure on $\CC^2$ defined by $$s_1\mapsto\left(\begin{array}{cc} 0&-1\\1&0 \end{array}\right),\quad s_2\mapsto \left(\begin{array}{cc} i&0\\0&-i \end{array}\right).$$ On the other hand, we have \begin{equation}\label{EAt} A^\top(t)\xi = \left(\begin{array}{c} \xi_1\\ \xi_2+tr_1\xi_1\\ \xi_3+tr_2\xi_2+t^2 r_1r_2\xi_1/2\end{array}\right). \end{equation} In particular, the sets \begin{eqnarray*} R^{(1)}&:=& \{\xi\in\RR^3\mid \xi_1=\xi_2=0\},\\ R^{(2)}&:=& \{\xi\in\RR^3\mid \xi_1=0,\ \xi_2\not=0\},\\ R^{(3)}&:=& \{\xi\in\RR^3\mid \xi_1\not=0\} \end{eqnarray*} are invariant under $A^\top(t)$ for all $t\in\RR$. Let us first consider $D_\theta$ for the orbit $\theta=\{\xi\}$ of an element $\xi\in R^{(1)}$. Then, according to (\ref{eq:mu_k_for_2dim_H}), the spectrum of~$D_\theta$ consists of the eigenvalues $$\mu_k^\pm(\xi)=\pm \pi \left( (2k+\dot{\eps}(1))^2+\xi_3^2\right)^{1/2},\ \ k\in\ZZ.$$ Now consider $\xi\in R^{(2)}$. Then $D_\xi$ has the form \begin{equation}\label{Elast} \left( \begin{array}{cc} i\partial_t & \bar\omega\\ \omega & -i\partial_t \end{array}\right) \end{equation} with $\omega(t)=a\omega_1t+\omega_0$, where $$ a=\pi |r_2\xi_2|,\quad \omega_1=\sgn(r_2\xi_2)\cdot i,\quad \omega_0= \pi i \xi_3.$$ According to (\ref{eq:lambda_for_2dim_H}) the spectrum of $D_\xi$ consists of the eigenvalues $\lambda_0=0$ and $$\lambda_k^\pm = \pm (2\pi |r_2\xi_2|k)^{1/2},\quad k\in\NN\setminus\{0\}.$$ Finally, take $\xi\in R^{(3)}$. Then $ D_\xi$ is of the form (\ref{Elast}) where $\omega(t)=i\pi (\xi_1r_1r_2t^2/2 +\xi_2r_2t +\xi_3)$. Hence $$D_\xi^2=\left( \begin{array}{cc} -\partial_t^2-\omega(t)^2 & -i\omega'(t)\\ -i\omega'(t) & -\partial_t^2 -\omega(t)^2 \end{array}\right).$$ Obviously, $D_\xi^2$ is time-independent diagonalisable. More exactly, $D_\xi^2$ is conjugate to $$\left( \begin{array}{cc} -\partial_t^2-\omega(t)^2 -i \omega'(t)&0\\ 0 & -\partial_t^2 -\omega(t)^2+i\omega'(t) \end{array}\right).$$ The operators $ -\partial_t^2-\omega(t)^2 \mp i\omega'(t)$ are of the form $$ P_{a,b,c}^\pm:= \-\partial_t^2+(at^2+bt+c)^2\pm (2at+b)$$ for $$a=\pi\xi_1r_1r_2\not=0,\quad b=\pi\xi_2r_2,\quad c=\pi\xi_3.$$ We consider the bijection $$L^2(\RR)\longrightarrow L^2(\RR),\quad \ph\longmapsto \tilde \ph, \ \tilde\ph(t) =\frac1{x^2}\ph(xt+y),$$ where $x=a^{1/3}$, $y=ba^{-2/3}/2$. We define $P_c^\pm:=P_{1,0,c}^\pm$. {\bf Claim.} The equation $P_{a,b,c}^\pm\tilde\ph=\tilde\lambda\tilde\ph$ is equivalent to $P_{c_1}^\pm\ph=\lambda\ph$, where $$c_1=-b^2a^{-4/3}/2 +ca^{-1/3},\quad \tilde \lambda = a^{2/3}\lambda.$$ Indeed, assume that $P_{c_1}^\pm\ph=\lambda\ph$. Then $\ph''(t)=\left( (t^2+c_1)^2\pm2t -\lambda\right)\ph(t)$ holds. Hence \begin{eqnarray*} \lefteqn{(P_{a,b,c}^\pm\tilde \ph)(t)\ =\ -(\partial_t^2\tilde\ph)(t) + \left( (at^2+bt+c)^2\pm(2at+b)\right)\tilde\ph(t)}\\ &&=\ \left( -x^2\left(((xt+y)^2+c_1)^2\pm2(xt+y)-\lambda\right) + (at^2+bt+c)^2\pm(2at+b)\right)\tilde\ph(t)\\ && =\ x^2\lambda\tilde\ph(t)\ =\ a^{2/3}\lambda\tilde\ph. \end{eqnarray*} The converse can be proven similarly using $\ph(t)=x^2\tilde\ph(t/x-y/x)$. It is well known that the Schr\"odinger operator $P_c^\pm$ having a polynomial potential of degree 4 has the following properties \cite{EGS,T}. The spectrum of $P_c^\pm$ is discrete. All eigenvalues are real and simple. They can be arranged into an increasing sequence $\lambda_0<\lambda_1<\dots \to \infty$ and satisfy $$\lambda_k\sim \left(\frac{\sqrt\pi\Gamma(7/4)\cdot k}{\Gamma(5/4)}\right)^{4/3}.$$ Obviously, $P_c^+$ and $P_c^-$ have the same eigenvalues. We will denote these eigenvalues by $\lambda_k(c)$, $k\in\NN$. Since $\dim \cH$ is even the spectrum of $D_\xi$ is symmetric. We conclude that ${\rm spec}(D_\xi)$ consists of the eigenvalues $$\pm \left(a^{2/3}\lambda_k (-4b^2 a^{-4/3}+ca^{-1/3})\right)^{1/2},\quad k\in\NN,$$ where $a=\pi\xi_1r_1r_2/2$, $b=\pi\xi_2r_2$, $c=\pi \xi_3$. Next we determine a set of representatives of the $\RR$-orbits in $\RR^3$ that intersect $\Sigma_{\eps'}$ and the number of $\ZZ$-orbits that are contained in them. Obviously, $${\cal R}^{(1)}:= R^{(1)}\cap \Sigma_{\eps'}$$ is the set of fixed points in $\Sigma_{\eps'}$ and $${\cal R}^{(2)}:= \{\xi\in\Sigma_{\eps'}\mid \xi_1=0,\ \xi_2\not=0,\ \xi_3=\eps_3\}$$ is a set of representatives of the $\RR$-orbits in $R^{(2)}$ that interset $\Sigma_{\eps'}$. For $\xi\in{\cal R}^{(2)}$ the $\RR$-orbit through $\xi$ contains $|r_2\xi_2|/2\,$ $\ZZ$-orbits. Now we turn to orbits contained in $R^{(3)}$. For a given number $k\in\ZZ\setminus\{0\}$ let $p,q\in \ZZ$, $q>0$ be such that $$\frac{|r_2|}{|r_1k|}=\frac pq, \quad (p,q)=1$$ and put $q(k):=q$. Moreover, for $l,q\in\NN$, $q>0$ we define $$M(l,q):=\{(m_1,m_2)\mid m_1,m_2\in\NN\setminus\{0\},\ m_1+m_2=l,\ q|m_1m_2\}.$$ We will show: \begin{enumerate} \item The set $$ {\cal R}^{(3)}:=\{\xi\in\Sigma_{\eps'}\mid 0\le\xi_2<|r_1\xi_1|,\ M(\xi_2, q(\xi_1)) =\emptyset\}$$ is a set of representatives of $\RR$-orbits in $R^{(3)}$ that intersect $\Sigma_{\eps'}$. \item For $\xi \in{\cal R}^{(3)}$ the number of $\ZZ$-orbits contained in the $\RR$-orbit of $\xi$ equals $$m(\xi_1,\xi_2):=\#\{k\in\NN\mid \xi_2+2k<|r_1\xi_1|,\ q(\xi_1)|k(k+\xi_2)\}.$$ \end{enumerate} Take $\xi\in R^{(3)}\cap \Sigma_{\eps'}$ and denote by $\theta$ the $\RR$-orbit of $\xi$. Using (\ref{EAt}) we see that $A^\top(t)\xi$ is in $\Sigma_{\eps'}$ if and only if $tr_1\xi_1$ and $t^2r_1r_2\xi_1/2+tr_2\xi_2$ are in $2\ZZ$. The latter condition is equivalent to \begin{equation}\label{Etq} t=\frac{2k}{r_1\xi_1},\quad q(\xi_1)|k(k+\xi_2) \end{equation} for some $k\in\ZZ$. Obviously, we may choose $\hat\xi=(\xi_1,\hat\xi_2,\hat\xi_3)\in\theta$ such that $0\le\hat \xi_2<|r_1\xi_1|$. Now we want to choose $\hat \xi$ is such a way that $\hat\xi_2\ge 0$ is minimal, which ensures the uniqueness of the representative. By (\ref{Etq}), $\hat\xi_2$ is minimal if and only if there does not exist an integer $k$, $-[\hat\xi_2/2]\le k\le-1$, such that $q(\xi_1)|k(k+\hat\xi_2)$. The latter condition is equivalent to $q(\xi_1)|(-k)(k+\hat\xi_2)$. Hence $\hat\xi_2$ is minimal if and only if $\hat\xi_2$ does not decompose as a sum $\hat\xi_2=m_1+m_2$ with $m_1, m_2\in\NN\setminus\{0\}$ and $q(\xi_1)|m_1m_2$. This proves the first assertion. The second one follows from (\ref{Etq}). Now we can give an expression for $m(D)$. In the following sums are taken over $\xi_i\in\eps_1+2\ZZ$, $i=1,2,3$. Moreover, we willtake another index of summation, namely $\xi_4\in\dot{\eps}(1)+2\ZZ$. Furthermore, $\kappa\in\{1,-1\}$. Then \begin{eqnarray*} \lefteqn{ m(D)= \sum_{\xi_3,\xi_4}\sum_\kappa\delta \left(\kappa\pi(\xi_3^2+\xi_4^2)^{1/2}\right)}\\ && +\sum_{\xi_2>0} |r_2\xi_2|\left(\delta(0)+\sum_{k=1}^\infty\sum_\kappa \delta(\kappa(2\pi k |r_2\xi_2|)^{1/2})\right)\\ &&+\sum_{\xi_1\not=0,\xi_2,\xi_3} \hspace{-9pt} m(\xi_1,\xi_2)\cdot\sum_{k=0}^\infty \sum_\kappa\delta\left( \kappa(\pi\xi_1\frac{r_1r_2}2)^{1/3}\lambda_k \Big((\pi\xi_1\frac{r_1r_2}2)^{-1/3}\pi( \xi_3-\frac{8\xi_2^2r_2}{\xi_1r_1})\Big)^{1/2}\right). \end{eqnarray*}
\section{Introduction} \label{sec_introduction} Ground state fidelity is defined as the overlap between two ground states \cite{Zanardi} \begin{equation} F(\lambda,\delta) = \left|\langle \lambda-\delta/2|\lambda+\delta/2\rangle\right|, \label{F} \end{equation} where $|\lambda\rangle$ is the ground state of the Hamiltonian $\hat H(\lambda)$, $\lambda$ is the parameter of that Hamiltonian, and $\delta$ is its shift. Since fidelity quantifies similarity of the ground states, it is a useful probe of quantum criticality. Indeed, the quantum phase transition happens when the ground state of the system can be fundamentally changed by a small variation of some parameter of the system's Hamiltonian \cite{Sachdev,ContinentinoBook,SachdevToday}. Since the ground states belonging to different phases have little in common, fidelity is expected to exhibit a marked drop across the critical point \cite{Zanardi}. This intuitive remark was studied in a surprisingly-large variety of physical models, for example, in numerous spin models (Ising, XY, XYX, XXZ, Heisenberg, Kitaev, Lipkin-Glick-Meshkov, etc.) \cite{GuReview}. The scaling theory of quantum phase transitions was employed to predict the dependence of fidelity on the distance $|\lambda-\lambda_c|$ from the critical point, the system size $M$, and the critical exponent $\nu$ characterizing the power-law divergence of the correlation length $\xi\sim|\lambda-\lambda_c|^{-\nu}$ \cite{ABQ2010,Polkovnikov,BDfid1}. The key insights provided by these studies can be briefly summarized as follows. In the limit of $\delta\to0$ taken under the fixed system size $M$ fidelity can be expanded as \begin{equation} F(\lambda,\delta)=1-\chi(\lambda)\frac{\delta^2}{2} + O\B{\delta^4}. \label{fidsus} \end{equation} Then, one finds that around the critical point $\chi \sim M^{2/d\nu}$, while far away from it $\chi\sim M/|\lambda-\lambda_c|^{2-d\nu}$ \cite{ABQ2010,Polkovnikov}. $\chi$ is known as fidelity susceptibility and $d$ stands for the system's dimensionality. On the other hand, in the limit of $M\to\infty$ taken at the fixed field shift $\delta$ one finds that near the critical point $\ln F \sim -M|\delta|^{d\nu}$, while far away from it $\ln F \sim -M\delta^2/|\lambda-\lambda_c|^{2-d\nu}$ \cite{BDfid1}; see also Ref.~\cite{Zhou} for a similar approach to fidelity in the thermodynamically-large systems and Ref. \cite{BDfid2} for some modifications to these scaling laws. We will study below the Bose-Hubbard model at unit filling factor \cite{BHearly,Fisher89}. This model describes interacting bosons in a lattice. It can be experimentally realized in optical lattices filled with ultra cold atoms. This was proposed in Ref. \cite{JakschPRL1998} and accomplished in a cubic lattice a few years later \cite{Greiner2002}. Soon by the appropriate modifications of the lattice potential, a one-dimensional (1D) version of the model was also realized \cite{Stoferle2004}. Since then the Bose-Hubbard model and its generalizations form standard starting points in describing cold atoms in periodic potentials (see Ref. \cite{Lewenstein12} for a recent review). The Bose-Hubbard model allows for the quantum phase transition between Mott insulator and superfluid states. Importantly, this transition is of Berezinskii-Kosterlitz-Thouless (BKT) type \cite{Fisher89}. The correlation length is infinite on the superfluid side and it exponentially diverges near the critical point on the Mott insulator side: $\ln\xi\sim1/\sqrt{\lambda_c-\lambda}$ \cite{PaiPRL1996,Monien1998}. This means that no critical exponent $\nu$ can be defined on either side of the transition and so the above scaling expressions cannot be directly used. As far as we know, two papers report results on fidelity of the Bose-Hubbard model \cite{Buonsante2007,Rigol2013}. Both of them discuss numerical simulations showing that the minimum of fidelity is strongly shifted from the critical point even in systems composed of a few hundreds of atoms (at unit filling factor). More importantly, no convincing argument for the extrapolation of the position of the critical point from the location of the minimum of fidelity in finite-size systems has been proposed so far. Therefore, the understanding of fidelity of the Bose-Hubbard model is incomplete, which motivates our numerical ``experiment'' on this model. We consider systems larger than those previously studied, describe an unexpected sensitivity of fidelity to the boundary conditions, and systematically study fidelity around its minimum. \section{Model} We study fidelity of the Bose-Hubbard model: \begin{equation*} \hat H = - J\sum_{i=1}^M \B{\hat a_i^\dag\hat a_{i+1} + {\rm h.c.}} +\frac{U}{2}\sum_{i=1}^M \hat n_i\B{\hat n_i-1}, \end{equation*} where $M$ is the number of lattice sites, $\hat n_i=\hat a^\dag_i\hat a_i$, and the creation/annihilation operators satisfy the bosonic commutation relations. The first term describes the tunnelling between lattice sites, while the second one accounts for on-site interactions \cite{homo}. The phase diagram of this model depends on the filling factor $N/M$ and the $J/U$ ratio ($N$ is the number of atoms in a lattice). For non-integer filling factors the system is always superfluid. When $N/M$ is integer, the system is in the Mott insulator phase for $J/U < (J/U)_c$ and in the superfluid phase for $J/U > (J/U)_c$. The position of the critical point was studied in numerous theoretical papers (see Sec. II of Ref.~\cite{Rigol2013} for the recent survey of these studies). The reported values for $(J/U)_c$ range from about $0.27$ to about $0.3$. The spread of these estimations clearly highlights the complexity of the 1D Bose-Hubbard model, which unlike its famous fermionic cousin \cite{Essler}, is not integrable. Thus, its analytical studies are necessarily approximate. We will discuss below some relevant approximations and critically evaluate their applicability to the computation of fidelity. The first simplification one can invoke is the Taylor expansion (\ref{fidsus}), where the $\lambda$ and $\delta$ dependence of fidelity separate out. One is then left with the computation of fidelity susceptibility, which can be exactly written as \cite{GuReview} \begin{equation} \chi(\lambda)= \sum_{S\neq0} \frac{|\langle \phi_S(\lambda)|\hat V|\phi_0(\lambda)\rangle|^2}{\BB{E_0(\lambda) - E_S(\lambda)}^2}, \label{perturb} \end{equation} where $\hat V = \partial\hat H/\partial\lambda$ and $|\phi_S(\lambda)\rangle$ is an eigenstate of $\hat H(\lambda)$ to the eigenvalue $E_S(\lambda)$ ($S\equiv0$ corresponds to the ground state). Eq.~(\ref{perturb}) can be in principle exactly evaluated when the eigenstates $|\phi_S(\lambda)\rangle$ and their eigenenergies $E_S(\lambda)$ are exactly known. This is possible in the Bose-Hubbard model only for $\lambda=J/U=0$ (deep Mott insulator limit) and $\lambda=U/J=0$ (deep superfluid limit). \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig01.eps} \caption{(color online) Fidelity in an open Bose-Hubbard chain for different system sizes and the parameter shift $\delta=0.02$. } \label{open_delta0_02_Ndo1024} \end{figure} In the former case, one sets $\hat V= -\sum_{i=1}^M \hat a_i^\dag\hat a_{i+1} + {\rm h.c.}$ and easily finds the eigenstates of $\hat H(J/U=0)=\frac{1}{2}\sum_{i=1}^M \hat n_i\B{\hat n_i-1}$, which immediately leads to \begin{equation} \chi(J/U=0)=2 \frac{N}{M}\B{\frac{N}{M}+1}M, \label{chiJ0} \end{equation} if the periodic boundary conditions and the integer filling factor are assumed. Thus $\chi(J/U=0)$ is extensive, i.e., it scales linearly with the system size at the fixed $N/M$. In the latter case, a non-trivial result is obtained. We refer the reader to Appendix \ref{AppA} for the details of its derivation and quote here only the final expression: \begin{equation} \chi(U/J=0) = \frac{N(N-1)}{5760M^2}\B{M^4+10M^2-11}, \label{chiU0} \end{equation} valid for periodic boundary conditions and arbitrary filling factor $N/M$. This result is super-extensive. The scaling $\chi(U/J=0)\sim M^4$ appears because the low-energy spectrum of a non-interacting bosonic gas in a lattice is quadratic in quasimomentum. Noting that the crossover from the quadratic to linear spectrum happens at \cite{Stoof2001} \begin{equation*} \frac{U}{J}\sim \sin^2(k/2), \end{equation*} the qualitative departures from Eq.~(\ref{chiU0}) are expected to happen at $U/J\sim \pi^2/M^2$. For $J/U\neq0,\infty$, Eq. (\ref{perturb}) cannot be efficiently used to compute fidelity, and so its usefulness for the analytical characterization of fidelity of the Bose-Hubbard model is very limited. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig02.eps} \caption{(color online) Fidelity in a periodic Bose-Hubbard model for different system sizes and $\delta=0.02$. } \label{per_delta0_02_Ndo64} \end{figure} The next approach in line that seems to be simple enough to provide some analytical insights about fidelity of the Bose-Hubbard model is the Bogolubov approach \cite{Stoof2001}. There are, however, two problems associated with it. First, it does not describe physics of the Mott insulator phase possessing a finite excitation gap. Second, its validity even in the superfluid phase is questionable in 1D lattices due to the sizable population of the non-condensed atoms \cite{Stoof2001}. Since we are interested in the BKT transition of the Bose-Hubbard model, our studies are restricted to the 1D model. Finally, the more advanced treatment of the superfluid phase is provided by the Luttinger liquid theory; see e.g. Ref.~\cite{MonienLL} for the brief discussion of this theory in the Bose-Hubbard context. Luttinger liquid theory can be regarded as an effective low energy theory of systems, whose spectrum is linear in momentum and whose correlations decay algebraically \cite{1D,Cazalilla2004}. The computation of fidelity in the Luttinger liquid theory was presented in Refs.~\cite{LLfidel,LLfidel1,LLfidel2}. While Refs. \cite{LLfidel,LLfidel1} provide a definite expression for fidelity, Ref.~\cite{LLfidel2} argues that the zero temperature result depends on an arbitrary cut-off. This prediction is verified by comparing the Luttinger liquid theory prediction for fidelity susceptibility to the actual result in the XXZ spin chain \cite{LLfidel2}. Similarly as the Bose-Hubbard model, the XXZ model also undergoes a BKT transition. Therefore, we assume that the findings of Ref. \cite{LLfidel2} are relevant for the Bose-Hubbard model as well. Given all these complications, we focus on the numerics. We use Matrix Product State (MPS) techniques for both periodic ($\hat a_{M+1}\equiv\hat a_{1}$) and open ($\hat a_{M+1}\equiv0$) boundary conditions. Boundary conditions significantly affect the complexity of the numerical computations. For open boundary conditions the usual DMRG algorithm can be used \cite{White1993,White1992,DMRG}. This algorithm can be naturally formulated in the MPS language \cite{Schollwock2011}. It has allowed us to find the ground states of the Bose-Hubbard model for lattices containing up to 2048 sites at the unit filling factor. We have limited the bond dimension of the tensors forming the MPS representation to the 200 largest singular values . This has allowed us to obtain converged results for fidelity. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig03.eps} \caption{(color online) Black solid line shows fidelity in a periodic chain, while the dashed red line shows fidelity in an open chain. In both cases $M=64$ and $\delta=0.02$. } \label{per_vs_open} \end{figure} To find the ground state for the periodic boundary conditions, we have used imaginary time evolution technique \cite{KubaDom}. The time evolution has been performed with the TEBD algorithm \cite{Wall2009}. In periodic chains the singular values decrease more slowly than in the open ones \cite{Verstraete2004,Pippan2010}. To obtain the converged fidelity, the bond dimension 220 of the MPS vectors was necessary even for relatively small periodic systems consisting of 64 lattice sites with unit filling. In both cases the local Hilbert space has been cut to the subspace allowing at most 6 particles per site. The accuracy of our computations is discussed in Appendix \ref{AppB}. Finally, we mention that for $M=12$ and both open and periodic chains we have computed fidelity via exact diagonalization and compared such results to the DMRG numerics. The two approaches agree with each other. Previous studies of fidelity of the Bose-Hubbard model at unit filling factor were restricted to $M\le12$ on periodic lattices (exact diagonalization) \cite{Buonsante2007} and $M\le120$ on the open lattices (DMRG) \cite{Rigol2013}. Ref.~\cite{Buonsante2007} reports results on both fidelity and fidelity susceptibility, while Ref. \cite{Rigol2013} focuses on fidelity susceptibility. \section{Fidelity} \label{sec_fidelity} From now on, we set $N=M$, i.e., we study the unit filling case. Typical results that we obtain in open and periodic systems are presented in Figs. \ref{open_delta0_02_Ndo1024} and \ref{per_delta0_02_Ndo64}, respectively. First, we notice that the minimum of fidelity lies on the Mott insulator side. This is an expected feature because the ground states near the critical point change more rapidly on the Mott insulator side of the transition (see e.g. the exponential dependence of the correlation length in the Mott phase near the critical point). \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig04.eps} \caption{(color online) The minimum of fidelity as a function of the system size (open boundary conditions). Pluses come from numerics done for $\delta=0.02$ and $M=32,64,128,256,512,1024,2048$. The line represents Eq.~(\ref{FJstar}) with $\alpha\simeq2.51$ obtained from the fit. } \label{Fwmin_0_02_open} \end{figure} Second, there are easy-to-notice differences between the periodic and open chain results. As shown in Fig. \ref{per_vs_open}, the minimum in the open chain is much more shallow and more distant from the critical point than the one in the periodic chain. As expected, periodic chains probe quantum criticality more robustly. However, the magnitude of the difference between the two cases is surprising. It prompts separate numerical studies of the periodic and open chains. \section{Open boundary conditions} \label{sec_open} We start the discussion from looking at the value of fidelity at the minimum. We denote the position of the minimum as $J^*/U$. Our numerics supports the following expression \begin{equation} F(J^*/U,\delta) = \exp\B{-\alpha M\delta^2}, \label{FJstar} \end{equation} where $\alpha$ is some constant. Note that since we work with $N/M=1$, it is impossible to decide whether there should be $M$ or $N$ in Eq. (\ref{FJstar}). The numerical evidence supporting Eq. (\ref{FJstar}) is discussed in detail in Figs. \ref{Fwmin_0_02_open} and \ref{Fwmin_Ln_open}. It is worth to stress that Eq. (\ref{FJstar}) works in these figures also when $F(J^*/U,\delta)\ll1$, i.e., when fidelity susceptibility fails to account for fidelity. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig05.eps} \caption{(color online) Scaling of the minimum of fidelity with the system size $M$ and the parameter shift $\delta$ (open boundary conditions). Panel (a): Pluses come from numerics done for $M=32,64,128,256,512,1024,2048$ \cite{remark_spline}. The line represents the fit: $\ln\BB{-\ln F\B{J^*/U}} = -7.142 + 1.032 \ln M$. Panel (b): Pluses come from numerics done for $\delta=0.005, 0.01, 0.02, 0.03, 0.04$ \cite{remark_spline}, while the line represents the fit: $\ln\BB{-\ln F\B{J^*/U}} = 7.833 + 1.994 \ln\delta$. Both fits accurately support the key features of Eq.~(\ref{FJstar}): $\ln F\B{J^*/U} \sim -M$ tested in panel (a) and $\ln F\B{J^*/U}\sim-\delta^2$ tested in panel (b). } \label{Fwmin_Ln_open} \end{figure} The natural question arising now is the following: Can we extrapolate the position of the critical point by studying the scaling of $J^*/U$ with the system size $M$? We were unable to find an extrapolation scheme that would give us the correct location of the critical point. We will mention below two attempts. First, we have tested \begin{equation} \frac{J^*(M)}{U} = \frac{J^*(\infty)}{U} - \frac{a}{M^b}, \label{Jfit} \end{equation} with fitted parameters $J^*(\infty)/U=0.2114\pm0.0003$, $a=0.53\pm0.02$, and $b=0.73\pm0.02$ obtained via standard {\it Mathematica} fitting procedure (Fig. \ref{Jstart_delta0_02_open}). Thus the extrapolated position of the minimum in the infinite system, $J^*(\infty)/U\simeq0.2114$, is nowhere near the expected location of the critical point $(J/U)_c$. The fit (\ref{Jfit}) was proposed in Ref. \cite{Buonsante2007} studying fidelity of the Bose-Hubbard model, but no theory supporting it was discussed there. This fit was also used to extrapolate the position of the maximum of fidelity susceptibility in the 2D Ising model in a transverse field \cite{ABQ2010}. The fit provided the correct location of the critical point and the critical exponent $\nu$ (following Ref. \cite{Paris}, it was verified in Ref. \cite{ABQ2010} that $b=1/\nu$ in the 2D Ising model in a transverse field; in a 1D Ising model in a transverse field $b=2/\nu$, which was shown in Refs. \cite{Paris,BDfid3}). In the Bose-Hubbard model, however, the critical exponent $\nu$ is undefined and it is unclear to us how to improve the fit (\ref{Jfit}) to properly extrapolate the position of the critical point. Second, we tried \begin{equation} \frac{J^*(M)}{U} = \frac{J^*(\infty)}{U} - a\frac{\ln M}{M^b} \label{Jfit_ln} \end{equation} to include logarithmic finite-size corrections near the BKT transition (Fig. \ref{Jstart_delta0_02_open}). The fit has provided $J^*(\infty)/U=0.2106\pm0.0001$, $a=0.375\pm0.007$, and $b=0.991\pm0.006$. While this value for $J^*(\infty)/U$ is very close to that obtained previously, the quality of this fit is better (the inset of Fig. \ref{Jstart_delta0_02_open}). This suggests that the finite-size correction to the position of the minimum of fidelity scales as $\ln M/M$. To quantify the difference between the fits, we provide chi-squared, i.e., the sum of the squared differences between the fitted curve and the numerical data. It equals about $2.7\times10^{-7}$ for the fit (\ref{Jfit}) and $3.9\times10^{-8}$ for the fit (\ref{Jfit_ln}). Thus, the fit (\ref{Jfit_ln}) is indeed better than the (\ref{Jfit}) one. Finally, we look closer at fidelity per site, i.e., $\ln F/M$, which is plotted in Fig. \ref{lnF_M}. As discussed in Sec. \ref{sec_introduction}, this quantity is expected to have a finite non-zero value in the thermodynamic limit in the systems, where the correlation length diverges algebraically. This was predicted and observed in several models \cite{Zhou,BDfid1,SenPRB2012,Adamski2013}. It is unclear from Fig. \ref{lnF_M} whether the same holds for the Bose-Hubbard model. Further studies focusing on larger systems are needed to settle the system-size dependence of fidelity per site. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig06.eps} \caption{(color online) Position of the minimum of fidelity as a function of the system size (open boundary conditions). X's show numerics done for $M=32,64,128,256,512,1024,2048$ and $\delta=0.02$ \cite{remark_spline}. The red (dashed) line is the fit (\ref{Jfit}) while the blue (solid) corresponds to (\ref{Jfit_ln}). The inset shows the zoom for large $M$ to enable comparison of both fits. } \label{Jstart_delta0_02_open} \end{figure} \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig07.eps} \caption{(color online) Fidelity per site for $\delta=0.02$ and three system sizes: $M=256,512,1024$ (black solid, red dashed, green dot-dashed, respectively). This plot is made from data from the lower panel of Fig. \ref{open_delta0_02_Ndo1024}. } \label{lnF_M} \end{figure} Next, we turn our attention to the fidelity susceptibility $\chi(J/U)$. Since fidelity (\ref{F}) is symmetric with respect to the $\delta\to-\delta$ transformation, we compute \begin{equation*} 2\frac{1-F(J/U,\delta)}{\delta^2} \end{equation*} for $\delta=0.005,0.01$ and all $J/U$'s of interest and fit it with \begin{equation} \chi(J/U) + \psi(J/U)\delta^2, \label{extrap} \end{equation} where $\chi(J/U)$ and $\psi(J/U)$ are the fitting parameters. We plot such obtained fidelity susceptibility in Fig. \ref{fidsus_extrap_open}. The deep Mott insulator regime, $J/U\to0$, can be computed exactly. For open boundary conditions we get $\chi(0)=4(M-1)$ at the unit filling factor. Similarly as in Ref.~\cite{Rigol2013}, we observe that the $\chi/M$ curves cross near the critical point, i.e., around $J/U\simeq0.287$ (the inset of Fig. \ref{fidsus_extrap_open}). Indeed, we estimate from Fig. 2 of Ref.~\cite{Rigol2013} that the crossing occurs there at $J/U\simeq 1/3.5\simeq0.286$, which is completely consistent with our result. One could speculate that the position of the critical point might be linked to such a crossing, i.e., to the point where $\chi\sim M$. One should be careful, however, as such a crossing is absent in periodic systems discussed in Sec. \ref{sec_periodic}. We have also studied the position of the maximum of fidelity susceptibility. For example, setting $\delta=0.02$ as in Fig. \ref{Jstart_delta0_02_open}, we have found that the minima of fidelity and maxima of fidelity susceptibility roughly coincide (as expected, the larger the system size is, the bigger the discrepancy is). We have fitted the position of the maximum of fidelity susceptibility for $M=32,64,128,256,512,1024,2048$ with the power-law (\ref{Jfit}) and again obtained $J^*(\infty)/U =0.2121\pm0.0002$. \section{Periodic boundary conditions} \label{sec_periodic} This section presents our results on fidelity in the periodic Bose-Hubbard model. Due to the numerical limitations, the system sizes studied here are restricted to $M\le64$, which is the factor of $2^5$ smaller than the range of the system sizes considered in Sec. \ref{sec_open}, but the factor of $2^2$ larger than the system sizes considered in Ref. \cite{Buonsante2007}. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig08.eps} \caption{(color online) Fidelity susceptibility per site (open boundary conditions). The black solid, red dashed, green dot-dashed lines represent data for $M=256,512, 1024$, respectively. The inset enlarges the part of the main plot, where the curves cross. } \label{fidsus_extrap_open} \end{figure} As in Sec. \ref{sec_open}, we find numerical evidence towards the proposed scaling of the position of the minimum of fidelity (\ref{FJstar}). This is presented in detail in Fig. \ref{periodic_M64_spline}. There is no significant difference here between the open and periodic results, compare Figs. \ref{Fwmin_Ln_open} and \ref{periodic_M64_spline}. On the other hand, the extrapolation of the position of the minimum of fidelity through Eq.~(\ref{Jfit}) provides a markedly different result with respect to what we have found in Sec. \ref{sec_open}. This extrapolation is discussed in Fig. \ref{Jstar_delta0_02_periodic}. The extrapolated position of the minimum in an infinite system is $J^*(\infty)/U=0.270\pm0.008$, which agrees with some of the previous estimations of the position of the critical point \cite{Rigol2013}. We see also from the fit that the convergence with the system size to the asymptotic result is slow: $b=0.44\pm0.05$ in formula (\ref{Jfit}). \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig09.eps} \caption{(color online) Scaling of the minimum of fidelity with the system size $M$ and the parameter shift $\delta$ in a periodic Bose-Hubbard chain. Upper panel: X's show numerics for $M=16,32,48,64$ \cite{remark_spline}. The line represents the fit $\ln\BB{-\ln F\B{J^*/U}} = -6.787 + 0.998\ln M$. Lower panel: X's show numerics for $\delta =0.005,0.01,0.02,0.03,0.04$. The line represents the fit $\ln\BB{-\ln F\B{J^*/U}} = 5.104 + 1.982\ln\delta$. } \label{periodic_M64_spline} \end{figure} \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig10.eps} \caption{(color online) Position of the minimum of fidelity as a function of the system size $M$ in the periodic Bose-Hubbard model. Stars show numerics done for $\delta=0.02$ \cite{remark_spline}, while the line represents Eq.~(\ref{Jfit}) with $J^*(\infty)/U\simeq0.27$, $a\simeq0.33$, and $b\simeq0.44$ (all coming from the fit). } \label{Jstar_delta0_02_periodic} \end{figure} Next, we focus on fidelity susceptibility $\chi(J/U)$, which is plotted in Fig. \ref{fidsus_extrap_periodic}. First, we notice again that $\chi/M\to4$ as $J/U\to0$. This follows from Eq. (\ref{chiJ0}) taken at $N/M=1$. Moreover, it is also seen that $\chi/M$ is about the same at the maximum for all three curves ($M=16,32,64$). Second, we see from Fig. \ref{fidsus_extrap_periodic} that the fidelity susceptibility curves do not cross near the critical point in the periodic chain. However, they do cross near the critical point in the open chain (the inset of Fig. \ref{fidsus_extrap_open}). This suggests that the inhomogeneities appearing near the edges in the open chain are responsible for the crossing in the open system (see also Sec. \ref{sec_discussion}). If this is indeed the case, it is even more puzzling why the crossing occurs near the critical point in Fig. \ref{fidsus_extrap_open}. Third, we have extrapolated the position of the maximum of fidelity susceptibility with Eq. (\ref{Jfit}). Using data for $M=16,32,48,64$, we have obtained from the fit (\ref{Jfit}) that $J^*(\infty)/U=0.289\pm0.008$, which again agrees with some earlier studies of the position of the critical point \cite{Rigol2013}. Finally, at the risk of stating the obvious, we mention that it is desirable to extend these computations to larger (more critical) systems. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig11.eps} \caption{(color online) Fidelity susceptibility per site (periodic boundary conditions). The black solid, red dashed, green dot-dashed lines represent data for $M=16,32, 64$, respectively. Fidelity susceptibility is obtained through the extrapolation procedure discussed around Eq. (\ref{extrap}). } \label{fidsus_extrap_periodic} \end{figure} \section{Discussion} \label{sec_discussion} Our results show that there is a significant difference between fidelity in the open and periodic chains. The ground states obtained in the two cases mainly differ by the occupation of the lattice sites, which is translationally invariant in a periodic chain and inhomogeneous near the edges in the open problem (Fig. \ref{gestosc}). We believe that this inhomogeneity makes the difference, but do not have the explanation of why it is so large. It is worth to realize that such an inhomogeneity did not cause much trouble in the determination of the location of the critical point through the studies of the decay of the correlation functions \cite{MonienLL}. The ground states in these studies were obtained through the open-chain DMRG simulations, so they were the same as in our calculations. As expected, the influence of inhomogeneities on the system properties near the center was marginal for large-enough systems. Thus, one could obtain reliable results by computing the two-point correlation functions near the center. This approach differs from the fidelity approach in one important aspect. Namely, the parts of the system near the center and those near the edges equally contribute to fidelity. We see no straightforward way to factor out the influence of the edges on fidelity. This complication follows from the simplicity of the fidelity approach, which democratically collapses all information about two ground state wave-functions into a single number. \begin{figure} \includegraphics[width=\columnwidth,clip=true]{fig12.eps} \caption{Density of atoms $\langle\hat n_i\rangle$ in the ground state of the open chain Bose-Hubbard model computed for $M=256$ and $J/U=0.21$, i.e., near the minimum of fidelity (Fig. \ref{open_delta0_02_Ndo1024}). } \label{gestosc} \end{figure} We would like to stress that better understanding of our numerical results should come from the analytical derivation of the finite system-size scaling of the position of the minimum of fidelity (maximum of fidelity susceptibility) in a BKT transition. The derivation of the robust system-size-dependent scaling expressions for a BKT transition is notoriously difficult. This is seen, e.g., from the spread of the estimates of the location of the critical point. Interestingly enough, our numerics shows that the scaling law capturing the behavior of fidelity should distinguish between the open and periodic boundary conditions. Finally, we mention that our numerically-supported scaling law for fidelity at its minimum, Eq. (\ref{FJstar}), also awaits analytical explanation. Summarizing, the key findings of this manuscript are the following. First, we have proposed and numerically verified an expression relating fidelity at the minimum, the system size, and the parameter shift (\ref{FJstar}). Second, we have found and numerically characterized the striking difference between fidelity in the open and periodic chains. Third, an exact analytical expression for fidelity susceptibility in the deep superfluid limit has been derived. We expect that these results should motivate further studies of fidelity in the Bose-Hubbard model ultimately leading to the complete understanding of the BKT transition in this model. \begin{center} {\bf ACKNOWLEDGMENTS} \end{center} We thank Dominique Delande for his comments about the the manuscript and suggesting logarithmic correction formula Eq.~\eqref{Jfit_ln}. BD and JZ acknowledge support of the Polish National Science Center grant DEC-2012/04/A/ST2/00088. MŁ acknowledges support the Polish National Science Center project no. 2013/08/T/ST2/00112 for the PhD thesis, special stipend of Smoluchowski Scientific Consortium ``Matter Energy Future'' and support by Jagiellonian University International Ph.D Studies in Physics of Complex Systems (Agreement No. MPD/2009/6 with Foundation for Polish Science). Simulations were performed at ACK Cyfronet AGH under the PL-Grid project and at IF UJ using the Deszno supercomputer purchased in the framework of the Polish Innovation Economy Operational Program (POIG.02.01.00-12-023/08).
\section{Introduction} Since decades the anomalous magnetic moments of electron and muon, $a_e$ and $a_\mu$, are used to perform precision tests of QED.\footnote{ See Refs.~\cite{Melnikov:2006sr,Jegerlehner:2009ry,Miller:2012opa} for comprehensive reviews.} In fact, in the case of the electron the experimental measurements and theoretical predictions have reached a precision which allows for the most precise extraction of the fine structure constant $\alpha$. In contrast to $a_e$ there is a sizable hadronic contribution to $a_\mu$ which involves as input measurements of the total cross section $\sigma(e^+e^-\to\mbox{hadrons})$ at low energies. Although all ingredients are measured and computed to high precision there is a discrepancy of about $3\sigma$ between the measured and predicted value for $a_\mu$~\cite{Jegerlehner:2011ti,Hagiwara:2011af,Davier:2010nc}. In this context it is interesting to mention that this difference is of the same order of magnitude as the four-loop QED contribution which to date has only been computed by one group~\cite{Aoyama:2012wk}. Thus, it is important to provide an independent cross check for this ingredient. First results have been obtained in Ref.~\cite{Laporta:1993ds,Aguilar:2008qj,Lee:2013sx}. In particular, in Ref.~\cite{Lee:2013sx} the contribution from Feynman diagrams containing two or three closed electron loops have been computed. In this letter we provide a further step towards the full four-loop QED corrections to $a_e$ and $a_\mu$ and compute the part induced by heavy leptons. In the case of the muon this means that Feynman diagrams have to be considered which contain closed tau loops and both closed muon and tau loops are present for $a_e$. Such contributions appear for the first time at two-loop order (cf. Fig.~\ref{fig::diags}) and have been computed in Refs.~\cite{Elend:1966,Passera:2004bj,Passera:2006gc}. Also the three-loop result is known in analytic form for arbitrary lepton masses~\cite{Samuel:1990qf,Laporta:1992pa,Laporta:1993ju,Czarnecki:1998rc,Kuhn:2003pu,Friot:2005cu} (see also~\cite{Passera:2004bj,Passera:2006gc}). At four loops, however, only numerical results are available~\cite{Kinoshita:2005zr,Aoyama:2007mn,Aoyama:2012wj,Aoyama:2012wk}. We want to cross-check these results using a different method which leads to analytic results for $a_e$ and $a_\mu$. It is based on asymptotic expansion~\cite{Smirnov:2013} in the ratio of the light and heavy lepton mass, $M_l$ and $M_h$, which leads to a factorization of the two-scale integrals into simpler ones with at most one mass scale. The latter can be computed analytically. We have computed three terms of the expansion in $M_l^2/M_h^2$. For the perturbative expansion of the QED corrections to $a_e$ and $a_\mu$ we take over the commonly used notation from Refs.~\cite{Aoyama:2012wj,Aoyama:2012wk} and write ($l=e,\mu$) \begin{eqnarray} a_l &=& \sum_{n\ge1} \left(\frac{\alpha}{\pi}\right)^n a_l^{(2n)} \,, \end{eqnarray} where $a_l^{(2n)}$ can be written in the form \begin{eqnarray} a_e^{(2n)} &=& A_{1,e}^{(2n)} + A_{2,e}^{(2n)}(M_e/M_\mu) + A_{2,e}^{(2n)}(M_e/M_\tau) + A_{3,e}^{(2n)}(M_e/M_\mu,M_e/M_\tau) \,,\nonumber\\ a_\mu^{(2n)} &=& A_{1,\mu}^{(2n)} + A_{2,\mu}^{(2n)}(M_\mu/M_e) + A_{2,\mu}^{(2n)}(M_\mu/M_\tau) + A_{3,\mu}^{(2n)}(M_\mu/M_e,M_\mu/M_\tau) \,. \end{eqnarray} In this paper we compute $A_{2,e}^{(2n)}(M_e/M_\mu)$, $A_{2,e}^{(2n)}(M_e/M_\tau)$, $A_{3,e}^{(2n)}(M_e/M_\mu,M_e/M_\tau)$ and $A_{2,\mu}^{(2n)}(M_\mu/M_\tau)$ to four-loop order. We have also computed the corresponding two- and three-loop results and found complete agreement with the literature. Before starting the actual calculation let us consider the parametrical size of our corrections. Actually, the heavy-lepton contribution decouples in the limit $M_h\to\infty$ and leads to a $M_l^2/M_h^2$ suppression. Thus the four-loop corrections to $a_\mu$ have the form $(\alpha/\pi)^4 \times M_\mu^2/M_\tau^2$ where $M_\mu^2/M_\tau^2$ is of order $10^{-3}$. On the other hand we have $\alpha/\pi \approx 2\cdot 10^{-3}$ which is of the same order of magnitude. Thus, from the parametric point of view the four-loop corrections induced by heavy leptons could be of the same order as the five-loop results obtained in Ref.~\cite{Aoyama:2012wk}. Note, however, that in practice the contributions involving electron loops are large ($A_{2,e}^{(10)}(M_e/M_\mu)$ is of order $10^3$) whereas the heavy-lepton contributions have coefficients which are at most of order $10$. In the case of $a_e$ the ratio of the lepton masses is much smaller than for $a_\mu$ (Note that $M_e^2/M_\mu^2 = {\cal O}(10^{-5})$, $M_e^2/M_\tau^2 = {\cal O}(10^{-8})$) and thus the corresponding corrections are less relevant. Nevertheless, for completeness we provide also those results. The remainder of the paper is structured as follows: In the next Section we briefly discuss some technical details which are important for our calculation. Section~\ref{sec::res} is devoted to the presentation and discussion of the results. In particular, we compare to the numerical results of Refs.~\cite{Aoyama:2012wj,Aoyama:2012wk}. We conclude in Section~\ref{sec::concl}. In the Appendix we present results for the on-shell counterterms for the fine structure constant, the lepton mass and the lepton wave function. \section{Some technical details} \begin{figure}[t] \begin{center} \leavevmode \epsfxsize=0.9\textwidth \epsffile[80 390 500 750]{feyndias.ps} \end{center} \caption[]{\label{fig::diags} Sample Feynman diagrams contributing to the electron and muon $(g-2)$ containing a heavy lepton at two, three and four loops. Thin and thick solid lines represent light and heavy leptons, respectively, and wavy lines denote photons. The symbols below the four-loop diagrams label the individual diagram classes and are taken over from Refs.~\cite{Aoyama:2012wj,Aoyama:2012wk}.} \end{figure} Typical Feynman diagrams to be considered for the heavy-lepton contribution of $a_e$ and $a_\mu$ are shown in Fig.~\ref{fig::diags}. At two-loop order only one diagram has to be considered.\footnote{Note that the contribution where the photon couples to the closed lepton loop vanishes due to Furry's theorem~\cite{Furry:1937zz}.} At three loop-order 60 and at four loops 1169 Feynman diagrams are generated. In the following discussion we denote the heavy lepton mass by $M_h$ and the light one by $M_l$. For the generation of the diagrams we use {\tt QGRAF}~\cite{Nogueira:1991ex} and transform the amplitudes with the help of {\tt q2e}~\cite{Harlander:1997zb,Seidensticker:1999bb} to a {\tt FORM}~\cite{Vermaseren:2000nd} readable output. In a next step we apply {\tt exp}~\cite{Harlander:1997zb,Seidensticker:1999bb} to perform an asymptotic expansion for $M_h\gg M_l$. At two-loop order [see Fig.~\ref{fig::ae}(a)] this leads to two so-called sub-diagrams which have to be Taylor-expanded in their external momenta. The first sub-diagram is given by the whole two-loop diagram which, after expansion, leads to two-loop vacuum integrals. The second contribution consists of a product of two one-loop diagrams. After expanding the one-loop vacuum integral in the external momentum one has to insert the result in the remaining one-loop on-shell integral and integrate over the second loop momentum. The described procedure is illustrated in the second line of Fig.~\ref{fig::ae}(a). Fig.~\ref{fig::ae}(b) shows a four-loop example which demonstrates the typical situation at this order: the original four-loop two-scale integral is transformed to a sum of products of $N$-loop vacuum integrals with scale $M_h$ and $(4-N)$-loop on-shell integrals with $q^2=M_l^2$ where $N=1,2,3$ or $4$ and $q$ is the momentum flowing through the external lepton line. All integrals only contain one mass scale and are thus significantly simpler than the original one. \begin{figure}[t] \begin{center} \begin{tabular}{c} \leavevmode \epsfxsize=0.9\textwidth \epsffile[80 600 500 750]{feyndias1.ps} \\ (a) \\ \leavevmode \epsfxsize=0.9\textwidth \epsffile[80 500 550 750]{feyndias2.ps} \\ (b) \end{tabular} \end{center} \caption[]{\label{fig::ae} Graphical examples for the application of the asymptotic expansion at two (a) and four (b) loops. Thick solid, thin solid and wavy lines represent heavy and light leptons and photons, respectively. In (b) only four representative sub-diagrams are shown. Altogether there are eight contributions.} \end{figure} Both vacuum and on-shell integrals are reduced to master integrals with the help of {\tt FIRE}~\cite{Smirnov:2008iw,Smirnov:2013dia}.\footnote{We thank A.V.~Smirnov and V.A.~Smirnov for allowing us to use the unpublished {\tt C++} version of {\tt FIRE}.} The master integrals are all known analytically and are taken from Refs. \cite{Laporta:2002pg,Chetyrkin:2004fq ,Kniehl:2005yc ,Schroder:2005db ,Schroder:2005hy ,Schroder:2005va ,Bejdakic:2006vg ,Chetyrkin:2006dh ,Kniehl:2006bf ,Kniehl:2006bg} and Refs.~\cite{Laporta:1996mq,Melnikov:2000qh,Lee:2010ik}, respectively. We renormalize our results in the on-shell scheme. For this purpose we need the counter\-term for the fine structure constant, the (light) lepton mass and lepton wave function to three loops. The corresponding analytic results for the case of a massless second lepton loop can be found in Ref.~\cite{Broadhurst:1991fi}, Refs.~\cite{Chetyrkin:1999ys,Chetyrkin:1999qi,Melnikov:2000qh,Marquard:2007uj} and~\cite{Melnikov:2000zc,Marquard:2007uj}, respectively. In our case the opposite limit of a heavy lepton is needed which we computed ourselves using the rules of asymptotic expansion as described above. The analytic expressions are presented in the Appendix for completeness. Our results for the leading term of the lepton mass counterterm agrees with Ref.~\cite{Sturm:2013uka} and the one for the charge counterterm is easily obtained from the general expression presented in Ref.~\cite{Lee:2013sx}. To our knowledge the three-loop result for the on-shell wave function renormalization constant is new. In addition, the heavy-lepton mass has to be renormalized in the two- and three-loop expression. The corresponding two-loop counterterm can be found in Refs.~\cite{Gray:1990yh,Bekavac:2007tk}. Note that the two-loop counterterm which has to be inserted into the two-loop vertex diagram of Fig.~\ref{fig::diags} involves contributions with a closed light lepton loop. The expansion of this contribution in $M_l\ll M_h$ contains both even and odd powers in $M_l/M_h$ which is the reason for the occurrence of odd expansion terms in $A^{(8)}_{2,\mu}$ (cf. Section~\ref{sec::res}). There are several checks on the correctness of our result. Besides the obvious ones like finiteness we have performed two independent calculations. In particular, two independent routines for the decomposition of the scalar products in the numerator and the preparation of the {\tt FIRE} input has been written. Furthermore for our calculation we have used general QED gauge parameter up to linear terms in $\xi$ and have checked that the final result of the leading term in the inverse heavy lepton expansion, i.e. the one proportional to $M_l^2/M_h^2$, is $\xi$-independent. Due to the complexity of the calculation we have used Feynman gauge for the higher order expansion terms. \section{\label{sec::res}Results and discussion} Let us in a first step present the analytic results of our calculation. The four-loop contribution to $a_\mu$ from Feynman diagrams involving a virtual tau lepton loop is given by \begin{eqnarray} A^{(8)}_{2,\mu}(M_\mu/M_\tau) &=& \left(\frac{M_\mu}{M_\tau}\right)^2 \bigg(\frac{37448693521}{2286144000}+\frac{89603}{16200}P_ +\frac{52}{675}P_5+\frac{4 \pi^2 \zeta_3}{15}+\frac{5771 \ln(2) \pi^4}{32400}\nonumber\\&& \qquad-\frac{3851 \pi^2}{3600}-\frac{25307 \zeta_5}{1440}-\frac{37600399 \pi^4}{27216000}+\frac{35590996657 \zeta_3}{508032000}\nonumber\\&& \qquad+\ln\frac{M_\mu^2}{M_\tau^2} \left(-\frac{38891}{12150}+\frac{19 \pi^2}{135}+\frac{3 \zeta_3}{2}\right)+\frac{359}{1080}\ln^2\frac{M_\mu^2}{M_\tau^2}\bigg)\nonumber\\&& + \left(\frac{M_\mu}{M_\tau}\right)^3 \frac{\pi^2}{90}\nonumber\\&& + \left(\frac{M_\mu}{M_\tau}\right)^4 \bigg(\frac{392783023945426851403}{73077446697615360000}-\frac{3355249339331 \pi^4}{2575112601600}\nonumber\\&& \qquad+\frac{74184592369}{14306181120} P_4 +\frac{557}{9450}P_5\nonumber\\&& \qquad-\frac{378681587 \pi^2}{114307200}-\frac{652 \ln(2) \pi^2}{1215}+\frac{26783 \ln(2) \pi^4}{226800}\nonumber\\&& \qquad+\frac{725750082915523417 \zeta_3}{10310750856806400}+\frac{66211 \pi^2 \zeta_3}{22680}-\frac{425983 \zeta_5}{30240}\nonumber\\&& \qquad+\ln\frac{M_\mu^2}{M_\tau^2} \left(-\frac{1922512966823}{1229031014400}+\frac{47899 \pi^2}{816480}+\frac{81782993 \zeta_3}{123863040}\right)\nonumber\\&& \qquad+\frac{193032971}{457228800}\ln^2\frac{M_\mu^2}{M_\tau^2}-\frac{24037}{362880}\ln^3\frac{M_\mu^2}{M_\tau^2}\bigg)\nonumber\\&& + \left(\frac{M_\mu}{M_\tau}\right)^5 \left(\frac{2671 \pi^2}{176400}+\frac{\pi^2}{140} \ln\frac{M_\mu^2}{M_\tau^2}\right)\nonumber\\&& + \left(\frac{M_\mu}{M_\tau}\right)^6 \bigg( \frac{326292200455466311953239}{4974581098834034688000}+\frac{4785889811617 \pi^2}{1234517760000}\nonumber\\&& \qquad+\frac{989648650006997}{191294078976000} P_4 +\frac{7001}{207900}P_5\nonumber\\&& \qquad-\frac{27903657664078117 \pi^4}{11477644738560000}+\frac{711883 \ln(2) \pi^4}{9979200}-\frac{148 \ln(2) \pi^2}{315}\nonumber\\&& \qquad+\frac{6446695611351419899 \zeta_3}{66315280711680000}-\frac{18533 \pi^2 \zeta_3}{6048}+\frac{179971 \zeta_5}{24192}\nonumber\\&& \qquad+\ln\frac{M_\mu^2}{M_\tau^2} \bigg(-\frac{2631561259843654279}{132735349555200000}+\frac{17955349 \pi^2}{489888000}\nonumber\\&& \qquad\qquad\qquad+\frac{314284167899 \zeta_3}{19818086400}\bigg)\nonumber\\&& \qquad+\frac{22710352067}{58786560000} \ln^2\frac{M_\mu^2}{M_\tau^2}-\frac{101799017}{979776000} \ln^3\frac{M_\mu^2}{M_\tau^2}\bigg)\nonumber\\&& + \left(\frac{M_\mu}{M_\tau}\right)^7 \left(\frac{79 \pi^2}{15120}+\frac{\pi^2}{60} \ln\frac{M_\mu^2}{M_\tau^2}\right) + {\cal O}\left( \left(\frac{M_\mu}{M_\tau}\right)^8 \right) \nonumber \\ &\approx& 0.0421670 + 0.0003257 + 0.0000015 \,, \label{eq::A8} \end{eqnarray} where $P_4=24a_4+\ln^4(2)-\ln^2(2) \pi^2$, $P_5=120a_5-\ln^5(2)+\frac{5}{3}\ln^3(2) \pi^2$, $a_n=\mbox{Li}_n(1/2)$ and $\zeta_n$ is Riemann's zeta function. In the last line of Eq.~(\ref{eq::A8}) the analytic expression has been evaluated numerically using \mbox{$M_\mu/M_\tau=5.94649(54) \cdot 10^{-2}$}~\cite{Beringer:1900zz}. Furthermore the contributions from $(M_\mu/M_\tau)^n$ and $(M_\mu/M_\tau)^{n+1}$ ($n=2,4,6$) have been combined. One observes a rapid convergence of the series in $M_\mu/M_\tau$ which suggests that with each additional order one gains two significant digits. To be conservative we take 10\% of the last term in Eq.~(\ref{eq::A8}) as error estimate which leads to our final result \begin{eqnarray} A^{(8)}_{2,\mu}(M_\mu/M_\tau)&\approx& 0.0424941(2)(53) \,, \label{eq::A8mu} \end{eqnarray} where the second uncertainty reflects the error in the input quantity $M_\mu/M_\tau$. The result in~(\ref{eq::A8mu}) agrees with the one from Ref.~\cite{Aoyama:2012wk} $A^{(8)}_{2,\mu}(M_\mu/M_\tau) = 0.04234(12)$, however, our number is significantly more precise. For completeness we also provide the numerical results for the two- and three-loop contributions which read\footnote{Since analytic expressions are available the uncertainties for the two- and three-loop results are due to the errors in the lepton masses.} \begin{eqnarray} A^{(4)}_{2,\mu}(M_\mu/M_\tau) &=& 7.8079(14) \cdot 10^{-5} \,,\nonumber\\ A^{(6)}_{2,\mu}(M_\mu/M_\tau) &=& 3.6063(12) \cdot 10^{-4} \,. \end{eqnarray} It is interesting to note that the three-loop coefficient is only a factor of five larger than the two-loop one whereas $A^{(8)}_{2,\mu}(M_\mu/M_\tau)$ is about 100 times larger than $A^{(6)}_{2,\mu}(M_\mu/M_\tau)$. Using $\alpha = 1/137.035999174$~\cite{Aoyama:2012wj} one finally obtains for the $\tau$-loop contribution to $a_\mu$ \begin{eqnarray} 10^{11} \times a_\mu\Big|_{\tau \rm loops} &=& 42.13 + 0.45 + 0.12 \,, \label{eq::amu_tau} \end{eqnarray} where the numbers on the right-hand side correspond to the two-, three- and four-loop contribution. The numbers in Eq.~(\ref{eq::amu_tau}) have to be compared with the universal contributions contained in $A_{1,\mu}$ which read~\cite{Aoyama:2012wk} \begin{eqnarray} 10^{11} \times a_\mu\Big|_{\rm univ.} \!\!\!&=&\!\!\! 116\,140\,973.21 - 177\,230.51 + 1\,480.42 - 5.56 + 0.06 \,, \end{eqnarray} where the individual terms on the right-hand side represent the results from one to five loops. \begin{table}[t] \begin{center} \begin{tabular}{c||l|l} group & \multicolumn{2}{c}{$10^2 \cdot A^{(8)}_{2,\mu}(M_\mu/M_\tau)$}\\ \hline & this work & \cite{Aoyama:2012wk}\\ \hline I(a) & \hphantom{$-$}0.00324281(2) & \hphantom{$-$}0.0032(0)\\ I(b) + I(c) + II(b) + II(c) & $-0.6292808(6)$ & $-0.6293(1)$\\ I(d) & \hphantom{$-$}0.0367796(4) & \hphantom{$-$}0.0368(0)\\ III & \hphantom{$-$}4.5208986(6) & \hphantom{$-$}4.504(14)\\ II(a) + IV(d) & $-2.316756(5)$ & $-2.3197(37)$\\ IV(a) & \hphantom{$-$}3.851967(3) & \hphantom{$-$}3.8513(11)\\ IV(b) & \hphantom{$-$}0.612661(5) & \hphantom{$-$}0.6106(31)\\ IV(c) & $-1.83010(1)$ & $-1.823(11)$ \end{tabular} \end{center} \caption{\label{tab::mu}Mass-dependent corrections to $a_\mu$ at four-loop order as obtained in this paper and the comparison with Refs.~\cite{Aoyama:2012wk}. The uncertainties assigned to our numbers correspond to 10\% of the highest available expansion terms, i.e., the ones of order $(M_\mu/M_\tau)^6$ and $(M_\mu/M_\tau)^7$. Uncertainties from the muon and tau lepton mass are not shown.} \end{table} The detailed comparison with Tab.~I of Ref.~\cite{Aoyama:2012wk} is shown in Tab.~\ref{tab::mu} where our result is split into eight different groups. In the first column the notation of~\cite{Aoyama:2012wk} is used to indicate the contributions which have to be summed\footnote{We add the uncertainties of Ref.~\cite{Aoyama:2012wk} in quadrature when adding results from different groups.} in order to compare with our numbers. Within the numerical uncertainties we observe good agreement. Note, however, that our results based on asymptotic expansion provide at least two more significant digits. Let us mention that the analytic result for the leading order expansion term of case IV(b) agrees with the result presented in Ref.~\cite{Kataev:2012kn} which has been obtained by transforming the result of Ref.~\cite{Boughezal:2011vw} to QED. Let us next turn to the anomalous magnetic moment of the electron. The numerical values for the two- and three-loop contributions read \begin{eqnarray} A^{(4)}_{2,e}(M_e/M_\mu) &=& 5.19738668(26) \cdot 10^{-7} \,,\nonumber\\ A^{(6)}_{2,e}(M_e/M_\mu) &=& - 7.37394162(27) \cdot 10^{-6} \,,\nonumber\\ A^{(4)}_{2,e}(M_e/M_\tau) &=& 1.83798(33) \cdot 10^{-9} \,,\nonumber\\ A^{(6)}_{2,e}(M_e/M_\tau) &=& - 6.5830(11) \cdot 10^{-8} \,, \end{eqnarray} where $M_e/M_\mu=4.83633166(12) \cdot 10^{-3}$ and $M_e/M_\tau=2.87592(26) \cdot 10^{-4}$~\cite{Beringer:1900zz} have been used. Inserting these values into Eq.~(\ref{eq::A8}) leads to the following four-loop results \begin{eqnarray} A^{(8)}_{2,e}(M_e/M_\mu) &\approx& (9.161259603 + 0.000711078 + 2.2 \cdot 10^{-8}) \cdot 10^{-4}\nonumber\\ &\approx& 9.161970703(2)(372) \cdot 10^{-4} \,,\nonumber\\ A^{(8)}_{2,e}(M_e/M_\tau) &\approx& (7.42923268609971 + 2.75209424 \cdot 10^{-6} + 3.2 \cdot 10^{-13}) \cdot 10^{-6}\nonumber\\ &\approx& 7.42924(0)(118) \cdot 10^{-6} \,, \label{eq::Aetaunum} \end{eqnarray} where the uncertainty has again been estimated by 10\% of the third term in the expansion and the parameter uncertainty is displayed separately. In Ref.~\cite{Aoyama:2012wj} one finds the results\footnote{Note that the entry for $A^{(8)}_{2,e}(M_e/M_\tau)$ in Tab.~I of Ref.\cite{Aoyama:2012wj} should be multiplied by a factor $1/100$. This misprint has been confirmed by the authors of Ref.~\cite{Aoyama:2012wj}.} $A^{(8)}_{2,e}(M_e/M_\mu)=9.222(66)\times 10^{-4}$ and $A^{(8)}_{2,e}(M_e/M_\tau)=7.38(12)\times 10^{-6}$ which agree with our numerical values. As far as the growth of the coefficients is concerned we observe the same pattern as for the muon: there is about one order of magnitude between two and three loops and a factor 100 between three and four loops. Note, however, that the three-loop result is negative for $a_e$. \begin{table}[t] \begin{center} \begin{tabular}{c||l|l} group & \multicolumn{2}{c}{$10^4 \cdot A^{(8)}_{2,e}(M_e/M_\mu)$}\\ \hline & this work & \cite{Aoyama:2012wj}\\ \hline I(a) & \hphantom{$-$}0.002264474414(6) & \hphantom{$-$}0.00226456(14)\\ I(b) + I(c) + II(b) + II(c) & $-1.21390182678(6)$ & $-1.21386(24)$\\ I(d) & \hphantom{$-$}0.02472687590(2) & \hphantom{$-$}0.024725(7)\\ III & \hphantom{$-$}8.1715251555(1) & \hphantom{$-$}8.1792(95)\\ II(a) + IV(d) & $-2.6414355180(7)$ & $-2.642(12)$\\ IV(a) & \hphantom{$-$}6.3578810372(3) & \hphantom{$-$}6.3583(44)\\ IV(b) & \hphantom{$-$}0.4157367168(5) & \hphantom{$-$}0.4105(93)\\ IV(c) & $-1.954826212(2)$ & $-1.897(64)$ \end{tabular} \end{center} \caption{\label{tab::e1}Muon mass dependent corrections to $a_e$ at four-loop order as obtained in this paper and the comparison with Refs.~\cite{Aoyama:2012wj}. The uncertainties assigned to our numbers correspond to 10\% of the highest available expansion terms, i.e., the ones of order $(M_e/M_\mu)^6$ and $(M_e/M_\mu)^7$. Uncertainties from the electron and muon mass are not shown.} \end{table} \begin{table}[t] \begin{center} \begin{tabular}{c||l|l} group & \multicolumn{2}{c}{$10^6 \cdot A^{(8)}_{2,e}(M_e/M_\tau)$}\\ \hline & this work & \cite{Aoyama:2012wj}\\ \hline I(a) & \hphantom{$-$}0.0008024665425029(1) & \hphantom{$-$}0.00080233(5)\\ I(b) + I(c) + II(b) + II(c) & $-0.9458168451136621(8)$ & $-0.94506(25)$\\ I(d) & \hphantom{$-$}0.0087455060010553(1) & \hphantom{$-$}0.008744(1)\\ III & \hphantom{$-$}6.059301961911502(2) & \hphantom{$-$}6.061(12)\\ II(a) + IV(d) & $-1.372489352896281(9)$ & $-1.3835(30)$\\ IV(a) & \hphantom{$-$}4.510496216222387(2) & \hphantom{$-$}4.5117(69)\\ IV(b) & \hphantom{$-$}0.147081582099596(4) & \hphantom{$-$}0.1431(95)\\ IV(c) & $-0.97888609657284(3)$ & $-1.02(11)$ \end{tabular} \end{center} \caption{\label{tab::e2}Tau lepton mass dependent corrections to $a_e$ at four-loop order as obtained in this paper and the comparison with Refs.~\cite{Aoyama:2012wj}. The uncertainties assigned to our numbers correspond to 10\% of the highest available expansion terms, i.e., the ones of order $(M_e/M_\tau)^6$ and $(M_e/M_\tau)^7$. Uncertainties from the electron and tau lepton mass are not shown. The result of Ref.~\cite{Aoyama:2012wj} for the contribution I(d) has been multiplied by $1/100$ [see footnote after Eq.~(\ref{eq::Aetaunum})].} \end{table} In Tabs.~\ref{tab::e1} and~\ref{tab::e2} our results are shown for the individual classes of Feynman diagrams. Due to the smallness of the expansion parameters our method provides an accuracy of at least eight significant digits. The comparison with the results of Ref.~\cite{Aoyama:2012wj} demonstrates good overall agreement. Note that we have applied the methods of Refs.~\cite{Baikov:1995ui,Baikov:2012rr,Baikov:2013ula}, where four- and five-loop contributions to $a_\mu$ from polarization function insertions have been computed, to cross check our result for case I(d). The quantity $A^{(8)}_{3,e}(M_e/M_\mu,M_e/M_\tau)$ has a more complicated structure since two different heavy masses are present. However, due to the strong hierarchy $M_\tau \gg M_\mu \gg M_e$ it is possible to apply the asymptotic expansion successively which again leads to one-scale vacuum and on-shell integrals. Our final result reads \begin{eqnarray} A^{(8)}_{3,e}(M_e/M_\mu, M_e/M_\tau) &=& \frac{M_e^2}{M_\tau^2} \bigg( -\frac{3123671}{1458000}-\frac{\pi^2}{270}+\frac{\pi^4}{30}-\frac{19 \zeta_3}{45}\nonumber\\&& \qquad+\ln\frac{M_\mu^2}{M_\tau^2} \left(\frac{271073}{291600}-\frac{3 \zeta_3}{2}\right)+\frac{89}{810}\ln^2\frac{M_\mu^2}{M_\tau^2} \bigg)\nonumber\\&& +\frac{M_e^2 M_\mu}{M_\tau^3} \frac{\pi^2}{90}\nonumber\\&& +\frac{M_e^2 M_\mu^2}{M_\tau^4} \bigg(-\frac{1213316893}{5834430000}+\frac{\pi^4}{3150}+\frac{1294 \zeta_3}{3675}-\frac{3}{280}\ln^3\frac{M_\mu^2}{M_\tau^2}\nonumber\\&& \qquad+\ln\frac{M_\mu^2}{M_\tau^2} \left(-\frac{9573107}{18522000}+\frac{\zeta_3}{70}\right)+\frac{130813}{1058400}\ln^2\frac{M_\mu^2}{M_\tau^2}\bigg)\nonumber\\&& +\frac{M_e^4}{M_\mu^2 M_\tau^2} \bigg( \frac{3304933}{14580000}+\frac{88 \pi^2}{6075}-\frac{107 \zeta_3}{360}+\frac{2533}{40500}\ln\frac{M_\mu^2}{M_\tau^2}\nonumber\\&& \qquad+\ln\frac{M_e^2}{M_\mu^2} \left(-\frac{3239}{121500}-\frac{79}{1350}\ln\frac{M_\mu^2}{M_\tau^2}\right)-\frac{7}{8100}\ln^2\frac{M_e^2}{M_\mu^2} \bigg)\nonumber\\&& +\frac{M_e^4}{M_\tau^4}\bigg(-\frac{19009349146181}{10081895040000}-\frac{37877173 \zeta_3}{76204800}-\frac{79 \pi^2}{58800}\nonumber\\&& \qquad-\frac{373}{40320} P_4+\frac{280111 \pi^4}{14515200}\nonumber\\&& \qquad+\ln\frac{M_e^2}{M_\mu^2} \bigg(\frac{441068819}{1714608000}-\frac{33487}{2721600}\ln\frac{M_\mu^2}{M_\tau^2}\nonumber\\&& \qquad\qquad\qquad\quad+\frac{1423}{38880}\ln^2\frac{M_\mu^2}{M_\tau^2}-\frac{\pi^2}{420}\bigg)\nonumber\\&& \qquad+\ln\frac{M_\mu^2}{M_\tau^2} \left(\frac{767814079}{750141000}-\frac{\pi^2}{420}-\frac{61849 \zeta_3}{80640}\right)\nonumber\\&& \qquad-\frac{3034811}{38102400}\ln^2\frac{M_\mu^2}{M_\tau^2}+\frac{1181}{40824}\ln^3\frac{M_\mu^2}{M_\tau^2}\bigg)\nonumber\\&& +\frac{M_e^2 M_\mu^3}{M_\tau^5} \frac{\pi^2}{90}+\frac{M_e^4 M_\mu}{M_\tau^5} \left(\frac{79 \pi^2}{19600}+\frac{\pi^2}{140}\ln\frac{M_e^2}{M_\tau^2}\right) \,. \end{eqnarray} After inserting numerical values for the lepton masses one obtains \begin{eqnarray} A^{(8)}_{3,e}(M_e/M_\mu, M_e/M_\tau) &\approx& ( 7.4426 + 0.0261 ) \cdot 10^{-7} \nonumber\\ &\approx& 7.4687(26)(10) \cdot 10^{-7} \,, \end{eqnarray} which has to be compared with $A^{(8)}_{3,e}(M_e/M_\mu, M_e/M_\tau) = 7.465(18) \cdot 10^{-7}$ as obtained in Ref.~\cite{Aoyama:2012wj}. Again good agreement is found, however, our analytic result is more precise by about an order of magnitude. It is interesting to note that the three-loop coefficient which is given by \begin{eqnarray} A^{(6)}_{3,e}(M_e/M_\mu, M_e/M_\tau) &=& 1.90982(34) \cdot 10^{-13} \,, \end{eqnarray} is more than six orders of magnitude smaller than the four-loop one which is due to the fact that the leading term is suppressed by $M_e^4/(M_\mu^2 M_\tau^2)$ whereas at four loops the suppression factor is only $M_e^2/M_\tau^2$. Note, however, that the overall contribution is very small. Similarly to the three-loop expression also the leading term of the four-loop contribution where three one-loop heavy lepton bubbles are inserted into the photon propagator (see class I(a) in Fig.~\ref{fig::diags}) is of order ${\cal O}(M_e^2/(M_\mu^2 M_\tau^2))$. Thus we compute for this contribution also the next term of the hard-mass procedure. It is given by \begin{eqnarray} \lefteqn{\delta A^{(8)}_{3,e}(M_e/M_\mu, M_e/M_\tau)\Bigg|_{I(a),M^6} =} \nonumber\\&& \frac{M_e^4 M_\mu^2}{M_\tau^6} \bigg( -\frac{1032407}{187535250}+\frac{1303}{297675}\ln\frac{M_\mu^2}{M_\tau^2}+\frac{4}{945}\ln^2\frac{M_\mu^2}{M_\tau^2} \nonumber\\ && \qquad\qquad+\ln\frac{M_e^2}{M_\mu^2} \bigg(-\frac{1039}{297675}+\frac{4}{945}\ln\frac{M_\mu^2}{M_\tau^2}\bigg) \bigg)\nonumber\\&& +\frac{M_e^6}{M_\mu^4 M_\tau^2} \bigg( \frac{204569}{30870000}+\frac{166}{18375}\ln\frac{M_e^2}{M_\mu^2}+\frac{1}{350}\ln^2\frac{M_e^2}{M_\mu^2} \bigg)\nonumber\\&& +\frac{M_e^6}{M_\mu^2 M_\tau^4} \bigg( \frac{959}{90000}+\frac{31}{5250}\ln\frac{M_e^2}{M_\mu^2}+\frac{1}{350}\ln^2\frac{M_e^2}{M_\mu^2} \bigg)\nonumber\\&& +\frac{M_e^6}{M_\tau^6} \bigg(\ln\frac{M_e^2}{M_\mu^2} \bigg(\frac{2735573}{187535250}-\frac{199}{297675}\ln\frac{M_\mu^2}{M_\tau^2}+\frac{2}{945}\ln^2\frac{M_\mu^2}{M_\tau^2}\bigg)\nonumber\\&& \quad\qquad+\frac{8 \zeta_3}{315}-\frac{118286321}{19691201250}+\frac{676036}{31255875}\ln\frac{M_\mu^2}{M_\tau^2}\nonumber\\&& \quad\qquad+\frac{394}{99225}\ln^2\frac{M_\mu^2}{M_\tau^2}+\frac{2}{945}\ln^3\frac{M_\mu^2}{M_\tau^2} \bigg) \,. \end{eqnarray} This term is included in the numerical values shown in Table~\ref{tab::e3} where our results are compared to the ones of Ref.~\cite{Aoyama:2012wj}. The quality of the agreement is as in the previous cases. \begin{table}[t] \begin{center} \begin{tabular}{c||l|l} group & \multicolumn{2}{c}{$10^7 \cdot A^{(8)}_{3,e}(M_e/M_\mu, M_e/M_\tau)$}\\ \hline & this work & ref.\\ \hline I(a) & \hphantom{$-$}0.00001199558(2) & \hphantom{$-$}0.000011994(1)\\ I(b) + I(c) & \hphantom{$-$}0.172910(24) & \hphantom{$-$}0.172874(21)\\ II(b) + II(c) & $-1.64747(17)$ & $-1.64866(67)$\\ IV(a) & \hphantom{$-$}8.9432(25) & \hphantom{$-$}8.941(17) \end{tabular} \end{center} \caption{\label{tab::e3}Lepton mass dependent corrections to $a_e$ at four-loop order induced by diagrams which contain at the same time the muon and tau lepton. The results obtained in this paper are compared to the ones of Refs.~\cite{Aoyama:2012wj}. The uncertainties assigned to our numbers correspond to 10\% of the highest available expansion terms and uncertainties from the lepton masses are not shown. } \end{table} \section{\label{sec::concl}Conclusions} Four-loop corrections induced by a heavy lepton to the anomalous magnetic moment of the electron and the muon have been computed. This includes tau lepton contributions to $a_\mu$ and contributions with virtual muons and tau leptons to $a_e$. With the help of an asymptotic expansion in the mass ratios we obtained analytic results. Their numerical evaluation leads to full agreement with the results of Refs.~\cite{Aoyama:2012wj,Aoyama:2012wk} which have been obtained with numerical methods. However, our results are more precise. Actually, the uncertainty is of the order of or even smaller than the one originating from the imprecise knowledge of the lepton masses. Due to the decoupling of heavy particles the heavy-lepton contributions are numerically quite small. \section*{Acknowledgements} We would like to thank M.~Nio for useful communications concerning Ref.~\cite{Aoyama:2012wj}. This work was supported by the DFG through the SFB/TR~9 ``Computational Particle Physics'' and by the EU Network {\sf LHCPHENOnet} PITN-GA-2010-264564. The Feynman diagrams were drawn with {\tt JaxoDraw}~\cite{Vermaseren:1994je,Binosi:2008ig}. \begin{appendix} \section*{\label{app::CTs}Appendix: On-shell counterterms} In this appendix we provide analytic results for the on-shell counterterms for $\alpha$, $M_l$ and $\psi_l$ where the latter stands for the lepton wave function. We concentrate on the contributions relevant for our calculation, i.e., the corrections originating from closed heavy lepton loops. In the formulae below we use the notation $L_x = \ln(\mu^2/M_x^2)$ with $x=e,\mu,\tau$ and $a_4=\mbox{Li}_4(1/2)$ and mark the contributions from closed electron, muon and tau loops by the labels $n_e=1$, $n_\mu=1$ and $n_\tau=1$. In our calculation we renormalize the coupling constant in a first step in the $\overline{\rm MS}$ scheme and switch to the on-shell scheme after having obtained a finite result. The relation between the fine structure constant defined in the $\overline{\rm MS}$ scheme, $\bar\alpha(\mu)\equiv \bar\alpha$, and the corresponding on-shell quantity reads \begin{eqnarray} \frac{\bar{\alpha}}{\alpha} &=& 1 +\frac{\alpha}{\pi} \sum_{i=e,\mu,\tau} \frac{L_i n_i}{3} +\left(\frac{\alpha}{\pi}\right)^2 \left[\left(\sum_{i=e,\mu,\tau} \frac{L_i n_i}{3}\right)^2 + \sum_{i=e,\mu,\tau} \left(\frac{15}{16}+\frac{L_i}{4}\right) n_i\right] \nonumber\\&& +\left(\frac{\alpha}{\pi}\right)^3 \Bigg[\left( \sum_{i=e,\mu,\tau} \frac{L_i n_i}{3} \right)^3 +\sum_{\scriptsize \begin{tabular}{c}$i,j=e,\mu,\tau$\\$M_i<M_j$\end{tabular}} n_i n_j \Bigg(-\frac{311}{1296}-\frac{\pi^2}{18}+L_i \left(\frac{15}{16}+\frac{5 L_j}{12}\right) \nonumber\\&& \qquad\qquad+\frac{23 L_j}{144}+\frac{\pi^2}{6} \left(\frac{M_i}{M_j}\right)+\left(-\frac{167}{150}+\frac{L_i}{45}-\frac{L_j}{45}\right) \left(\frac{M_i}{M_j}\right)^2+\frac{\pi^2}{6} \left(\frac{M_i}{M_j}\right)^3\nonumber\\&& \qquad\qquad+\bigg(-\frac{23353331}{37044000}-\frac{29 L_i^2}{420}+L_i \left(-\frac{28967}{88200}+\frac{29 L_j}{210}\right)\nonumber\\&& \qquad\qquad\qquad+\frac{28967 L_j}{88200}-\frac{29 L_j^2}{420}-\frac{\pi^2}{18}\bigg) \left(\frac{M_i}{M_j}\right)^4\nonumber\\&& \qquad\qquad+\left(\frac{5288963}{62511750}+\frac{2 L_i^2}{315}+L_i \left(\frac{4609}{99225}-\frac{4 L_j}{315}\right)-\frac{4609 L_j}{99225}+\frac{2 L_j^2}{315}\right) \left(\frac{M_i}{M_j}\right)^6\Bigg)\nonumber\\&& \qquad+\sum_{i=e,\mu,\tau} n_i \left(\frac{77}{576}-\frac{L_i}{32}+\frac{5 \pi^2}{24}-\frac{\ln(2) \pi^2}{3}+\frac{\zeta_3}{192}\right)\nonumber\\&& \qquad+\sum_{i=e,\mu,\tau} n_i^2 \left(-\frac{695}{648}+\frac{79 L_i}{144}+\frac{5 L_i^2}{24}+\frac{\pi^2}{9}+\frac{7 \zeta_3}{64}\right)\Bigg] \,, \end{eqnarray} where terms up to ${\cal O}(M_i^{8}/M_j^{8})$ are included. In the case of the heavy lepton contributions to $a_e$ this formula can immediately be applied, in the case of $a_\mu$ one has to set $n_e=0$. The bare and on-shell renormalized lepton mass and wave function are related by \begin{eqnarray} M_l^{\rm bare} &=& Z_{m,l}^\text{OS} \, M_l \,,\nonumber\\ \psi_l^{\rm bare} &=& Z_{2,l}^\text{OS} \, \psi_l \,, \end{eqnarray} where the renormalization constants for the muon mass and wave function are given by \begin{eqnarray} Z_{m,\mu}^\text{OS} &=& 1 + \frac{\bar{\alpha}}{\pi} \Bigg[-1-\frac{3}{4 \epsilon}-\frac{3 L_\mu}{4}+\epsilon \left(-2-L_\mu-\frac{3 L_\mu^2}{8}-\frac{\pi^2}{16}\right)\nonumber\\&& \qquad+\epsilon^2 \left(-4-\frac{L_\mu^2}{2}-\frac{L_\mu^3}{8}-\frac{\pi^2}{12}+L_\mu \left(-2-\frac{\pi^2}{16}\right)+\frac{\zeta_3}{4}\right)\nonumber\\&& \qquad+\epsilon^3 \bigg(-8-\frac{L_\mu^3}{6}-\frac{L_\mu^4}{32}-\frac{\pi^2}{6}-\frac{3 \pi^4}{640}+L_\mu^2 \left(-1-\frac{\pi^2}{32}\right)\nonumber\\&& \qquad\qquad+L_\mu \left(-4-\frac{\pi^2}{12}+\frac{\zeta_3}{4}\right)+\frac{\zeta_3}{3}\bigg)\Bigg]\nonumber\\&& +\left(\frac{\bar{\alpha}}{\pi}\right)^2 \Bigg[\frac{1}{\epsilon^2} \left(\frac{9}{32}-\frac{n_\mu}{8}-\frac{n_\tau}{8}\right)+\frac{1}{\epsilon} \left(\frac{45}{64}+\frac{9 L_\mu}{16}+\frac{5 n_\mu}{48}+\frac{5 n_\tau}{48}\right)\nonumber\\&& \qquad+\frac{199}{128}+\frac{45 L_\mu}{32}+\frac{9 L_\mu^2}{16}-\frac{17 \pi^2}{64}+\frac{\ln(2) \pi^2}{2}-\frac{3 \zeta_3}{4}\nonumber\\&& \qquad+n_\mu \left(\frac{143}{96}+\frac{13 L_\mu}{24}+\frac{L_\mu^2}{8}-\frac{\pi^2}{6}\right)\nonumber\\&& \qquad+n_\tau \bigg(-\frac{89}{288}+\frac{13 L_\tau}{24}+\frac{L_\mu L_\tau}{4}-\frac{L_\tau^2}{8}+\left(\frac{19}{150}+\frac{L_\mu}{15}-\frac{L_\tau}{15}\right) \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad+\left(\frac{1389}{78400}+\frac{9 L_\mu}{560}-\frac{9 L_\tau}{560}\right) \left(\frac{M_\mu}{M_\tau}\right)^4+\left(\frac{997}{198450}+\frac{2 L_\mu}{315}-\frac{2 L_\tau}{315}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\nonumber\\&& \qquad+\left(\frac{1229}{627264}+\frac{5 L_\mu}{1584}-\frac{5 L_\tau}{1584}\right) \left(\frac{M_\mu}{M_\tau}\right)^8\bigg)\nonumber\\&& \qquad+\epsilon \bigg(\frac{677}{256}-12 a_4+\frac{45 L_\mu^2}{32}+\frac{3 L_\mu^3}{8}-\frac{\ln^4(2)}{2}-\frac{205 \pi^2}{128}+3 \ln(2) \pi^2\nonumber\\&& \quad\qquad-\ln^2(2) \pi^2+\frac{7 \pi^4}{40}-\frac{135 \zeta_3}{16}+L_\mu \left(\frac{199}{64}-\frac{17 \pi^2}{32}+\ln(2) \pi^2-\frac{3 \zeta_3}{2}\right)\nonumber\\&& \quad\qquad+n_\mu \left(\frac{1133}{192}+\frac{17 L_\mu^2}{24}+\frac{L_\mu^3}{8}-\frac{227 \pi^2}{288}+\ln(2) \pi^2+L_\mu \left(\frac{175}{48}-\frac{5 \pi^2}{16}\right)-\frac{7 \zeta_3}{2}\right)\nonumber\\&& \quad\qquad+n_\tau \bigg(\frac{869}{1728}+\frac{7 L_\tau}{144}+\frac{L_\mu^2 L_\tau}{8}+\frac{3 L_\tau^2}{8}-\frac{L_\tau^3}{8}+\frac{13 \pi^2}{288}+L_\mu \left(\frac{L_\tau}{3}+\frac{L_\tau^2}{8}+\frac{\pi^2}{48}\right)\nonumber\\&& \quad\qquad+\left(-\frac{701}{3375}+\frac{L_\mu^2}{30}+L_\mu \left(\frac{1}{30}+\frac{L_\tau}{15}\right)+\frac{11 L_\tau}{50}-\frac{L_\tau^2}{10}\right) \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \quad\qquad+\left(\frac{20481}{10976000}+\frac{9 L_\mu^2}{1120}+L_\mu \left(\frac{27}{1120}+\frac{9 L_\tau}{560}\right)+\frac{111 L_\tau}{9800}-\frac{27 L_\tau^2}{1120}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \quad\qquad+\left(\frac{584149}{125023500}+\frac{L_\mu^2}{315}+L_\mu \left(\frac{5}{378}+\frac{2 L_\tau}{315}\right)-\frac{631 L_\tau}{198450}-\frac{L_\tau^2}{105}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\nonumber \\ &&\qquad+\left(\frac{176625767}{60857153280}+\frac{5 L_\mu^2}{3168}+L_\mu \left(\frac{1}{126}+\frac{5 L_\tau}{1584}\right)-\frac{8821 L_\tau}{2195424}-\frac{5 L_\tau^2}{1056}\right) \left(\frac{M_\mu}{M_\tau}\right)^8\bigg)\bigg)\Bigg]\nonumber\\&& +\left(\frac{\bar{\alpha}}{\pi}\right)^3 \Bigg[\frac{1}{\epsilon^3} \left(-\frac{9}{128}+\frac{3 n_\mu}{32}-\frac{n_\mu^2}{36}-\frac{n_\mu n_\tau}{18}+\frac{3 n_\tau}{32}-\frac{n_\tau^2}{36}\right)\nonumber\\&& \qquad+\frac{1}{\epsilon^2} \bigg(-\frac{63}{256}-\frac{27 L_\mu}{128}+\left(-\frac{5}{192}+\frac{3 L_\mu}{32}\right) n_\mu+\frac{5 n_\mu^2}{216}\nonumber\\&& \qquad\qquad+\frac{5 n_\mu n_\tau}{108}+\left(-\frac{5}{192}+\frac{3 L_\mu}{32}\right) n_\tau+\frac{5 n_\tau^2}{216}\bigg)\nonumber\\&& \qquad+\frac{1}{\epsilon}\Bigg(-\frac{457}{512}-\frac{189 L_\mu}{256}-\frac{81 L_\mu^2}{256}+\frac{35 n_\mu^2}{1296}\nonumber\\&& \qquad\qquad+\frac{35 n_\mu n_\tau}{648}+\frac{35 n_\tau^2}{1296}+\frac{111 \pi^2}{512}-\frac{3 \ln(2) \pi^2}{8}+\frac{9 \zeta_3}{16}\nonumber\\&& \qquad\qquad+n_\tau \bigg(\frac{79}{128}+\frac{3 L_\mu^2}{64}+L_\mu \left(\frac{3}{64}-\frac{3 L_\tau}{16}\right)-\frac{13 L_\tau}{32}+\frac{3 L_\tau^2}{32}+\frac{\pi^2}{128}-\frac{\zeta_3}{4}\nonumber\\&& \qquad\qquad\quad+\left(-\frac{19}{200}-\frac{L_\mu}{20}+\frac{L_\tau}{20}\right) \left(\frac{M_\mu}{M_\tau}\right)^2+\left(-\frac{4167}{313600}-\frac{27 L_\mu}{2240}+\frac{27 L_\tau}{2240}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad\qquad\quad+\left(-\frac{997}{264600}-\frac{L_\mu}{210}+\frac{L_\tau}{210}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\bigg)\nonumber\\&& \qquad\qquad+n_\mu \left(-\frac{281}{384}-\frac{23 L_\mu}{64}-\frac{3 L_\mu^2}{64}+\frac{17 \pi^2}{128}-\frac{\zeta_3}{4}\right)\Bigg)\nonumber\\&& \qquad-\frac{14225}{3072}-3 a_4-\frac{567 L_\mu^2}{512}-\frac{81 L_\mu^3}{256}-\frac{\ln^4(2)}{8}-\frac{6037 \pi^2}{3072}+5 \ln(2) \pi^2+\frac{5 \ln^2(2) \pi^2}{4}\nonumber\\&& \qquad-\frac{73 \pi^4}{480}+\frac{5 \zeta_5}{8}+\frac{153 \zeta_3}{128}-\frac{\pi^2 \zeta_3}{16}+L_\mu \left(-\frac{1371}{512}+\frac{333 \pi^2}{512}-\frac{9 \ln(2) \pi^2}{8}+\frac{27 \zeta_3}{16}\right)\nonumber\\&& \qquad+n_\tau \bigg(\frac{6367}{2304}-4 a_4+\frac{L_\mu^3}{64}+L_\mu^2 \left(\frac{3}{128}-\frac{9 L_\tau}{32}\right)-\frac{9 L_\tau^2}{32}+\frac{3 L_\tau^3}{32}\nonumber\\&& \qquad\qquad-\frac{\ln^4(2)}{6}-\frac{23 \pi^2}{768}+\frac{\ln^2(2) \pi^2}{6}+\frac{11 \pi^4}{360}-\frac{29 \zeta_3}{16}\nonumber\\&& \qquad\qquad+L_\mu \left(\frac{415}{384}-\frac{21 L_\tau}{32}-\frac{\pi^2}{128}\right)+L_\tau \left(-\frac{1}{64}+\frac{5 \pi^2}{24}-\frac{\ln(2) \pi^2}{3}-\frac{\zeta_3}{4}\right)\nonumber\\&& \qquad\qquad+\left(\frac{M_\mu}{M_\tau}\right)^2 \bigg(\frac{8153}{12150}-\frac{31 L_\mu^2}{360}+L_\mu \left(-\frac{67}{4050}+\frac{L_\tau}{45}\right)\nonumber\\&& \qquad\qquad\qquad\qquad\qquad-\frac{4349 L_\tau}{16200}+\frac{23 L_\tau^2}{360}+\frac{2 \pi^2}{135}-\frac{77 \zeta_3}{144}\bigg)\nonumber \\ &&\qquad+\left(\frac{M_\mu}{M_\tau}\right)^4 \bigg(\frac{13231711}{98784000}-\frac{17 L_\mu^2}{1120}+L_\mu \left(\frac{1907}{44800}-\frac{13 L_\tau}{2240}\right)\nonumber\\&& \qquad\qquad\qquad\qquad-\frac{517 L_\tau}{6272}+\frac{47 L_\tau^2}{2240}+\frac{\pi^2}{105}-\frac{147 \zeta_3}{1024}\bigg)\nonumber\\&& \qquad+\left(\frac{M_\mu}{M_\tau}\right)^6 \bigg(\frac{3752184623}{90016920000}-\frac{8 L_\mu^2}{2025}+L_\mu \left(\frac{925261}{35721000}-\frac{181 L_\tau}{28350}\right)\nonumber\\&& \qquad\qquad\qquad\qquad-\frac{664523 L_\tau}{17860500}+\frac{293 L_\tau^2}{28350}+\frac{32 \pi^2}{6075}-\frac{119 \zeta_3}{1920}\bigg)\bigg)\nonumber\\&& +n_\tau^2 \bigg(\frac{1685}{7776}+\frac{31 L_\tau}{108}-\frac{13 L_\tau^2}{72}-\frac{L_\mu L_\tau^2}{12}+\frac{L_\tau^3}{18}-\frac{7 \zeta_3}{18}\nonumber\\&& \qquad+\left(\frac{23}{324}-\frac{19 L_\tau}{225}-\frac{2 L_\mu L_\tau}{45}+\frac{2 L_\tau^2}{45}\right) \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad+\left(-\frac{119}{24000}+L_\mu \left(-\frac{1}{100}-\frac{3 L_\tau}{280}\right)-\frac{71 L_\tau}{39200}+\frac{3 L_\tau^2}{280}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad+\left(-\frac{1594}{496125}+L_\mu \left(-\frac{1}{175}-\frac{4 L_\tau}{945}\right)+\frac{704 L_\tau}{297675}+\frac{4 L_\tau^2}{945}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\bigg)\nonumber\\&& +n_\mu n_\tau \bigg(-\frac{1327}{3888}-\frac{13 L_\mu L_\tau}{36}-\frac{L_\mu^2 L_\tau}{12}+\frac{L_\tau^3}{36}+L_\tau \left(-\frac{5}{8}+\frac{\pi^2}{9}\right)+\frac{2 \zeta_3}{9}\nonumber\\&& \qquad+\left(-\frac{1541}{3375}-\frac{11 L_\mu}{45}-\frac{L_\mu^2}{45}+\frac{4 L_\tau}{25}+\frac{L_\tau^2}{45}+\frac{4 \pi^2}{135}\right) \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad+\left(-\frac{1833259}{16464000}-\frac{9 L_\mu^2}{560}+L_\mu \left(-\frac{3551}{39200}+\frac{3 L_\tau}{140}\right)+\frac{193 L_\tau}{2450}-\frac{3 L_\tau^2}{560}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad+\left(-\frac{1997398}{93767625}-\frac{2 L_\mu^2}{315}+L_\mu \left(-\frac{8021}{297675}+\frac{8 L_\tau}{945}\right)+\frac{7024 L_\tau}{297675}-\frac{2 L_\tau^2}{945}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\bigg)\nonumber\\&& +n_\mu \bigg(-\frac{5257}{2304}+\frac{8 a_4}{3}-\frac{117 L_\mu^2}{128}-\frac{11 L_\mu^3}{64}+\frac{\ln^4(2)}{9}-\frac{1327 \pi^2}{6912}+\frac{37 \zeta_3}{96}\nonumber\\&& \qquad+\frac{5 \ln(2) \pi^2}{36}-\frac{\ln^2(2) \pi^2}{9}+\frac{91 \pi^4}{2160}+L_\mu \left(-\frac{1145}{384}+\frac{221 \pi^2}{384}-\frac{\ln(2) \pi^2}{3}-\frac{\zeta_3}{4}\right)\bigg)\nonumber\\&& +n_\mu^2 \left(-\frac{9481}{7776}-\frac{13 L_\mu^2}{72}-\frac{L_\mu^3}{36}+\frac{4 \pi^2}{135}+L_\mu \left(-\frac{197}{216}+\frac{\pi^2}{9}\right)+\frac{11 \zeta_3}{18}\right)\Bigg] \,, \end{eqnarray} \begin{eqnarray} Z_{2,\mu}^\text{OS} &=& 1 + \frac{\bar{\alpha}}{\pi} \Bigg[-1-\frac{3}{4 \epsilon}-\frac{3 L_\mu}{4}+\epsilon \left(-2-L_\mu-\frac{3 L_\mu^2}{8}-\frac{\pi^2}{16}\right)\nonumber\\&& \qquad+\epsilon^2 \left(-4-\frac{L_\mu^2}{2}-\frac{L_\mu^3}{8}-\frac{\pi^2}{12}+L_\mu \left(-2-\frac{\pi^2}{16}\right)+\frac{\zeta_3}{4}\right)\nonumber\\&& \qquad+\epsilon^3 \bigg(-8-\frac{L_\mu^3}{6}-\frac{L_\mu^4}{32}-\frac{\pi^2}{6}-\frac{3 \pi^4}{640}+L_\mu^2 \left(-1-\frac{\pi^2}{32}\right)\nonumber\\&& \qquad\qquad+L_\mu \left(-4-\frac{\pi^2}{12}+\frac{\zeta_3}{4}\right)+\frac{\zeta_3}{3}\bigg)\Bigg]\nonumber\\&& +\left(\frac{\bar{\alpha}}{\pi}\right)^2 \Bigg[\frac{9}{32 \epsilon^2}+\frac{1}{\epsilon} \left(\frac{51}{64}+\frac{9 L_\mu}{16}+\left(\frac{1}{16}+\frac{L_\mu}{4}\right) n_\mu+\left(\frac{1}{16}+\frac{L_\tau}{4}\right) n_\tau\right)\nonumber\\&& \qquad+\frac{433}{128}+\frac{51 L_\mu}{32}+\frac{9 L_\mu^2}{16}-\frac{49 \pi^2}{64}+\ln(2) \pi^2-\frac{3 \zeta_3}{2}\nonumber\\&& \qquad+n_\mu \left(\frac{947}{288}+\frac{11 L_\mu}{24}+\frac{3 L_\mu^2}{8}-\frac{5 \pi^2}{16}\right)\nonumber\\&& \qquad+n_\tau \bigg(\frac{\pi^2}{48}-\frac{5}{96}+\frac{11 L_\tau}{24}+\frac{L_\mu L_\tau}{4}+\frac{L_\tau^2}{8}+\frac{1}{15}\left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad\qquad+\left(-\frac{129}{78400}-\frac{9 L_\mu}{560}+\frac{9 L_\tau}{560}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad\qquad+\left(-\frac{367}{99225}-\frac{4 L_\mu}{315}+\frac{4 L_\tau}{315}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\nonumber\\&& \qquad\qquad+\left(-\frac{569}{209088}-\frac{5 L_\mu}{528}+\frac{5 L_\tau}{528}\right) \left(\frac{M_\mu}{M_\tau}\right)^8\bigg)\nonumber\\&& \qquad+\epsilon \bigg(\frac{211}{256}+\frac{51 L_\mu^2}{32}+\frac{3 L_\mu^3}{8}-\ln^4(2)-\frac{339 \pi^2}{128}+\frac{23 \ln(2) \pi^2}{4}-2 \ln^2(2) \pi^2\nonumber\\&& \qquad\qquad+L_\mu \left(\frac{433}{64}-\frac{49 \pi^2}{32}+2 \ln(2) \pi^2-3 \zeta_3\right)-\frac{297 \zeta_3}{16}+\frac{7 \pi^4}{20}-24 a_4\nonumber\\&& \qquad\qquad+n_\mu \bigg(\frac{17971}{1728}+\frac{5 L_\mu^2}{8}+\frac{7 L_\mu^3}{24}-\frac{445 \pi^2}{288}+2 \ln(2) \pi^2\nonumber\\&& \qquad\qquad\qquad\quad+L_\mu \left(\frac{1043}{144}-\frac{29 \pi^2}{48}\right)-\frac{85 \zeta_3}{12}\bigg)\nonumber\\&& \qquad\qquad+n_\tau \bigg(\frac{89}{576}+\frac{L_\mu^2 L_\tau}{8}+\frac{7 L_\tau^2}{24}+\frac{L_\tau^3}{24}+\frac{11 \pi^2}{288}-\frac{\zeta_3}{12}+L_\tau \left(\frac{9}{16}+\frac{\pi^2}{24}\right)\nonumber \\ &&\qquad+L_\mu \left(\frac{L_\tau}{3}+\frac{L_\tau^2}{8}+\frac{\pi^2}{48}\right)+\left(-\frac{7}{75}+\frac{2 L_\tau}{15}\right) \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad+\left(\frac{49659}{10976000}-\frac{9 L_\mu^2}{1120}+L_\mu \left(-\frac{27}{1120}-\frac{9 L_\tau}{560}\right)+\frac{51 L_\tau}{2450}+\frac{27 L_\tau^2}{1120}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad+\left(-\frac{71329}{62511750}-\frac{2 L_\mu^2}{315}+L_\mu \left(-\frac{5}{189}-\frac{4 L_\tau}{315}\right)+\frac{1891 L_\tau}{99225}+\frac{2 L_\tau^2}{105}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\nonumber\\&& \qquad+\bigg(-\frac{55373867}{20285717760}-\frac{5 L_\mu^2}{1056}+L_\mu \left(-\frac{1}{42}-\frac{5 L_\tau}{528}\right)\nonumber\\&& \qquad\qquad+\frac{13441 L_\tau}{731808}+\frac{5 L_\tau^2}{352}\bigg) \left(\frac{M_\mu}{M_\tau}\right)^8\bigg)\bigg)\Bigg]\nonumber\\&& +\left(\frac{\bar{\alpha}}{\pi}\right)^3 \Bigg[-\frac{9}{128 \epsilon^3}+\frac{1}{\epsilon^2} \bigg(-\frac{81}{256}-\frac{27 L_\mu}{128}+\left(-\frac{7}{192}-\frac{3 L_\mu}{16}\right) n_\mu+\frac{n_\mu^2}{72}\nonumber\\&& \qquad\qquad\qquad\qquad\qquad+\frac{n_\mu n_\tau}{36}+\left(-\frac{7}{192}-\frac{3 L_\tau}{16}\right) n_\tau+\frac{n_\tau^2}{72}\bigg)\nonumber\\&& \qquad+\frac{1}{\epsilon}\bigg(-\frac{1039}{512}-\frac{243 L_\mu}{256}-\frac{81 L_\mu^2}{256}+\frac{303 \pi^2}{512}-\frac{3 \ln(2) \pi^2}{4}+\frac{9 \zeta_3}{8}\nonumber\\&& \qquad\qquad+\left(-\frac{5}{432}-\frac{L_\mu^2}{12}\right) n_\mu^2+\left(-\frac{5}{216}-\frac{L_\mu L_\tau}{6}\right) n_\mu n_\tau+\left(-\frac{5}{432}-\frac{L_\tau^2}{12}\right) n_\tau^2\nonumber\\&& \qquad\qquad+n_\tau \bigg(\frac{85}{128}+L_\mu \left(-\frac{3}{64}-\frac{3 L_\tau}{8}\right)-\frac{13 L_\tau}{32}-\frac{3 L_\tau^2}{32}-\frac{\pi^2}{64}\nonumber\\&& \qquad\qquad\qquad-\frac{1}{20} \left(\frac{M_\mu}{M_\tau}\right)^2+\left(\frac{387}{313600}+\frac{27 L_\mu}{2240}-\frac{27 L_\tau}{2240}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad\qquad\qquad+\left(\frac{367}{132300}+\frac{L_\mu}{105}-\frac{L_\tau}{105}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\bigg)\nonumber\\&& \qquad\qquad+n_\mu \left(-\frac{707}{384}-\frac{29 L_\mu}{64}-\frac{15 L_\mu^2}{32}+\frac{15 \pi^2}{64}\right)\bigg)\nonumber\\&& \qquad-\frac{10823}{3072}-\frac{729 L_\mu^2}{512}-\frac{81 L_\mu^3}{256}-\frac{5 \ln^4(2)}{12}-\frac{58321 \pi^2}{9216}-10 a_4\nonumber\\&& \qquad+\frac{685 \ln(2) \pi^2}{48}+3 \ln^2(2) \pi^2-\frac{41 \pi^4}{120}-\frac{739 \zeta_3}{128}+\frac{\pi^2 \zeta_3}{8}\nonumber\\&& \qquad+L_\mu \left(-\frac{3117}{512}+\frac{909 \pi^2}{512}-\frac{9 \ln(2) \pi^2}{4}+\frac{27 \zeta_3}{8}\right)-\frac{5 \zeta_5}{16}\nonumber \\ &&+n_\tau^2 \bigg(-\frac{35}{2592}-\frac{11 L_\tau^2}{72}-\frac{L_\mu L_\tau^2}{12}-\frac{L_\tau^3}{12}-\frac{L_\tau \pi^2}{72}-\frac{2 L_\tau}{45} \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad+\left(-\frac{121}{24000}+L_\mu \left(\frac{1}{100}+\frac{3 L_\tau}{280}\right)-\frac{349 L_\tau}{39200}-\frac{3 L_\tau^2}{280}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad+\left(\frac{353}{496125}+L_\mu \left(\frac{2}{175}+\frac{8 L_\tau}{945}\right)-\frac{2668 L_\tau}{297675}-\frac{8 L_\tau^2}{945}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\bigg)\nonumber\\&& +n_\mu n_\tau \bigg(-\frac{35}{1296}-\frac{L_\mu^2 L_\tau}{4}+L_\tau \left(-\frac{481}{216}+\frac{5 \pi^2}{24}\right)\nonumber\\&& \qquad+L_\mu \left(-\frac{11 L_\tau}{36}-\frac{L_\tau^2}{12}-\frac{\pi^2}{72}\right)+\left(-\frac{37}{90}-\frac{2 L_\mu}{45}+\frac{2 \pi^2}{45}\right) \left(\frac{M_\mu}{M_\tau}\right)^2\nonumber\\&& \qquad+\left(-\frac{317689}{16464000}-\frac{3 L_\mu}{1120}+\frac{3 L_\mu^2}{560}+\frac{37 L_\tau}{9800}-\frac{3 L_\tau^2}{560}\right) \left(\frac{M_\mu}{M_\tau}\right)^4\nonumber\\&& \qquad+\left(-\frac{159337}{18753525}+\frac{151 L_\mu}{14175}+\frac{4 L_\mu^2}{945}-\frac{2437 L_\tau}{297675}-\frac{4 L_\tau^2}{945}\right) \left(\frac{M_\mu}{M_\tau}\right)^6\bigg)\nonumber\\&& +n_\mu^2 \left(-\frac{8425}{2592}-\frac{11 L_\mu^2}{72}-\frac{L_\mu^3}{6}+\frac{2 \pi^2}{45}+L_\mu \left(-\frac{481}{216}+\frac{5 \pi^2}{24}\right)+\frac{7 \zeta_3}{3}\right)\nonumber\\&& +n_\tau \bigg(\frac{4285}{2304}+L_\mu^2 \left(-\frac{3}{128}-\frac{3 L_\tau}{8}\right)-\frac{L_\tau^2}{8}-\frac{L_\tau^3}{32}-\frac{29 \pi^2}{768}+\frac{3 \zeta_3}{16}\nonumber\\&& \qquad+L_\mu \left(\frac{87}{128}-\frac{31 L_\tau}{32}-\frac{3 L_\tau^2}{16}-\frac{\pi^2}{32}\right)+L_\tau \left(-\frac{133}{192}+\frac{95 \pi^2}{192}-\frac{2 \ln(2) \pi^2}{3}+\zeta_3\right)\nonumber\\&& \qquad+\left(\frac{M_\mu}{M_\tau}\right)^2 \left(-\frac{461}{1296}-\frac{71 L_\mu}{360}+\frac{17 L_\tau}{360}+\frac{\pi^2}{45}\right)\nonumber\\&& \qquad+\left(\frac{M_\mu}{M_\tau}\right)^4 \bigg(-\frac{2229623}{8232000}+\frac{3 L_\mu^2}{80}+L_\mu \left(-\frac{28607}{940800}-\frac{87 L_\tau}{2240}\right)\nonumber\\&& \qquad\qquad\qquad\qquad+\frac{3209 L_\tau}{94080}+\frac{3 L_\tau^2}{2240}+\frac{\pi^2}{105}+\frac{147 \zeta_3}{1024}\bigg)\nonumber\\&& \qquad+\left(\frac{M_\mu}{M_\tau}\right)^6 \bigg(-\frac{15665719421}{90016920000}+\frac{1913 L_\mu^2}{56700}+L_\mu \left(-\frac{1587179}{71442000}-\frac{1103 L_\tau}{28350}\right)\nonumber\\&& \qquad\qquad\qquad\qquad+\frac{2181719 L_\tau}{71442000}+\frac{293 L_\tau^2}{56700}+\frac{16 \pi^2}{6075}+\frac{119 \zeta_3}{960}\bigg)\bigg)\nonumber\\&& +n_\mu \bigg(-\frac{76897}{6912}+12 a_4-\frac{143 L_\mu^2}{128}-\frac{19 L_\mu^3}{32}+\frac{\ln^4(2)}{2}-\frac{11551 \pi^2}{20736}+\frac{7 \ln(2) \pi^2}{18}\nonumber\\&& \qquad-\frac{\ln^2(2) \pi^2}{2}+\frac{31 \pi^4}{720}+\frac{1763 \zeta_3}{288}+L_\mu \left(-\frac{2891}{384}+\frac{233 \pi^2}{192}-\frac{2 \ln(2) \pi^2}{3}+\zeta_3\right)\bigg)\Bigg] \,, \nonumber\\ \end{eqnarray} {where we include terms up to ${\cal O}(1/M_\tau^{8})$.} The mass and wave function renormalization constants for the electron can be constructed from the above results by replacing $M_\mu$ by $M_e$, $L_\mu$ by $L_e$ and $n_\mu$ by $n_e$. Moreover the terms proportional to $n_\tau$ and $n_\tau^2$ have to be duplicated and afterwards the replacements $n_\tau\to n_\mu$, $M_\tau \to M_\mu$ and $L_\tau \to L_\mu$ have to be performed in one of the expressions. Furthermore, one has to add the contributions involving simultaneously virtual muon and tau loops which are given by \begin{eqnarray} \delta Z_{m,e}^\text{OS} &=&\left(\frac{\bar{\alpha}}{\pi}\right)^3 \bigg[-\frac{1}{18 \epsilon^3}+\frac{5}{108 \epsilon^2}+\frac{35}{648 \epsilon}-\frac{1327}{3888}+\frac{2 \zeta_3}{9}\nonumber\\&& \quad+\frac{31 L_\tau}{54}-\frac{13 L_\mu L_\tau}{36}-\frac{L_e L_\mu L_\tau}{6}+\frac{L_\mu^2 L_\tau}{12}+\frac{L_\tau^3}{36}\nonumber\\&& \quad+\frac{M_\mu^2}{M_\tau^2} \left( -\frac{47}{150}-\frac{L_\mu}{5}+\frac{L_\tau}{5} \right)+\frac{M_e^2}{M_\mu^2} \left( -\frac{19 L_\tau}{225}-\frac{2 L_e L_\tau}{45}+\frac{2 L_\mu L_\tau}{45} \right)\nonumber\\&& \quad+\frac{M_e^2}{M_\tau^2} \left( \frac{1937}{10125}-\frac{2 L_\mu}{45}-\frac{2 L_e L_\mu}{45}+\frac{L_\mu^2}{45}-\frac{L_\tau}{25}+\frac{L_\tau^2}{45} \right)\nonumber\\&& \quad+\frac{M_e^2 M_\mu^2}{M_\tau^4} \left( -\frac{107}{3675}-\frac{L_\mu}{35}+\frac{L_\tau}{35} \right)+\frac{M_e^4}{M_\mu^2 M_\tau^2} \left( -\frac{33}{1000}-\frac{L_e}{50}+\frac{L_\mu}{50} \right)\nonumber\\&& \quad+\frac{M_\mu^4}{M_\tau^4} \left(-\frac{224261}{2058000}-\frac{529 L_\mu}{9800}-\frac{3 L_\mu^2}{280}+\frac{529 L_\tau}{9800}+\frac{3 L_\mu L_\tau}{140}-\frac{3 L_\tau^2}{280}\right)\nonumber\\&& \quad+\frac{M_e^4}{M_\mu^4} \left( -\frac{463 L_\tau}{39200}-\frac{3 L_e L_\tau}{280}+\frac{3 L_\mu L_\tau}{280} \right)\nonumber\\&& \quad+\frac{M_e^4}{M_\tau^4} \left( \frac{39379}{1029000}-\frac{9 L_\mu}{1120}-\frac{3 L_e L_\mu}{280}+\frac{3 L_\mu^2}{560}-\frac{37 L_\tau}{9800}+\frac{3 L_\tau^2}{560} \right) \bigg] \,, \end{eqnarray} \begin{eqnarray} \delta Z_{2,e}^\text{OS} &=&\left(\frac{\bar{\alpha}}{\pi}\right)^3 \bigg[ \frac{1}{36 \epsilon^2}+\frac{1}{\epsilon} \left(-\frac{5}{216}-\frac{L_\mu L_\tau}{6}\right)-\frac{35}{1296}-\frac{11 L_\mu L_\tau}{36}\nonumber\\&& \quad-\frac{L_e L_\mu L_\tau}{6}-\frac{L_\mu^2 L_\tau}{12}-\frac{L_\mu L_\tau^2}{12}-\frac{L_\mu \pi^2}{72}-\frac{L_\tau \pi^2}{72}\nonumber\\&& \quad-\frac{2 L_\tau}{45} \frac{M_e^2}{M_\mu^2}-\frac{2 L_\mu}{45} \frac{M_e^2}{M_\tau^2}+\left( \frac{13}{1000}+\frac{L_e}{50}-\frac{L_\mu}{50} \right)\frac{M_e^4}{M_\mu^2 M_\tau^2}\nonumber\\&& \quad+\frac{M_e^4}{M_\mu^4} \left( \frac{43 L_\tau}{39200}+\frac{3 L_e L_\tau}{280}-\frac{3 L_\mu L_\tau}{280} \right)\nonumber\\&& \quad+\frac{M_e^4}{M_\tau^4} \left( -\frac{39379}{1029000}-\frac{3 L_\mu}{1120}+\frac{3 L_e L_\mu}{280}-\frac{3 L_\mu^2}{560}+\frac{37 L_\tau}{9800}-\frac{3 L_\tau^2}{560} \right) \bigg] \,. \end{eqnarray} These formulae include only terms up to quartic order in the inverse heavy mass since the corresponding contributions to $a_e$ are only computed up to this order. \end{appendix}
\section*{Introduction} One of the most essential problems in graphene physics is a problem of electronic transport. Electronic excitations in graphene strongly interact with each other. The strength of the interaction is controlled by the substrate dielectric permittivity so the strongest interaction is in suspended graphene. The conductivity of suspended graphene has been an open problem for many years. Many theoretical models \cite{Son:07:1, Khvech} have predicted the existence of the mass gap in the case of free graphene sheet which appears due to spontaneous breaking of sublattice symmetry. This theoretical prediction have been verified numerically within the framework of effective field theory of graphene \cite{Lahde, Hands, Buividovich:2012uk}. Resent experimental study \cite{Mayorov} of suspended graphene have shown that this material remains in metallic state even without any substrate. This discrepancy was solved in the paper \cite{arxiv}. It was shown that if one takes into the account the screening of the Coulomb potential at short distances, the tight-binding model of graphene predicts the semimetal state for suspended graphene. There are several order parameters and possible fermion condensates which were discussed in the context of phase transition phenomenon in graphene. For example, in the papers\cite{Buividovich:2012uk,arxiv,Rebbi, Buividovich:2012nx} the effect of spontaneous polarization of spins was studied. From this point of view, the dielectric phase is an antiferromagnetic state. The another possibility is excitonic condensation broadly discussed both in theoretical papers and lattice simulations by means of effective graphene field model. From microscopic point of view, two sublattices of the hexagonal lattice acquire opposite charges in this phase. It is also important to emphasize that the study of this excitonic order parameter within the framework of Hybrid Monte-Carlo simulations (like it was done in \cite{arxiv} and \cite{ Buividovich:2012nx} for the antiferromagnetic order parameter) is impossible now due to the sign problem in fermionic determinant. So in this paper we have used simplified Monte Carlo method formulated in terms of occupation numbers where this problem doesn't arise. \section{Graphene model in terms of occupation numbers} Let us consider an ideal monolayer graphene. Carbon atoms in it form two dimensional hexagonal lattice. There is a 4-valency carbon atom in every site of this lattice. Three electrons of each carbon atom participate in chemical bonding ($\sigma$-bound) with the neighboring atoms and the fourth electron ($\pi$-orbital) causes the conductivity of graphene. We use the following Hamiltonian to describe the electric properties of this material \cite{Semenoff1}: \begin{equation} \label{fullH} H = H_\kappa+H_H+H_C =-\kappa\sum_{x,\rho_b,\sigma}\left(\psi^\dagger_{\sigma,x+\rho_b} \psi_{\sigma, x} + h.c.\right) + { 1 \over 2} \sum_{x} q_x~V_{x, x}~q_x + { 1 \over 2} \sum_{x\neq x'} q_x V_{x, x'} q_{x'} \, . \end{equation} Different terms of Hamiltonian correspond to different physical phenomena. $H_\kappa$ corresponds to hoppings from one site to the nearest neighboring ones. $\psi_{\sigma,x}^\dagger$ and $\psi_{\sigma,x}$ are the creation and annihilation operators for the electrons at the site $x$ and up or down spin, $\rho_b, \,\, b=1,2,3$ are the vectors between the site $x$ and three its neighbors. The term $H_H$ is the on-site Coulomb interaction, $q_x$ is the operator of electrical charge for the site $x$: \begin{equation} q_x = 1-\psi^\dagger_{\uparrow,x} \psi_{\uparrow,x} - \psi^\dagger_{\downarrow,x} \psi_{\downarrow,x}. \label{density}\end{equation} ``One'' in this formula represents the fact that the charge of the lattice site is equal to 1 in the absence of electrons on $\pi$-orbitals. Finally, the last term in (1) is a Coulomb interaction of electrons at different sites of the lattice. $V_{x, x'}$ is a Coulomb potential screened by $\sigma$-orbitals at short distances. We will discuss this phenomenon more properly in the next part of the paper. It's convenient to introduce electron and hole excitations. Let us consider the creation and annihilation operators for the ``electrons'' ($a$ operators) and ``holes'' ($b$ operators) at the site $x$: \begin{equation} a^\dagger_x = \psi^\dagger_{x,\uparrow} \,\, , \,\,\,\, a_x = \psi_{x,\uparrow} \,\, , \,\,\,\, b^\dagger_x =\psi_{x,\downarrow} \,\, , \,\,\,\, b_x =\psi^\dagger_{x,\downarrow} \,\, . \label{ab} \end{equation} The charge operator in terms of operators $a_x$ and $b_x$ can be expressed as: \begin{equation} q_x=1-a^\dagger_x a_x - b_x\ b^\dagger_x = b^\dagger_x b_x - a^\dagger_x a_x. \end{equation} There are four different states for every site $x$: 1) The state $\vert \cdot\,\cdot \rangle$. There is no electrons on the $\pi$-orbital. 2) The states $\vert \uparrow\,\cdot \rangle$ and $\vert \downarrow \cdot \rangle$. There is only one electron on the $\pi$-orbital. 3) The state $\vert \uparrow\,\downarrow \rangle$. There are two electrons on the $\pi$-orbital. Each state in this set is also an eigenvector of ``electrons'' number operator and ``holes'' number operator at the site $x$: \begin{equation} a^\dagger_x a_x \vert n_x, \, m_x \rangle = m_x \vert n_x, \, m_x \rangle; \,\,\,\,\,\,\,\,\,\,\, b^\dagger_x b_x \vert n_x, \, m_x \rangle = n_x \vert n_x, \, m_x \rangle \, ;\,\,\,\,\,\,\,\,\,\,\, n_x\,,\,m_x =0,1.\label{c} \end{equation} Let us choose ``vacuum'' state as $\vert \downarrow\,\cdot \rangle$ with $n_x=0, \, m_x=0$ (there is no difference for Monte Carlo method what state is called as ``vacuum''). Then $\vert \uparrow\,\downarrow \rangle$ with $n_x=1, \, m_x=0$ is ``electron'' excitation, $\vert \uparrow\,\cdot \rangle$ with $n_x=1, \, m_x=1$ is ``electron+hole'' state, and $\vert\cdot\,\cdot \rangle$ with $n_x=0, \, m_x=1$ is ``hole'' excitation. These vectors are eigenstates of charge operator with eigenvalues $q_x=n_x-m_x$. As seen from the above, there are two states with the same charges (equal to ``0''). One of them is the ``vacuum'' state $\vert \downarrow \, \cdot \rangle$, and the other is the ``electron+hole'' state $\vert \uparrow \, \cdot \rangle$. There is also one state with charge ``-1'' ($ \vert \downarrow \, \uparrow \rangle$) and one state with charge ``+1'' ($\vert \cdot \, \cdot \rangle$). Now we can define the state of the whole system $ \vert S \rangle $ as an antisymmetrized product of one-electron states: $ \vert S \rangle=\vert \{n_x\}\{ m_x\}\rangle$, where $\{n_x\}$ and $\{ m_x\}$ are the distributions of occupation numbers for ''electrons'' and '' holes'' respectively at the whole lattice. Let us discuss the meaning of three different parts of the full Hamiltonian (\ref{fullH}). The values of electron-electron interaction potentials for suspended graphene one can see in \cite{Kac}. The energy scale in hopping part of the Hamiltonian is defined by the value of $\kappa=2.7$ eV. The energy scale of interaction part is defined by the Coulomb potential which is 9.3 eV in the case of on-site interaction. It is 3 times larger than the hopping parameter $\kappa$. The eigenstates of particle number operator are also eigenstates of Hubbard and Coulomb terms of Hamiltonian (\ref{fullH}) but not of the $H_\kappa$ term. Because of this fact and the fact that interaction part of the Hamiltonian gives possibly the main contribution to the energy, we will neglect $H_\kappa$. We can improve this method further by perturbation theory if necessary. So the partition function has the form: \begin{eqnarray} Z=\sum_{\{n\}\{m\}}e^{-\frac{1}{T}H(\{n\}\{m\})} = \sum_{\{n\}\{m\}}\exp\left\{ -\frac{1}{2 T}\left( \sum_x V_{xx}(n_x-m_x)^2+ \sum_{x\neq y}V_{xy}(n_x-m_x)(n_y-m_y)\right)\right\} \nonumber \end{eqnarray} The average value of an observable quantity can be easily calculated if the eigenstates of particles number operators are also eigenstates of the operator of this observable. In this case, its average value can be represented as: \begin{eqnarray} \left< O \right> = \frac{ \sum_{\{q_x\}} O(\{q_x\}) e^{-\beta H} \prod_x (1+\delta_{q_x,0})}{ \sum_{\{q_x\}} e^{-\beta H}} ,\nonumber \\ H = {1 \over 2 } \left( \sum_x V_{xx} (q_x)^2 + \sum_{x\neq y} V_{xy}(q_x)(q_y) \right), \label{aver} \end{eqnarray} where $\beta = 1/T$, and $(1+\delta_{q_x,0})$ corresponds to the existence of two different states with $q_x=0$. \section{Simple model: we take into account only interactions in the 1st coordination radius} Let us consider the model with the following partition function \begin{equation} Z = \sum_{\{q_x\}} \prod_x (1+\delta_{q_x,0}) \exp^{-\beta H(\{q_x\})}, \, \, \, \, \mbox{where} \, \, \, \, H(\{q_x\}) = {1 \over 2} \left( V_{00} \sum_{x} q^2_x + V_{01} \sum_{x, \rho_b} q_x q_{x+\rho_b} \right). \label{simp_ham} \end{equation} Only interaction with neighbouring sites is included into the Hamiltonian. The first term in Hamiltonian corresponds to the interaction between electrons at one lattice site. The second term corresponds to the Coulomb interaction between the nearest neighbours. The vectors $\rho_b$, $b=1,2,3$ connect the nearest sites with each other. The relation between $V_{00}$ and $V_{01}$ is the key quantity which defines the behaviour of the model. Let us consider the system at large $\beta$, when the ground state plays the main role in the partition function. We will consider firstly the limit $V_{00} \gg V_{01}$. In this case the self energy contribution, which corresponds to the on-site interaction term, dominates the interaction between neighbouring sites. The configuration with zero charges $q_x=0$ at all sites is the ground state of the system. In the opposite case $V_{00} \ll V_{01}$ the second term in the Hamiltonian (\ref{simp_ham})will dominate the first term and the ground state is the configuration where sites at different sublattices acquire opposite charges: $q_x = - q_{x+\rho_b}$. This situation is illustrated at the figure \ref{fig1}, where different charges are marked with different colors. Every term $q_x q_{x+\rho_b}$ gives a negative contribution into the energy. This contributions leads to the negative energy of such ''chiral domain'' in the limit $V_{01} \gg V_{00}$ and this configuration becomes a ground state because the zero charges configuration has zero energy. So, there is a critical value of $V^c_{00}$ where one ground state is replaced by another one. \begin{figure} \centering \includegraphics[width=.3\textwidth]{fig1_eps.eps} \caption{The system with charge separation.} \label{fig1} \end{figure} The value of $V^c_{00}$ at large $\beta$ can be found from the following considerations. Zero energy of the ''chiral domain'' configuration marks the point of the phase transition. We can calculate the ''chiral domain'' energy per one lattice cell. One graphene lattice cell contains two sites and tree links between sites. The condition of zero energy per lattice cell can be written as follows: $V_{00} = 3 \cdot V_{01}$. It is impossible to find the critical $V^c_{00}$ analytically if we take into account finite temperature and interaction at all radii. So, we'll use Monte-Carlo to study this problem further in the next section. \section{Monte Carlo investigation of the phase diagram} Let us turn to the numerical study of the models (\ref{aver}) and (\ref{simp_ham}) by means of Monte-Carlo method. We have used the heat bath algorithm for our calculations, taking into account double weight of the states with zero charge $q_x=0$. We used hexagonal lattice with $36 \times 36$ sites and periodical boundary conditions. We study the following observable to mark the phase transition: \begin{equation} \langle O \rangle = \langle q_x q_{x+\rho_b} \rangle. \end{equation} This average is very convenient for the detection of the transition between two vacua. Indeed, this quantity is close to zero in case disordered system. Otherwise, it is equal to $-1$ in the chiral domain phase, because $q_x$ and $q_{x+\rho_b}$ belong to different sublattices and have opposite charges. We generated 100 statically independent configurations for every value of $V_{00}$ and $\beta=1/T$. We changed the value of $V_{00}$ and hence the ratio between $V_{01}$ and ${V_{00}}$. The values of $V_{x y}$ at large distances correspond to ordinary Coulomb interaction. Its strength is defined by the $V_{01}$ potential. So we vary the value of ${V_{00}}$ potential leaving the other constant. At the beginning, we consider the simple model (\ref{simp_ham}) where only the nearest neighbours interact with each other. The phase transition is marked by the maximum of derivation ${\partial \langle O \rangle} \over { \partial T} $. The phase diagram is shown in the figure \ref{fig2}. There are tree different curves that corresponds to three different starts of Markov process during Monte-Carlo calculation. The bottom curve corresponds to start from configuration filled with zero charges. The top curve corresponds to chiral domain starting configuration. The curve between them corresponds to start from a special type of configuration: half of the lattice is filled with chiral domain and the rest of the lattice is filled with zero charges. The results of Monte Carlo algorithm shouldn't depend on the initial configuration for infinite thermalization time (infinite length of the Markov process). But in practice it's impossible to provide infinite time for thermalization. That is why we use that peculiar configuration for start. In this case, the long termalization is not necessary. We can only measure the change of the domain volume to determine the phase which will be a result of the thermalization process. The accuracy of this method can be confirmed by the correspondence between low-temperature extrapolation presented in the the figure \ref{fig2} and the theoretical value of the critical on-site interaction potential $V^c_{00} = 3 \cdot V_{01}=16.5$~eV. \begin{figure} \hspace*{-4.0cm} \centering \includegraphics[width=0.6\textwidth,angle=-90]{fig2_eps.eps} \vspace{-70pt} \caption{Phase diagram in the plane $(T, V_{00})$ for the model which takes into account only interaction of the nearest neighbours. Three different kinds of initial configurations for termalization were used in the simulations.} \label{fig2} \end{figure} Now let us investigate the system with interaction at all coordination radii. In this case we again use the start from lattice that is half-filled with domain. We detect the change of the chiral domain volume. If it becomes smaller, the domain melt, so the system turns into plasma. If domain grows, the systems turns into the phase with ordered charges. The result one can see in figure \ref{fig3}. The comparison of figures \ref{fig2} and \ref{fig3} allows to trace the influence of Coulomb interaction at large coordination radii. \begin{figure} \hspace*{-4.0 cm}\centering \includegraphics[width=0.6\textwidth,angle=-90]{fig3_eps.eps} \vspace{-70pt} \caption{Phase diagram in the plane $(T, V_{00})$ for the model with long range interaction. Start from halfdomain configuration.} \label{fig3} \end{figure} \section*{Discussions and conclusion} We investigated the phase diagram in the plane $ (T , V_ { 00 }) $ for the excitonic phase transition at hexagonal lattice. The main observation is that the value of the on-site interaction $V_{00}$ strongly influences the ground state of the system. Decreasing of $V_{00}$ can lead to the change of the system's ground state from zero charges configuration to the configuration with spontaneous charge separation between sublattices. The last phase is a dielectric with nonzero excitonic condensate. In the real graphene the $\sigma$ - orbital screening decreases the on-site potential $V_{00}$. But this screening is not enough strong for the phase transition even at zero temperature. The phase diagram strongly depends also on the interaction at large distances. So the real physical situation is a combination of two effects: the on-site interaction screened by $\sigma$ - orbitals and long range Coulomb interaction. The phase transition is very sensitive to both these factors. We obtained that free graphene is a conductor even at zero temperature. In order to achieve the dielectric state with nonzero excitonic condensate, the $\sigma$ - orbital screening should be stronger or something should block the Coulomb interaction between electrons at large distances.
\section{Introduction} \par{Pambuccian has offered two axiom systems $\Sigma$ and $\Sigmafano$~\cite{pambuccian2001two} for André's central translation structures~\cite{andre1961parallelstrukturen}. ($\Sigmafano$ is $\Sigma$ together with the Fano principle that diagonals of parallelograms are not parallel.) In this note we further develop Pambuccian's work by showing that:} \begin{itemize} \item {$\Sigma^{\prime}$ has a dependent axiom. Without this axiom, the axiom system is indeed independent.} \item {Without the dependent axiom, finite independence models exist for all axioms, whereas most of the independence proofs offered in~\cite{pambuccian2001two} used infinite models. In particular, one of the independence proofs was accomplished by proof-theoretic methods rather than by a independence model; but for this axiom, too, there is a finite independence model;} \item {Without its dependent axiom, $\Sigma^{\prime}$ is not just independent, but completely independent.} \end{itemize} \par{For the sake of completeness, we repeat here the definitions of the axiom systems $\Sigma$ and $\Sigmafano$. Both are based on classical one-sorted first-order logic with identity. Variables are intended to range over points. A single language is used for both axiom systems; the language has three constants, $a_{0}$, $a_{1}$, and $a_{2}$, a ternary relation $L$ for collinearity, and a single ternary function symbol $\tau$ for central translations: $\taufunc{a}{b}{c}$ is the image of $c$ under the translation that shifts $a$ to $b$. With the the binary operation symbol $\sigma$ understood as} \[ \sigmafunc{a}{b} = \taufunc{b}{a}{a}, \] \par{(essentially a point reflection), the axioms of $\Sigma$ are as follows:} \begin{description} \item[A3] $a \ne b \wedge \collinear{a}{b}{c} \wedge \collinear{a}{b}{d} \rightarrow \collinear{a}{c}{d}$ \item[B1] $\collinear{a}{b}{c} \rightarrow \collinear{b}{a}{c}$ \item[B2] $\taufunc{a}{b}{c} = \taufunc{a}{c}{b}$ \item[B3] $\collinear{a}{b}{\sigmafunc{a}{b}}$ \item[B4] $\collinear{a}{b}{c} \rightarrow \collinear{x}{\taufunc{a}{b}{x}}{\taufunc{a}{c}{x}}$ \item[B5] $\taufunc{a}{b}{x} = x \rightarrow a = b$ \item[B6] $\taufunc{a}{b}{x} = \taufunc{c}{\taufunc{a}{b}{x}}{x}$ \item[B7] $\neg \collinear{a_{0}}{a_{1}}{a_{2}}$ \end{description} \par{(The appearance of $\Aaxiom{3}$ without $\Aaxiom{1}$ and $\Aaxiom{2}$ is not an error. In the the official definition of $\Sigma$ from~\cite{pambuccian2001two}, rather than duplicating an axiom from André's axiom system, the names of whose axioms all have the prefix ``A'', under a new name, it is simply reused and the ``B'' axioms are offered.) The axiom system $\Sigma^{\prime}$ is $\Sigma$ together with} \begin{description} \item[B8] $\sigmafunc{a}{b} = b \rightarrow a = b$ \end{description} \par{$\Baxiom{8}$ captures the Fano principle that diagonals of parallelograms are not parallel.} \par{Universal quantifiers will usually be suppressed.} \section{A dependent axiom} \par{It was claimed in~\cite{pambuccian2001two} that $\Sigmafano$ is independent. This requires qualification: $\Baxiom{5}$ is indeed an independent axiom of $\Sigma$ (Proposition~\ref{b5-independent-of-sigma}), but $\Baxiom{5}$ can be proved from $\subtract{\Sigma}{\Baxiom{5}}$ with the help of the Fano principle $\Baxiom{8}$ (Proposition~\ref{b5-dependent}).} \begin{Proposition}\label{b5-independent-of-sigma}$\notproves{\subtract{\Sigma}{\Baxiom{5}}}{\Baxiom{5}}$.\end{Proposition} \begin{proof} \par{Consider the domain $\{1,2\}$ and interpret $\collinearsymbol$ and $\tausymbol$ according to Table~\ref{tab:b5-counterexample}. A counterexample to} \[ \taufunc{a}{b}{x} = x \rightarrow a = b \] \par{is provided by $(a,b,x) \coloneqq (1,2,2)$.} \begin{table} \begin{tabular}{lll} \hline Triple & $\tausymbol$ & $\collinearsymbol$\\ \hline $(1,1,1)$ & $2$ & $-$\\ $(1,1,2)$ & $2$ & $+$\\ $(1,2,1)$ & $2$ & $-$\\ $(1,2,2)$ & $2$ & $+$\\ $(2,1,1)$ & $2$ & $-$\\ $(2,1,2)$ & $2$ & $+$\\ $(2,2,1)$ & $2$ & $-$\\ $(2,2,2)$ & $2$ & $-$\\ \end{tabular}\caption{\label{tab:b5-counterexample}A model of $\add{\subtract{\Sigma}{\Baxiom{5}}}{\neg\Baxiom{5}}$} \end{table} For each pair $(a,b)$, ${\tausymbol}_{a,b}$ fails to be a transitive action because $1$ is never a value of $\tausymbol$. \end{proof} \begin{Lemma}\label{b5-dependent-lemma-2}$\proves{\Sigmafano}{\taufunc{a}{b}{c} = \taufunc{d}{c}{\taufunc{a}{b}{d}}}$ \end{Lemma} \begin{proof}{The desired conclusion follows from $\Baxiom{2}$ and $\Baxiom{6}$ ($\Baxiom{5}$ is not needed).}\end{proof} \begin{Lemma}\label{b5-dependent-lemma-3}$\proves{\Sigmafano}{\taufunc{a}{b}{a} = b}$\end{Lemma} \begin{proof}This is equation (3) of~\cite{pambuccian2001two}. It is derived without the help of $\Baxiom{5}$. (Indeed, $\{ \Baxiom{2}, \Baxiom{6}, \Baxiom{8} \}$ suffices).\end{proof} \begin{Proposition}\label{b5-dependent}$\proves{\subtract{\Sigmafano}{\Baxiom{5}}}{\Baxiom{5}}$.\end{Proposition} \begin{proof} \par{Suppose $\taufunc{a}{b}{x} = x$. From Lemmas~\ref{b5-dependent-lemma-2} and~\ref{b5-dependent-lemma-3}, we have} \begin{equation}\label{eq:1} \taufunc{u}{v}{w} = \taufunc{u}{w}{v}, \end{equation} \par{as well as} \begin{equation}\label{eq:2} \taufunc{\taufunc{t}{u}{v}}{w}{u} = \taufunc{v}{t}{w} \end{equation} \par{for all $t$, $u$, $v$, and $w$. Thus, by~(\ref{eq:1}), $\taufunc{a}{x}{b} = x$, whence (by Lemma~\ref{b5-dependent-lemma-2} and~(\ref{eq:2}),} \begin{equation*} \taufunc{u}{a}{v} = \taufunc{u}{b}{v} \end{equation*} \par{for all $u$ and $v$. Lemma~\ref{b5-dependent-lemma-3} then gives us the desired conclusion $a = b$.} \end{proof} \par{In light of Proposition~\ref{b5-dependent}, we define:} \begin{Definition}$\Sigmasharpened \coloneqq \Sigmafano \setminus \{ \Baxiom{5} \}$.\end{Definition} \par{In the following, the theory in focus is $\Sigmasharpened$ rather than $\Sigmafano$.} \section{Small finite independence models}\label{sec:finite-independence-models} \par{The next several propositions show that every axiom of $\Sigmasharpened$ has a finite independence model. (Incidentally, the cardinalities of the independence models are minimal: when it is claimed that there is a independence model for $\phi$ of cardinality $n$, it is also claimed that $\phi$ is true in every model of $\subtract{\Sigmasharpened}{\phi}$ of cardinality less than $n$.)} \par{In the independence models that follow we give only the interpretation of the predicate $\collinearsymbol$ and the function $\tausymbol$. Strictly speaking, this is not enough, because we need to interpret the constants $a_{0}$, $a_{1}$, and $a_{2}$ so that $\Baxiom{7}$ holds. But in the independence models there is always at least one triangle; from any one, an interpretation of the constants $a_{0}$, $a_{1}$, and $a_{2}$ can be chosen so that $\Baxiom{7}$ is satisfied.} \begin{Proposition}\label{a3-finite-model}There exists an independence model for $\Aaxiom{3}$ of cardinality~$3$.\end{Proposition} \begin{proof} \par{Without $\Aaxiom{3}$ one cannot prove} \[ \collinear{a}{a}{b} \wedge \collinear{a}{b}{a} \wedge \collinear{a}{b}{b}, \] \par{the failure of which opens the door to geometrically counterintuitive models. Consider the domain $\{1,2,3\}$, interpret $\collinearsymbol$ and $\tausymbol$ as in Table~\ref{tab:a3-counterexample}. Note that the model is ``collinear'' in the sense that $\collinear{\alpha(1)}{\alpha(2)}{\alpha(3)}$ for any permutation $\alpha$ of $\{ 1, 2 3 \}$; nonetheless, for many pairs $(u,v)$ of distinct points, the various collinearity statements one can make about $u$ and $v$ are false. A counterexample to} \[ a \ne b \wedge \collinear{a}{b}{c} \wedge \collinear{a}{b}{d} \rightarrow \collinear{a}{c}{d} \] \par{is $(a,b,c,d) \coloneqq (3,2,1,1)$.} \begin{table}[h] \begin{tabular}{lll|lll|lll} \hline Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$\\ \hline $(1,1,1)$ & $1$ & $+$ & $(2,1,1)$ & $3$ & $+$ & $(3,1,1)$ & $2$ & $-$\\ $(1,1,2)$ & $2$ & $-$ & $(2,1,2)$ & $1$ & $-$ & $(3,1,2)$ & $3$ & $+$\\ $(1,1,3)$ & $3$ & $+$ & $(2,1,3)$ & $2$ & $+$ & $(3,1,3)$ & $1$ & $+$\\ $(1,2,1)$ & $2$ & $+$ & $(2,2,1)$ & $1$ & $+$ & $(3,2,1)$ & $3$ & $+$\\ $(1,2,2)$ & $3$ & $-$ & $(2,2,2)$ & $2$ & $+$ & $(3,2,2)$ & $1$ & $+$\\ $(1,2,3)$ & $1$ & $+$ & $(2,2,3)$ & $3$ & $-$ & $(3,2,3)$ & $2$ & $-$\\ $(1,3,1)$ & $3$ & $-$ & $(2,3,1)$ & $2$ & $+$ & $(3,3,1)$ & $1$ & $-$\\ $(1,3,2)$ & $1$ & $+$ & $(2,3,2)$ & $3$ & $+$ & $(3,3,2)$ & $2$ & $+$\\ $(1,3,3)$ & $2$ & $+$ & $(2,3,3)$ & $1$ & $-$ & $(3,3,3)$ & $3$ & $+$\\ \end{tabular}\caption{\label{tab:a3-counterexample}A model of $\add{\subtract{\Sigmasharpened}{\Aaxiom{3}}}{\neg\Aaxiom{3}}$} \end{table} \end{proof} \begin{Proposition}\label{b1-finite-model}There exists an independence model for $\Baxiom{1}$ of cardinality~$3$.\end{Proposition} \begin{proof} \par{The difficulty here is that} \[ \collinear{a}{b}{a} \] \par{fails without $\Baxiom{1}$. Consider the domain $\{ 1,2,3 \}$ and the interpretations of $\collinearsymbol$ and $\tausymbol$ as in Table~\ref{tab:b1-counterexample}. From the standpoint of $\collinearsymbol$, the model is nearly trivialized; the only triples $(a,b,c)$ where $\collinear{a}{b}{c}$ fails are where $a \ne b$ and $a = c$. An example where} \[ \collinear{a}{b}{c} \rightarrow \collinear{b}{a}{c} \] \par{fails is $(a,b,c) \coloneqq (1,2,3)$.} \begin{table} \begin{tabular}{lll|lll|lll} \hline Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$\\ \hline $(1,1,1)$ & $1$ & $+$ & $(2,1,1)$ & $3$ & $+$ & $(3,1,1)$ & $2$ & $+$\\ $(1,1,2)$ & $2$ & $+$ & $(2,1,2)$ & $1$ & $-$ & $(3,1,2)$ & $3$ & $+$\\ $(1,1,3)$ & $3$ & $+$ & $(2,1,3)$ & $2$ & $+$ & $(3,1,3)$ & $1$ & $-$\\ $(1,2,1)$ & $2$ & $-$ & $(2,2,1)$ & $1$ & $+$ & $(3,2,1)$ & $3$ & $+$\\ $(1,2,2)$ & $3$ & $+$ & $(2,2,2)$ & $2$ & $+$ & $(3,2,2)$ & $1$ & $+$\\ $(1,2,3)$ & $1$ & $+$ & $(2,2,3)$ & $3$ & $+$ & $(3,2,3)$ & $2$ & $-$\\ $(1,3,1)$ & $3$ & $-$ & $(2,3,1)$ & $2$ & $+$ & $(3,3,1)$ & $1$ & $+$\\ $(1,3,2)$ & $1$ & $+$ & $(2,3,2)$ & $3$ & $-$ & $(3,3,2)$ & $2$ & $+$\\ $(1,3,3)$ & $2$ & $+$ & $(2,3,3)$ & $1$ & $+$ & $(3,3,3)$ & $3$ & $+$\\ \end{tabular}\caption{\label{tab:b1-counterexample}A model of $\add{\subtract{\Sigmasharpened}{\Baxiom{1}}}{\neg\Baxiom{1}}$} \end{table} \end{proof} \begin{Proposition}\label{b2-finite-model}There exists an independence model for $\Baxiom{2}$ of cardinality~$3$.\end{Proposition} \begin{proof} \par{Consider the domain $\{ 1,2,3 \}$ and the interpretations of $\collinearsymbol$ and $\tausymbol$ are as in Table~\ref{tab:b2-counterexample}. A counterexample to} \[ \taufunc{a}{b}{c} = \taufunc{a}{c}{b} \] \par{in this structure is $(a,b,c) \coloneqq (1,3,1)$.} \begin{table} \begin{tabular}{lll|lll|lll} \hline Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$\\ \hline $(1,1,1)$ & $1$ & $+$ & $(2,1,1)$ & $1$ & $+$ & $(3,1,1)$ & $1$ & $+$\\ $(1,1,2)$ & $1$ & $+$ & $(2,1,2)$ & $1$ & $+$ & $(3,1,2)$ & $1$ & $+$\\ $(1,1,3)$ & $1$ & $-$ & $(2,1,3)$ & $1$ & $-$ & $(3,1,3)$ & $1$ & $-$\\ $(1,2,1)$ & $2$ & $+$ & $(2,2,1)$ & $2$ & $+$ & $(3,2,1)$ & $2$ & $+$\\ $(1,2,2)$ & $2$ & $+$ & $(2,2,2)$ & $2$ & $+$ & $(3,2,2)$ & $2$ & $+$\\ $(1,2,3)$ & $2$ & $-$ & $(2,2,3)$ & $2$ & $-$ & $(3,2,3)$ & $2$ & $-$\\ $(1,3,1)$ & $2$ & $+$ & $(2,3,1)$ & $1$ & $+$ & $(3,3,1)$ & $2$ & $+$\\ $(1,3,2)$ & $2$ & $+$ & $(2,3,2)$ & $1$ & $+$ & $(3,3,2)$ & $2$ & $+$\\ $(1,3,3)$ & $2$ & $-$ & $(2,3,3)$ & $1$ & $-$ & $(3,3,3)$ & $2$ & $-$\\ \end{tabular}\caption{\label{tab:b2-counterexample}A model of $\add{\subtract{\Sigmasharpened}{\Baxiom{2}}}{\neg\Baxiom{2}}$} \end{table} \end{proof} \begin{Proposition}\label{b3-finite-model}There exists an independence model for $\Baxiom{3}$ of cardinality~$1$.\end{Proposition} \begin{proof} \par{$\Baxiom{3}$ is the only axiom that outright asserts that some points are collinear. Thus, if every triple $(a,b,c)$ of points constitutes a triangle (that is, $\neg\collinear{a}{b}{c}$ holds for all $a$, $b$, and $c$), then clearly $\Baxiom{3}$ would be falsified. So take a $1$-element structure and interpret $\collinearsymbol$ so that $\collinear{a}{b}{c}$ for the unique triple $(a,b,c)$ of the structure is false. The interpretations of $\tau$, $a_{0}$, $a_{1}$, and $a_{2}$ are forced. One can check that all axioms, except $\Baxiom{3}$, are satisfied.} \end{proof} \par{Pambuccian was unable to find an independence model of $\Baxiom{4}$. To show that $\Baxiom{4}$ is independent of the other axioms, methods of structural proof analysis~\cite{negri2008structural} were employed. Specifically, an analysis of all possible formal derivations starting from $\Sigma$ was made, and by a syntactic-combinatorial argument it was found that $\Baxiom{4}$ is underivable from $\subtract{\Sigma}{\Baxiom{4}}$. By the completeness theorem, then, there must exist an independence model. Here is one:} \begin{Proposition}\label{b4-finite-model}There exists an independence model for $\Baxiom{4}$ of cardinality~$27$.\end{Proposition} \begin{proof} \par{For lack of space, we omit an explicit description of the $27 \times 27 \times 27$ table. The model was found by the finite model-finding program {\macefour}~\cite{prover9-mace4}.} \end{proof} \par{Consideration of $\Baxiom{5}$ is skipped because it is a dependent axiom of $\Sigma$ (and in any case is not officially an axiom of $\Sigmasharpened$).} \begin{Proposition}\label{b6-finite-model}There exists a independence model for $\Baxiom{6}$ of cardinality~$3$.\end{Proposition} \begin{proof} \par{Consider the domain $\{ 1,2,3 \}$ and the interpretations of $\collinearsymbol$ and $\tausymbol$ as in Table~\ref{tab:b6-counterexample}. A counterexample to} \[ \taufunc{a}{b}{x} = \taufunc{c}{\taufunc{a}{b}{x}}{x} \] is given by $(a,b,c,x) \coloneqq (2,2,1,2)$. \begin{table} \begin{tabular}{lll|lll|lll} \hline Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$ & Triple & $\tausymbol$ & $\collinearsymbol$\\ \hline $(1,1,1)$ & $2$ & $+$ & $(2,1,1)$ & $1$ & $+$ & $(3,1,1)$ & $2$ & $+$\\ $(1,1,2)$ & $2$ & $+$ & $(2,1,2)$ & $1$ & $+$ & $(3,1,2)$ & $2$ & $+$\\ $(1,1,3)$ & $2$ & $-$ & $(2,1,3)$ & $2$ & $-$ & $(3,1,3)$ & $2$ & $-$\\ $(1,2,1)$ & $2$ & $+$ & $(2,2,1)$ & $1$ & $+$ & $(3,2,1)$ & $2$ & $+$\\ $(1,2,2)$ & $2$ & $+$ & $(2,2,2)$ & $1$ & $+$ & $(3,2,2)$ & $2$ & $+$\\ $(1,2,3)$ & $2$ & $-$ & $(2,2,3)$ & $2$ & $-$ & $(3,2,3)$ & $2$ & $-$\\ $(1,3,1)$ & $2$ & $+$ & $(2,3,1)$ & $2$ & $+$ & $(3,3,1)$ & $2$ & $+$\\ $(1,3,2)$ & $2$ & $+$ & $(2,3,2)$ & $2$ & $+$ & $(3,3,2)$ & $2$ & $+$\\ $(1,3,3)$ & $2$ & $-$ & $(2,3,3)$ & $1$ & $-$ & $(3,3,3)$ & $2$ & $-$\\ \end{tabular}\caption{\label{tab:b6-counterexample}A model of $\add{\subtract{\Sigmasharpened}{\Baxiom{6}}}{\neg\Baxiom{6}}$} \end{table} \end{proof} \par{We are unable to improve upon the independence proof for $\Baxiom{7}$ given in~\cite{pambuccian2001two}: the independence model given there has cardinality~$1$.} \begin{Proposition}\label{b8-finite-model}There exists an independence model for $\Baxiom{8}$ of cardinality~2.\end{Proposition} \begin{proof} Indeed, a suitable structure is already available: the countermodel $M$ for $\Baxiom{5}$ (over $\Sigma$) also falsifies $\Baxiom{8}$; since $\Baxiom{5}$ is not an axiom of $\Sigmasharpened$, $M$ works. A counterexample to \[ \sigmafunc{a}{b} = b \rightarrow a = b \] is given by $(a,b) \coloneqq (1,2)$. In this structure, the value of $\tau$, and hence $\sigmasymbol$, is always $2$. \end{proof} \section{Complete independence} The notion of completely independent set was proposed by E.~H.~Moore~\cite{moore1910introduction}. It is a considerably stronger property of an axiom system than the familiar notion of independence. \begin{Definition An axiom system $X$ is said to be completely independent if, for all subsets $A$ of $X$, the set $A \cup \{ \phi \in X \setminus A : \neg\phi \}$ is satisfiable. \end{Definition} If an axiom system $X$ is completely independent then it is also dependent: for every sentence $\phi$ of $X$, we have that $X \setminus \{ \phi \} \cup \{ \neg \phi \}$ is satisfiable, or (by the completeness theorem), that $\notproves{X \setminus \{ \phi \}}{\phi}$. When an axiom system is completely independent, no Boolean combination of its axioms can be proved from the other axioms. \begin{Theorem} $\Sigma$ is completely independent. \end{Theorem} \begin{proof} \par{Since $\Sigma$ has $8$ axioms, by following the definition of complete independence one sees that there are $2^{8}$ sets of formulas to check for satisfiability. (Such an enumeration of cases is best executed mechanically rather than by hand; we were assisted by the {\tipi} program~\cite{alama2012tipi}.) But for all cases, very small finite models can be found with the help of a finite model-finder for first-order classical logic (e.g., \paradox~\cite{claessen2003new}). For lack of space we do not present all the models here.} \end{proof} \begin{Theorem}$\Sigmasharpened$ is completely independent.\end{Theorem} \begin{proof} \par{As with $\Sigma$, $\Sigmasharpened$ has $8$ axioms, so again one has $2^{8}$ sets of formulas to check for satisfiability. Except for two cases, very small finite models can be found almost immediately. The only cases---both involving $\Baxiom{4}$---that cannot be immediately dispensed with are:} \begin{enumerate} \item {$\Sigmasharpened \setminus \{ \Baxiom{4} \} \cup \{ \neg\Baxiom{4} \}$, and} \item {$\Sigmasharpened \setminus \{ \Baxiom{4} , \Baxiom{7} \} \cup \{ \neg\Baxiom{4} , \neg\Baxiom{7} \}$.} \end{enumerate} \par{Case (1): The satisfiability of $\Sigmasharpened \setminus \{ \Baxiom{4} \} \cup \{ \neg\Baxiom{4} \}$ is, by the completeness theorem, the same thing as the independence of $\Baxiom{4}$. The proof in~\cite{pambuccian2001two} works. Recall that the smallest independence model for this axiom (27) is much larger than the other independence models (which are all size 3 or less).} \par{Case (2): $\Sigmasharpened \setminus \{ \Baxiom{4} , \Baxiom{7} \} \cup \{ \neg\Baxiom{4} , \neg\Baxiom{7} \}$ is satisfiable. Take a model $M$ of $\subtract{\Sigmasharpened}{\Baxiom{4}}$ in which $\Baxiom{4}$ is false. Since $\Baxiom{4}$ fails, there exists points $a$, $b$, $c$, and $d$ in $M$ such that $\collinear{a}{b}{c}$ but $\neg\collinear{d}{\taufunc{a}{b}{d}}{\taufunc{a}{c}{d}}$. An appropriate model is obtained from $M$ by changing $M$'s interpretation of $a_{0}$, $a_{1}$, and $a_{2}$ to, respectively, $a$, $\taufunc{a}{b}{d}$, and $\taufunc{a}{c}{d}$.} \end{proof} \par{The treatment of case (2) in the preceding proof might be regarded as somewhat odd. The model at work there did contain triangles (that is, it was ``nonlinear''), yet it falsified $\Baxiom{7}$. But if $\Baxiom{7}$ is false, shouldn't the model be ``linear''?} \par{It is worth noting that $\Sigmasharpened$ is just barely completely independent; a seemingly innocent change to one it is axioms destroys complete independence. Consider the existential generalization of $\Baxiom{7}$} \begin{description} \item[B7$^{\prime}$] {$\exists a,b,c \left [ \neg\collinear{a}{b}{c} \right ]$} \end{description} \par{and let $\Sigmasharpened^{\prime}$ be $\add{\subtract{\Sigmasharpened}{\Baxiom{7}}}{\Baxiom{7^{\prime}}}$.} \par{Intuitively, $\Baxiom{7^{\prime}}$ says the same thing as $\Baxiom{7}$. Every model of $\Sigmasharpened$ is, by ignoring the interpretations of $a_{0}$, $a_{1}$, and $a_{2}$, a model of $\Sigmasharpened^{\prime}$. And every model of $\Sigmasharpened^{\prime}$ can be extended to a model of $\Sigmasharpened$ by choosing, for the interpretation of $a_{0}$, $a_{1}$, and $a_{2}$, any witness to the truth of $\Baxiom{7^{\prime}}$. Nonetheless, the two theories are subtly different:} \begin{Proposition}\label{not-completely-independent}$\Sigmasharpened^{\prime}$ is independent but not completely independent.\end{Proposition} \begin{proof} \par{The independence proofs for $\Sigmasharpened$ are easily adapted to $\Sigmasharpened^{\prime}$. As for complete independence, note that $\Baxiom{7^{\prime}}$ is true in every model of $\add{\subtract{\Sigmasharpened^{\prime}}{\Baxiom{4}}}{\neg\Baxiom{4}}$. Thus, $\Sigmasharpened^{\prime} \setminus \{ \Baxiom{4}, \Baxiom{7^{\prime}} \} \cup \{ \neg\Baxiom{4}, \neg\Baxiom{7^{\prime}} \}$ is unsatisfiable.} \end{proof} \par{In other words, if $\Baxiom{4}$ is false, then there is a triangle, i.e., $\Baxiom{7^{\prime}}$ holds. This illustrates a relationship among the axioms of $\Sigmasharpened^{\prime}$ that the notion of complete independence is intended to rule out.} \par{$\Sigmasharpened^{\prime}$ is quite far from being completely independent. Although $\Baxiom{7^{\prime}}$ cannot be proved from the other axioms, with seven exceptions (see Table~\ref{tab:compatible-with-neg-b7-prime}), if any of $\Sigmasharpened^{\prime}$'s axiom is negated, $\Baxiom{7^{\prime}}$ becomes provable; that is, the other $2^{8} - 7$ Boolean combinations are incompatible with $\neg\Baxiom{7^{\prime}}$. In a rough sense, then, $\Baxiom{7^{\prime}}$ is ``almost'' a theorem of $\subtract{\Sigmasharpened^{\prime}}{\Baxiom{7^{\prime}}}$.} \begin{center} \begin{table}[h] \begin{tabular}{llllllll} \hline $\Aaxiom{3}$ & $\Baxiom{1}$ & $\Baxiom{2}$ & $\Baxiom{3}$ & $\Baxiom{4}$ & $\Baxiom{5}$ & $\Baxiom{6}$ & $\Baxiom{8}$\\ \hline $+$ & $+$ & $+$ & $-$ & $+$ & $+$ & $-$ & $-$\\ $+$ & $+$ & $+$ & $+$ & $+$ & $+$ & $-$ & $-$\\ $+$ & $+$ & $+$ & $-$ & $+$ & $+$ & $+$ & $-$\\ $+$ & $+$ & $+$ & $-$ & $+$ & $+$ & $-$ & $+$\\ $+$ & $+$ & $+$ & $+$ & $+$ & $+$ & $+$ & $-$\\ $+$ & $+$ & $+$ & $+$ & $+$ & $+$ & $-$ & $+$\\ $+$ & $+$ & $+$ & $-$ & $+$ & $+$ & $+$ & $+$\\ \end{tabular}\caption{\label{tab:compatible-with-neg-b7-prime}Boolean combinations of axioms of $\Sigmasharpened$ compatible with $\neg\Baxiom{7^{\prime}}$.} \end{table} \end{center} \par{The difference between $\Baxiom{7}$ and $\Baxiom{7^{\prime}}$ is now clear. $\Baxiom{7}$ makes an assertion about three specific (though undetermined) points which are not mentioned anywhere else in the axioms and are thus ``semantically inert''. By contrast, $\Baxiom{7^{\prime}}$ is a purely existential sentence that can ``interact'' with the other axioms (specifically, $\Baxiom{4}$).} \par{Similarly, in the foundations of logic, a result similar to Proposition~\ref{not-completely-independent} was discovered by Dines~\cite{dines1915complete}: among several axioms, only one (also having the flavor of a minimal-cardinality principle) was an obstacle to complete independence.} \par{Interestingly, the axiom system $\Gamma$ from which $\Sigma$ is derived has many dependent axioms. $\Gamma$ is, moreover, very far from being completely independent. Thus $\Sigma$ is, from a certain methodological standpoint, to be preferred to $\Gamma$.} \bibliographystyle{amsplain}
\section{Introduction}\label{intro_sec} For a finite algebra $\m a$, write $d_{\m a}(n)=g$ if $g$ is the least size of a generating set for $\m a^n$, and write $h_{\m a}(g)=n$ if the largest power of $\m a$ that is $g$-generated is $\m a^n$. The functions $d_{\m a}$ and $h_{\m a}$ map natural numbers to natural numbers and are related by \[ d_{\m a}(n) \leq g \wrel{\Longleftrightarrow} \textrm{$\m a^n$ is $g$-generated} \wrel{\Longleftrightarrow} n \leq h_{\m a}(g), \] which asserts that $d_{\m a}$ is the lower adjoint of $h_{\m a}$ and $h_{\m a}$ is the upper adjoint of $d_{\m a}$. It follows that $d_{\m a}, h_{\m a}\colon \omega\to\omega$ are increasing functions, which are inverse bijections between their images: \[ \im(d_{\m a}) \xrightleftharpoons[d]{h} \im(h_{\m a}); \] and, moreover, each determines the other. These functions make sense for partial algebras and infinite algebras, too. The study of the functions $d_{\m a}$ and $h_{\m a}$ has a long history, which we briefly survey. \subsection{The $\phi$-function of a group} In the 1936 paper \cite{hall}, Philip Hall generalizes the Euler $\phi$-function from number theory by defining $\phi_k(G)$ to be the number of $k$-tuples $\wec{t}=(t_1,\ldots,t_k)$ for which $\{t_1,\ldots,t_k\}$ is a generating set of the group $G$. The classical Euler $\phi$-function is therefore $\phi(k)=\phi_1(\mathbb Z_k)$. Hall calls two generating $k$-tuples $\wec{t}_1$ and $\wec{t}_2$ ``equivalent'' if there is an automorphism $\alpha$ of $G$ which applied coordinatewise to $\wec{t}_1$ yields $\wec{t}_2$. The automorphism group of $G$ acts freely on generating $k$-tuples, hence the number of equivalence classes of generating $k$-tuples is $\phi_k(G)/|\textrm{Aut}(G)|$. Hall denotes $\phi_k(G)/|\textrm{Aut}(G)|$ by $d_k(G)$, an unfortunate conflict with more recent notation since $\phi_k(G)/|\textrm{Aut}(G)|$ is closer to the $h$-function than to the $d$-function. Indeed, if $G$ is a finite simple nonabelian group, then $h_G(k) = \phi_k(G)/|\textrm{Aut}(G)|$. Hall calls the function $\phi_k(G)/|\textrm{Aut}(G)|$ ``intrinsically more interesting'' than $\phi_k(G)$, and derives a formula for it in the case where $G$ is a finite simple nonabelian group, namely \begin{equation}\label{hall_formula} h_{G}(k) = \frac{1}{|\textrm{Aut}(G)|}\sum_{H\leq G} \mu(H)|H|^k \end{equation} where $\mu$ is the M\"obius function of the subgroup lattice of $G$. This calculation is the first result of our topic. \subsection{Non-Hopf kernels} A group is \emph{Hopfian} if every surjective endomorphism is an isomorphism, and non-Hopfian otherwise. A group $N$ is a \emph{non-Hopf kernel} of $G$ if it is isomorphic to the kernel of a surjective endomorphism of $G$ that is not an isomorphism. In the 1969 paper \cite{dey}, I.~M.~S.~Dey investigates the problem of determining which groups are non-Hopf kernels. Dey notes that every nontrivial group is a non-Hopf kernel, since, for example, the kernel of the shift \[ N^{\omega}\to N^{\omega}\colon (n_0,n_1,n_2,\ldots)\mapsto (n_1,n_2,n_3\ldots) \] is isomorphic to $N$. Dey restricts attention to non-Hopf kernels of finitely generated groups, and notes the following: a finite complete group is not a non-Hopf kernel of a finitely generated group. ($N$ is complete if it is centerless and $\textrm{Aut}(N)=\textrm{Inn}(N)$.) His reasoning goes like this: if $N$ is complete and a non-Hopf kernel of $G$, then $C_G(N)$ is a normal complement to $N$. By the non-Hopf property, $C_G(N)\cong G$, so \[ G\cong N\times G\cong N^2\times G\cong N^3\times G\cong \cdots. \] If $G$ is finitely generated, say by $g$ elements, then so are the quotient groups $N^n$ for all finite $n$. But this contradicts the local finiteness of the variety ${\mathcal V}(N)$. Specifically, the $g$-generated groups in this variety have size at most $|N|^{|N|^g}$. Thus, Dey's paper draws attention to the (easy) fact that if $N$ is finite, then the number of elements required to generate $N^n$ goes to infinity as $n$ goes to infinity. (In symbols, $\lim_{n\to\infty}(d_{N}(n))=\infty$.) \subsection{Growth rates of groups} In the 1974 paper \cite{wiegold1}, James Wiegold cites Dey's work on non-Hopf kernels as the inspiration for his investigation into the question ``What are the ways in which \ldots [$d_G(n)$] \ldots can tend to infinity [when $G$ is a finite group]?'' Wiegold inverts Hall's formula (\ref{hall_formula}) to show that, for $n>0$, $d_{G}(n)$ is one of the three natural numbers nearest \[ \log_{|G|}(n) + \log_{|G|}(|\textrm{Aut}(G)|) \] when $G$ is a finite simple nonabelian group, so in this case $d_G(n)$ is asymptotically equivalent to $\log(n)$. He shows that $d_G(n)$ has logarithmic upper and lower bounds whenever $G$ is a finite perfect group. ($G$ is perfect if $[G,G]=G$.) He shows also that $d_G(n)$ agrees with a linear function for large $n$ if $G$ is a finite imperfect group. Thus, he establishes that $d_G(n)$ tends to infinity as a logarithmic or linear function when $G$ is a finite group. \subsection{Growth rates of groups, semigroups and group expansions} Wiegold's paper initiated a program of research into growth rates of groups including, for example, \cite{erfanian1,erfanian2,erfanian3,erfanian4,erfanian5, erfanian_rezaei, erfanian_wiegold, kimmerle, lennox_wiegold, meier_wiegold, obraztsov, stewart_wiegold, wiegold2, wiegold3, wiegold4, wiegold_wilson,wise}. The program expanded to include the investigation of growth rates of semigroups, in \cite{pollak, wiegold_sem}, and later to include the investigation of growth rates of more general algebraic structures, in \cite{glass_riedel, riedel}. Some of the questions being investigated about growth rates of finite algebras are related to the following theorems of Wiegold: \begin{enumerate} \item[(I)] A finite perfect group has growth rate that is logarithmic ($d_{\m a}(n)\in \Theta(\log(n))$), while a finite imperfect group has growth rate that is linear ($d_{\m a}(n)\in \Theta(n)$). \item[(II)] A finite semigroup with identity has growth rate that is logarithmic or linear, while a finite semigroup without identity has growth rate that is exponential ($d_{\m a}(n)\in 2^{\Theta(n)}$), \cite{wiegold_sem}. \end{enumerate} \noindent Herbert Riedel partially extends Item (I) to congruence uniform varieties in \cite{riedel} by proving that finite algebras in such varieties that are perfect (in the sense of modular commutator theory) have logarithmic growth rate. The paper \cite{quick_ruskuc} by Martyn Quick and Nikola Ru\v skuc extends Item (I) to any variety of rings, modules, $k$-algebras or Lie algebras, but also falls short of extending Item (I) to arbitrary congruence uniform varieties. \subsection{Our work}\label{our_work} We got interested in growth rates of finite algebras after reading Remark 4.15 of \cite{quick_ruskuc}, which states that \emph{``At present no finite algebraic structure is known for which the $d$--sequence does not have one of logarithmic, linear or exponential growth.''} We found some of these missing algebras. (Theorem~\ref{example}.) Our interest in growth rates was later strengthened upon learning about paper \cite{chen}, by Hubie Chen, which links growth rates with the constraint satisfaction problem by giving a polynomial time reduction from the quantified constraint satisfaction problem to the ordinary constraint satisfaction problem for algebras with $d_{\m a}(n)\in O(n^k)$ for some $k$. Our new algebras are relevant to this investigation. Our work is currently a 3-paper series, of which this is the first. \subsubsection This paper} The results from \cite{quick_ruskuc}, about growth rates in varieties of classical algebraic structures, can be presented in a stronger way. Let $\Sigma$ be a set of identities. If $\m a$ is an algebra in a language $\mathcal K$, then say that $\m a$ \emph{realizes} $\Sigma$ if there is a way to interpret the function symbols occurring in $\Sigma$ as $\mathcal K$-terms in such a way that each identity in $\Sigma$ holds in $\m a$. What is really proved in \cite{quick_ruskuc} is that if $\Sigma_{\text{Grp}}$ is the set of identities axiomatizing the variety of groups and $\m a$ is a finite algebra realizing $\Sigma_{\text{Grp}}$, then $\m a$ has a logarithmic growth rate if it is perfect and has a linear growth rate if it is imperfect. Although the results of \cite{quick_ruskuc} are stated for only a few specific varieties of group expansions, the results hold for \emph{any} variety of group expansions. The main results of this paper are also best expressed in the terminology of algebras realizing a set of identities. Call a term \emph{basic} if it contains at most one nonnullary function symbol. An identity $s\approx t$ is basic if the terms on both sides are. This paper is an investigation into the restrictions imposed on growth rates of finite algebras by a set $\Sigma$ of basic identities. A new concept that emerges from this investigation is the notion of a pointed cube term. If $\Sigma$ is a set of identities in a language $\mathcal L$, then an $\mathcal L$-term $F(x_1,\ldots,x_m)$ is a \emph{$p$-pointed, $k$-cube term} for the variety axiomatized by $\Sigma$ if there is a $k\times m$ matrix $M$ consisting of variables and $p$ distinct constant symbols, with every column of $M$ containing a symbol different from $x$, such that \begin{equation}\label{cube} \Sigma\models F(M)\approx \left( \begin{matrix} x\\ \vdots\\ x \end{matrix} \right). \end{equation} (\ref{cube}) is meant to be a compact representation of a sequence of $k$ row identities of a special kind. For example, \begin{equation}\label{m} \Sigma\models {m}\left(\begin{matrix} x&y&y\\ y&y&x \end{matrix} \right) \approx \left(\begin{matrix} x\\ x \end{matrix} \right), \end{equation} which is the assertion that $\Sigma\models m(x,y,y)\approx x$ and $\Sigma\models m(y,y,x)\approx x$, witnesses that $m(x_1,x_2,x_3)$ is a $3$-ary, $0$-pointed, $2$-cube term. The basic identities (\ref{m}) define what is called a \emph{Maltsev term}. For another example, \begin{equation}\label{b} \Sigma\models {B}\left(\begin{matrix} 1&x\\ x&1 \end{matrix} \right) \approx \left(\begin{matrix} x\\ x \end{matrix} \right), \end{equation} which is the assertion that $\Sigma\models B(1,x)\approx x$ and $\Sigma\models B(x,1)\approx x$, witnesses that $B(x_1,x_2)$ is a $2$-ary, $1$-pointed, $2$-cube term. As a final example, \begin{equation}\label{majority} \Sigma\models {M}\left(\begin{matrix} y&x&x\\ x&y&x\\ x&x&y \end{matrix} \right) \approx \left(\begin{matrix} x\\ x\\ x \end{matrix} \right), \end{equation} which is the assertion that $M$ is a majority term for the variety axiomatized by $\Sigma$, witnesses that $M(x_1,x_2,x_3)$ is a $3$-ary, $0$-pointed, $3$-cube term. To state our main results, let $\Sigma$ be a set of basic identities. We show that \begin{enumerate} \item[(1)] The growth rate of any partial algebra can be realized as the growth rate of a total algebra (Corollary~\ref{partial_cor}). If the partial algebra is finite, then the total algebra can be taken to be finite. \item[(2)] A function $D\colon \omega\to\omega^+$ arises as the $d$-function of a countably infinite algebra if and only if (i)~$D$ is increasing and satisfies (ii)~$D(0)=0$ or $1$, and (iii)~\mbox{$D(2)>0$} (Theorem~\ref{big}). \item[(3)] If $\Sigma$ does not entail the existence of a pointed cube term, then $\Sigma$ imposes no restriction on growth rates of algebras (Theorem~\ref{nonrestrictive_thm}). That is, for every algebra $\m a$ there is an algebra $\m b$ realizing $\Sigma$ such that $d_{\m b}=d_{\m a}$. The algebra $\m b$ can be taken to be finite if $\m a$ is finite and the set $\Sigma$ involves only finitely many distinct constants. \item[(4)] If $\Sigma$ entails the existence of a $p$-pointed cube term, $p\geq 1$, then any algebra $\m a$ realizing $\Sigma$ such that $\m a^{p+k-1}$ is finitely generated has growth rate that is bounded above by a polynomial (Theorem~\ref{pointed_polynomial}). This is a nontrivial restriction. \item[(5)] There exist finite algebras with pointed cube terms whose growth rate is asymptotically equivalent to a polynomial of any prescribed degree (Theorem~\ref{example}). \item[(6)] Any function that arises as the growth rate of an algebra with a pointed cube term also arises as the growth rate of an algebra without a pointed cube term (Theorem~\ref{avoid_thm}). \end{enumerate} In addition to these items we give a new proof of Kelly's Completeness Theorem for basic identities (Theorem~\ref{kelly}). We give a procedure, based on this theorem, for deciding if a finite set of basic identities implies the existence of a pointed cube term (Corollary~\ref{decide_cor}). \subsubsection{Our second paper, \cite{paper2}} We investigate growth rates of algebras with a $0$-pointed $k$-cube term, which we shall just call a ``$k$-cube term''. Such terms were first identified in \cite{bimmvw} in connection with investigations into constraint satisfaction problems, while an equivalent type of term was identified independently in \cite{parallelogram} in connection with investigations into compatible relations of algebras. We show in \cite{paper2} that if $\m a$ has a $k$-cube term and $\m a^{k}$ is finitely generated, then $d_{\m a}(n)\in O(\log(n))$ if $\m a$ is perfect, while $d_{\m a}(n)\in O(n)$ if $\m a$ is imperfect. One can strengthen `Big Oh' to `Big Theta' if $\m a$ is finite. This extends Wiegold's result (I) for groups to a setting that includes, as special cases, any finite algebra with a Maltsev term (in particular, any finite algebra in a congruence uniform variety) or any finite algebra with a majority term. \subsubsection{Our third paper, \cite{paper3}} We investigate growth rates of finite solvable algebras. Our original aim was to show that the only growth rates exhibited by such algebras are linear or exponential functions. We do prove this for finite nilpotent algebras and we prove it for finite solvable algebras with a pointed cube term, but the general case of a finite solvable algebra without a pointed cube term remains open. \section{Preliminaries}\label{prelim_sec} \subsection{Notation}\label{notation} $[n]$ denotes the set $\{1,\ldots,n\}$. A tuple in $A^n$ may be denoted $(a_1,\ldots,a_n)$ or $\wec{a}$, and may be viewed as a function $\wec{a}\colon [n]\to A$. A tuple $(a,a,\ldots,a)\in A^n$ with all coordinates equal to $a$ may be denoted $\hat{a}$. The size of a set $A$, the length of a tuple $\wec{a}$, and the length of a string $\sigma$ are denoted $|A|$, $|\wec{a}|$ and $|\sigma|$. Structures are denoted in bold face font, e.g. $\m a$, while the universe of a structure is denoted by the same character in italic font, e.g., $A$. The subuniverse of $\m a$ generated by a subset $G\subseteq A$ is denoted $\lb G\rangle$. We will use Big Oh notation. If $f$ and $g$ are real-valued functions defined on some subset of the real numbers, then $f\in O(g)$ and $f=O(g)$ both mean that there are positive constants $M$ and $N$ such that $|f(x)|\leq M|g(x)|$ for all $x>N$. We write $f\in \Omega(g)$ and $f=\Omega(g)$ to mean that there are positive constants $M$ and $N$ such that $|f(x)|\geq M|g(x)|$ for all $x>N$. Finally, $f\in \Theta(g)$ and $f=\Theta(g)$ mean that both $f\in O(g)$ and $f\in \Omega(g)$ hold. \subsection{Easy estimates}\label{estimates} \begin{thm}\label{basic_estimates} Let $\m a$ be an algebra. \begin{enumerate} \item[(1)] $d_{\m a^k}(n) = d_{\m a}(kn)$. \item[(2)] If $\m b$ is a homomorphic image of $\m a$, then $d_{\m b}(n) \leq d_{\m a}(n)$. \item[(3)] If $\m b$ is an expansion of $\m a$ (equivalently, if $\m a$ is a reduct of $\m b$), then $d_{\m b}(n) \leq d_{\m a}(n)$. \item[(4)] {\rm (From \cite{quick_ruskuc})} If $\m b$ is the expansion of $\m a$ obtained by adjoining all constants, then \[ d_{\m a}(n) - d_{\m a}(1)\leq d_{\m b}(n) \leq d_{\m a}(n). \] \end{enumerate} \end{thm} \begin{proof} For $(1)$, both $d_{\m a^k}(n)$ and $d_{\m a}(kn)$ represent the number of elements in a smallest size generating set for $(\m a^k)^n\cong \m a^{kn}$. For $(2)$, if $\varphi\colon \m a\to \m b$ is surjective and $G\subseteq A^n$ is a smallest size generating set for $\m a^n$, then $\varphi(G)$ is a generating set for $\m b^n$. Hence $d_{\m b}(n)\leq |\varphi(G)|\leq |G| = d_{\m a}(n)$. For $(3)$, if $G\subseteq A^n$ is a smallest size generating set for $\m a^n$, then $G$ is also a generating set for $\m b^n$. Hence $d_{\m b}(n)\leq |G| = d_{\m a}(n)$. For $(4)$, the right-hand inequality $d_{\m b}(n) \leq d_{\m a}(n)$ follows from $(3)$. Now let $G\subseteq A^n$ be a smallest size generating set for $\m b^n$ and let $H\subseteq A$ be a smallest size generating set for $\m a$. For each $a\in H$ let $\hat{a} = (a,a,\ldots,a)\in A^n$ be the associated constant tuple, and let $\widehat{H}$ be the set of these. Every tuple of $A^n$ is generated from $G$ by polynomial operations of $\m a$ acting coordinatewise, hence is generated from $G\cup \widehat{H}$ by term operations of $\m a$ acting coordinatewise. This proves $ d_{\m a}(n) \leq |G| + |H| = d_{\m b}(n) + d_{\m a}(1), $ from which the left-hand inequality follows. \end{proof} The next theorem will not be used later in the paper, except that in Section~\ref{problems} one should know that the $d$-function of a finite algebra is bounded below by a logarithmic function and above by an exponential function. \begin{thm}\label{first_bounds} If $\m a$ is a finite algebra of more than one element and $n>0$, then \[ \lceil \log_{|A|}(n) \rceil\leq d_{\m a}(n) \leq |A|^n \] and \[ \lfloor \log_{|A|}(n) \rfloor\leq h_{\m a}(n) \leq |A|^n. \] Hence $d_{\m a}(n), h_{\m a}(n)\in \Omega(\log(n))\cap 2^{O(n)}$. Moreover, \begin{enumerate} \item[(1)] $d_{\m a}(n) \in O(\log(n))$ iff $h_{\m a}(n) \in 2^{\Omega(n)}$. \item[(2)] $d_{\m a}(n) \in O(n)$ iff $h_{\m a}(n) \in \Omega(n)$, and $d_{\m a}(n) \in \Omega(n)$ iff $h_{\m a}(n) \in O(n)$. \item[(3)] $d_{\m a}(n) \in 2^{\Omega(n)}$ iff $h_{\m a}(n) \in O(\log(n))$. \end{enumerate} \end{thm} \begin{proof} It follows from Theorem~\ref{basic_estimates}~(3) that, among all algebras with universe $A$, the algebra with only projection operations for its term operations has the smallest $d$-function and the algebra with all finitary operations as term operations has the largest $d$-function. These two algebras are also extremes for the $h$-function. If $\m a$ has no nontrivial term operations, then every element of $A^n$ is a required generator, so $d_{\m a}(n) = |A|^n$. In this case, $h_{\m a}(n)= \lfloor \log_{|A|}(n)\rfloor$ for $n>0$, since $h$ is the upper adjoint of $d$. Now assume that $\m a$ has all finitary operations as term operations. The $n$-generated free algebra in the variety generated by $\m a$ is isomorphic to $\m a^{|A|^n}$ (Theorem 3 of \cite{foster}). Since the largest $n$-generated algebra in this variety is a power of $\m a$, it is also the largest $n$-generated power of $\m a$ in the variety; we obtain that $h_{\m a}(n) = |A|^n$. In this case, $d_{\m a}(n)= \lceil \log_{|A|}(n)\rceil$ for $n>0$, since $d$ is the lower adjoint of $h$. The fact that $d_{\m a}$ is the lower adjoint of $h_{\m a}$ suggests an asymmetry, in that \begin{equation}\label{adjoint1} d_{\m a}(n)\leq k\Longleftrightarrow n\leq h_{\m a}(k), \end{equation} relates an upper bound of $d_{\m a}$ to a lower bound of $h_{\m a}$. But the fact that these functions are defined between totally ordered sets allows us to rewrite (\ref{adjoint1}) as \begin{equation}\label{adjoint2} h_{\m a}(k) < n\Longleftrightarrow k < d_{\m a}(n), \end{equation} which almost exactly reverses condition (\ref{adjoint1}) on $d_{\m a}$ and $h_{\m a}$. Using this fact and the following claim, one easily verifies Items (1)--(3). \begin{clm} If $f, g\colon [a,\infty)\to \mathbb R$ are increasing functions that tend to infinity as $x$ tends to infinity, then $\lfloor f(n)\rfloor < d_{\m a}(n)\leq \lceil g(n)\rceil$ holds for all large $n$ iff $\lfloor g^{-1}(n)\rfloor \leq h_{\m a}(n) < \lceil f^{-1}(n)\rceil$ holds for all large $n$. \end{clm} \cproof Allow ``$\forall^{\infty} N$'' to stand for ``for all large $n$'', i.e., for ``$(\exists N)(\forall n>N)$''. We have \[ \begin{array}{rl} \forall^{\infty} N(d_{\m a}(n)\leq \lceil g(n)\rceil) &\Longrightarrow\;\; \forall^{\infty} N(n\leq h_{\m a}(\lceil g(n)\rceil))\\ &\Longrightarrow\;\; \forall^{\infty} N(\lfloor g^{-1}(n)\rfloor \leq h_{\m a}(\lceil g(\lfloor g^{-1}(n)\rfloor)\rceil))\\ &\Longrightarrow\;\; \forall^{\infty} N(\lfloor g^{-1}(n)\rfloor \leq h_{\m a}(n)), \end{array} \] because the monotonicity of $g$ guarantees that $\lceil g(\lfloor g^{-1}(n)\rfloor)\rceil\leq n$. The reverse implication is proved the same way, as are both implications in $\lfloor f\rfloor < d \Leftrightarrow h < \lceil f^{-1}\rceil$.~\hfill\rule{1.3mm}{3mm} \end{proof} Recall that the \emph{free spectrum} of a variety $\mathcal V$ is the function $f_{\mathcal V}(n):=|F_{\mathcal V}(n)|$ whose value at $n$ is the cardinality of the $n$-generated free algebra in $\mathcal V$. \begin{thm}\label{spec1} If $\m a$ is a nontrivial finite algebra and $f_{\mathcal V}$ is the free spectrum of the variety $\mathcal V = {\mathcal V}(\m a)$, then $ h_{\m a}(n)\leq \log_{|A|}(f_{\mathcal V}(n)) $ for $n>0$. In particular, \begin{enumerate} \item[(1)] if $f_{\mathcal V}(n)\in O(n^k)$ for some fixed $k\in\mathbb Z^+$, then $d_{\m a}(n)\in 2^{\Theta(n)}$; \item[(2)] if $f_{\mathcal V}(n)\in 2^{O(n)}$, then $d_{\m a}(n)\in \Omega(n)$. \end{enumerate} \end{thm} \begin{proof} Assume that $n>0$. The algebra $\m a^{h_{\m a}(n)}$ is $n$-generated, hence a quotient of the $n$-generated free algebra $\m f_{\mathcal V}(n)$. This proves that $|A|^{h_{\m a}(n)}\leq f_{\mathcal V}(n)$, or $h_{\m a}(n)\leq \log_{|A|}(f_{\mathcal V}(n))$. If $f_{\mathcal V}(n)\in O(n^k)$ for some fixed $k\in\mathbb Z^+$, then $\log(f_{\mathcal V}(n))\in O(\log(n))$, hence $h_{\m a}(n)\in O(\log(n))$. Theorem~\ref{first_bounds} proves that $d_{\m a}(n)\in 2^{\Omega(n)}$ holds when $h_{\m a}(n)$ is bounded like this and that $d_{\m a}(n)\in 2^{O(n)}$ holds just because $\m a$ is finite, so $d_{\m a}(n)\in 2^{\Theta(n)}$. If $f_{\mathcal V}(n)\in 2^{O(n)}$, then $\log(f_{\mathcal V}(n))\in O(n)$, hence $h_{\m a}(n)\in O(n)$. It follows from Theorem~\ref{first_bounds}~(2) that $d_{\m a}(n)\in \Omega(n)$. \end{proof} \begin{cor}\label{abelian_cor} Let $\m a$ be a nontrivial finite algebra and let $\m b$ be a nontrivial homomorphic image of $\m a^k$ for some $k$. \begin{enumerate} \item[(1)] If $\m b$ is strongly abelian (or even just strongly rectangular), then $d_{\m a}(n)\in 2^{\Theta(n)}$. \item[(2)] If $\m b$ is abelian, then $d_{\m a}(n)\in \Omega(n)$. \end{enumerate} \end{cor} \begin{proof} For (1), Theorem~5.3 of \cite{strnil} proves that a finite strongly rectangular algebra generates a variety with free spectrum bounded above by a polynomial. By Theorem~\ref{spec1}, $d_{\m a}(n)\in 2^{\Theta(n)}$ in this case. The strong abelian property is more restrictive than the strong rectangular property by Lemma~2.2~(11) of \cite{strnil}. For (2), any finite abelian algebra generates a variety $\mathcal V$ whose free spectrum satisfies $f_{\mathcal V}(n)\in 2^{O(n)}$, according to \cite{berman_mck}, so Theorem~\ref{spec1}~(2) completes the argument. \end{proof} Recall that an algebra is \emph{affine} if it is polynomially equivalent to a module. It is known that $\m a$ is affine iff $\m a$ is abelian and has a Maltsev term iff $\m a$ is abelian and has a Maltsev polynomial. \begin{thm}\label{affine_prep} If $\m a^2$ is a finitely generated affine algebra, then $d_{\m a}(n)\in O(n)$. If, moreover, $\m a$ is finite and has more than one element, then $d_{\m a}(n)\in \Theta(n)$. \end{thm} \begin{proof} The theorem is true under the weaker assumption that $\m a$ (rather than $\m a^2$) is finitely generated, provided $\m a$ is a module rather than an arbitrary affine algebra. To see this, suppose that $\m m$ is a module generated by a finite subset $G$. The set of tuples in $\m m^n$ with exactly one nonzero entry, which is taken from $G$, is a generating set for $\m m^n$ of size $\leq |G|\cdot n$. Hence $d_{\m m}(n)\in O(n)$. If, moreover, $\m m$ is finite and has more than one element, then Corollary~\ref{abelian_cor}~(2) proves that $d_{\m m}(n)\in \Omega(n)$, so $d_{\m m}(n)\in \Theta(n)$. It now follows from Theorem~\ref{basic_estimates}~(4) that if $\m a$ is an algebra that is polynomially equivalent to a finitely generated module, then $d_{\m a}(n)\in O(n)$, and $d_{\m a}(n)\in\Theta(n)$ if $\m a$ is finite and nontrivial. Unfortunately, not every finitely generated affine algebra is polynomially equivalent to a finitely generated module. But if $\m a$ is affine and $\m a^2$ is finitely generated, then the linearization $\m a^2/\Delta$ (see \cite{freese-mckenzie} pp.~114) is also finitely generated and term equivalent to a reduct of the underlying module of $\m a$. Hence when $\m a^2$ is finitely generated, then $\m a$ is polynomially equivalent to a finitely generated module, and the conclusions of the theorem hold. \end{proof} \section{General growth rates} \subsection{Growth rates of partial algebras}\label{partial_subsection} A partial algebra is a set equipped with a set of partial operations. A total algebra is considered to be a partial algebra, but, of course, some partial algebras are not total. The definitions of functions $d_{\m a}$ and $h_{\m a}$ make sense when $\m a$ is a partial algebra, as does the problem of determining growth rates of partial algebras. Theorem~\ref{basic_estimates}~(3), which relates the growth rate of an algebra to that of a reduct, holds in exactly the same form if a ``reduct of $\m b$'' is interpreted to mean an algebra $\m a$ with the same universe as $\m b$ whose basic partial operations are obtained from \emph{some} of the term partial operations of $\m b$ by possibly restricting their domains. We will learn in this subsection that a function arises as the growth rate of a partial algebra if and only if it arises as the growth rate of a total algebra. \begin{df}\label{one-point} Let $\m a = \lb A; P\rangle$ be a partial algebra with universe $A$ and a set $P$ of partial operations on $A$. The \emph{one-point completion} of $\m a$ is the total algebra whose universe is $A_{0}:=A\cup \{0\}$, where $0$ is some element not in $A$, and whose operations $P_0 = \{p_0\;|\;p\in P\}\cup \{\wedge\}$ are defined as follows. \begin{enumerate} \item[(1)] If $p\in P$ is a partial $m$-ary operation on $A$ with domain $D\subseteq A^m$, then the total operation $p_{0}\colon (A_{0})^m\to A_{0}$ is defined by \[ p_{0}(\wec{a}) = \begin{cases} p(\wec{a}) & \textrm{if $\wec{a}\in D$;}\\ 0 & \textrm{otherwise.} \end{cases} \] \item[(2)] A meet operation $\wedge$ on $A_{0}$ is defined by \[ a\wedge b = \begin{cases} a & \textrm{if $a=b$;}\\ 0 & \textrm{otherwise.} \end{cases} \] \end{enumerate} \end{df} \begin{thm}\label{partial} Let $\m a$ be a partial algebra of more than one element, and let $\m a_0$ be its one-point completion. \begin{enumerate} \item[(1)] Any generating set for $\m a^n$ is a generating set for $\m a_0^n$, and \item[(2)] Any generating set for $\m a_0^n$ contains a generating set for $\m a^n$. \end{enumerate} In particular, least size generating sets for $\m a^n$ and $\m a_0^n$ have the same size, and if $\m a^n$ or $\m a_0^n$ have any minimal generating sets, then they are the same. \end{thm} \begin{proof} In this paragraph we prove (1). If $G\subseteq A^n$ is a generating set for $\m a^n$, then as a subset of $\m a_0^n$ it will generate (in exactly the same manner) all tuples in $A_0^n$ which have no $0$'s. If $\wec{z}\in A_0^n$ is an arbitrary tuple and $a, b\in A$ are distinct, let $\wec{z}_a$ and $\wec{z}_b$ be the tuples obtained from $\wec{z}$ by replacing all $0$'s with $a$ and $b$, respectively. Then $\wec{z}_a, \wec{z}_b \in A^n$, so they are generated by $G$, and $\wec{z} = \wec{z}_a\wedge \wec{z}_b$, so $\wec{z}$ is also generated by $G$. Hence $G$ generates all of $\m a_0^n$. Now we prove (2). Assume that $H\subseteq A_0^n$ is a generating set for $\m a_0^n$. If $\wec{a}\in A_0^n$, let $Z(\wec{a})\subseteq [n]$ be the \emph{zero set} of $\wec{a}$, by which we mean the set of coordinates where $\wec{a}$ is $0$. It is easy to see that for any basic operation $F$ of $\m a_0$ it is the case that \begin{equation}\label{generation} Z(\wec{a}_1)\cup \cdots \cup Z(\wec{a}_m) \subseteq Z(F(\wec{a}_1,\ldots,\wec{a}_m)), \end{equation} since $0$ is absorbing for every basic operation. If the right-hand side is empty, then the left-hand side is empty as well; i.e., tuples with empty zero sets can be generated only by tuples with empty zero sets. Said a different way, if $H\subseteq A_0^n$ generates $\m a_0^n$, then $H\cap A^n$ suffices to generate all tuples in $A^n$. If you consider how $H$ generates elements of $A^n$ in the algebra $\m a_0^n$, it is clear that $H$ generates those elements in the algebra $\m a^n$ in exactly the same way, so $H$ is a generating set for $\m a^n$. \end{proof} \begin{cor}\label{partial_cor} If $\m a$ is a partial algebra and $\m a_0$ is its one-point completion, then $d_{\m a_0}(n)=d_{\m a}(n)$ for all $n\in\omega$. \qed \end{cor} \subsection{Growth rates of countably infinite algebras} In this section we characterize the $d$-functions of countably infinite algebras. We will see that there are a few obvious properties that these functions have, and that any function $D\colon \omega\to \omega^+$ that has these properties may be realized as a $d$-function. One obvious property of $d$-functions is that they are increasing: $m\leq n$ implies $d_{\m a}(m)\leq d_{\m a}(n)$. The $d$-function of a countably infinite algebra is an increasing function from the ordered set of natural numbers, $\omega$, to the ordered set $\omega^+=\omega\cup\{\omega\}=\{0,1,\ldots,\omega\}$, where $d_{\m a}(n)=\omega$ means that $\m a^n$ is not finitely generated. $d$-functions also have special initial values. $\m a^0$ is a 1-element algebra, so $\m a^0$ is $0$-generated if $\m a$ has a nullary term and is $1$-generated if $\m a$ has no nullary term. Thus $d_{\m a}(0)=0$ or $1$, with the cases distinguished according to whether $\m a$ has a nullary term. Finally, if $\m a$ has more than one element, then $d_{\m a}(2) > 0$, since any $0$-generated subalgebra of $\m a^2$ is contained in the diagonal and the diagonal is a proper subalgebra of $\m a^2$ when $|A|>1$. We now prove: \begin{thm}\label{big} If $D\colon \omega\to \omega^+$ \begin{enumerate} \item[(i)] is increasing, \item[(ii)] satisfies $D(0)=0$ or $1$, and \item[(iii)] satisfies $D(2) > 0$, \end{enumerate} then there is a countably infinite total algebra $\m a$ such that $d_{\m a}(n)=D(n)$ for all $n\in\omega$. \end{thm} \begin{proof} We construct a partial algebra $\m a$ such that $d_{\m a}(n) = D(n)$ for all $n\in \omega$. By Corollary~\ref{partial_cor} the one-point completion of $\m a$ (Definition~\ref{one-point}) will be a total algebra with the same growth rate. First we describe the universe of our partial algebra. Start with a countably infinite set $X$. This set will be a subset of the universe of $\m a$, and its main function is to ensure that the constructed algebra is infinite. Next, for any algebra $\m b$, $d_{\m b}(0)=0$ happens exactly when $\m b$ has a nullary term. Hence if $D(0)=0$ and we wish to represent $D$ as $d_{\m a}$ for some $\m a$, then we must ensure that $\m a$ has a nullary term. So let $Y=\{y\}$ be a singleton set. If we need our algebra to have a nullary term, we will introduce a term with value $y$. Finally, for each nonzero $n\in\omega$ where $D(n)$ is finite, let $M^{(n)} = [z_{i,j}^{(n)}]$ be an $n\times D(n)$ matrix of elements such that all entries of all $M^{(n)}$'s are different from each other and are different from the elements of $X\cup Y$. Let $Z = \{z_{i,j}^{(n)}\}$ be the set of all entries appearing in these matrices, and take $A:=X\cup Y\cup Z$ to be the universe of the partial algebra. If $D(0)=0$, then we introduce a nullary operation whose value is $y$. We may introduce more nullary operations later in the case $D(0)=0$, but if $D(0)=1$ then we do not introduce any nullary operations throughout the construction. For each nonzero $n\in\omega$ where $D(n)$ is finite and for each tuple $\wec{b}\in A^n$, introduce a $D(n)$-ary partial operation $F_{\wec{b}}$ for which $F_{\wec{b}}(M^{(n)})=\wec{b}$. This means that $F_{\wec{b}}$ has domain of size $n$, consisting of the $n$ rows of $M^{(n)}$, and that $F_{\wec{b}}(z_{i,1}^{(n)},\ldots,z_{i,D(n)}^{(n)})=b_i$ for each $i=0,1,\ldots,n$. It is worth mentioning how to interpret the instructions of the previous paragraph in the case where $n=1$ and $D(n)=0$. Here $M^{(n)}$ is defined to be a $1\times 0$ matrix, and for each $\wec{b}\in A^1=A$ we are instructed to add a partial operation $F_{\wec{b}}$ with the property that $F_{\wec{b}}(M)=\wec{b}$. One should view $F_{\wec{b}}$ as a nullary partial operation with range $\wec{b}$. Hence, in the case $(n,D(n))=(1,0)$ we are to add nullary operations naming each element of $A$. [Consider how one might interpret the instructions of the previous paragraph in the case where $n=2$ and $D(n)=0$, if such were permitted by the assumptions on $D$. We would be instructed to add nullary partial operations to $\m a$ with range $\wec{b}$ for each $\wec{b}\in A^2$. Such nullary operations do not exist for those $\wec{b}\in A^2$ off of the diagonal, so we would be unable to adhere to the instructions if we allowed $D(2)=0$. This is the place in our construction where we make use of the assumption that $D(2)>0$.] Our partial algebra is $A$ equipped with all partial operations of the type described in the previous three paragraphs. Observe that $d_{\m a}(0) = 0$ iff $\m a$ has a nullary term iff $D(0)=0$, so $d_{\m a}(0)=D(0)$. Observe that if $D(n)=\omega$ for some $n>0$, then none of the partial operations has $n$ distinct elements of $A$ in its image. Hence every tuple $\wec{b}\in A^n$ with distinct coordinates must appear in any generating set for $\m a^n$. This proves that $d_{\m a}(n)=\omega$ whenever $D(n)=\omega$. Observe that if $D(1)=0$, then we have added nullary operations to $\m a$ naming each element of $A$, so $d_{\m a}(1)=0$, too. Now we consider generating sets for $\m a^n$ when $n>0$ and $D(n)$ is finite and positive. In this case, $F_{\wec{b}}(M^{(n)}) = \wec{b}$ whenever $\wec{b}\in A^n$, so the columns of $M^{(n)}$ form a generating set of size $D(n)$ for $\m a^n$. The following claim will help us to prove that there is no smaller generating set for $\m a^n$. \begin{clm} If $n>0$ and a subset $G\subseteq A^n$ has fewer than $D(n)$ tuples whose coordinates are distinct, then the same is true for $\lb G\rangle$. \end{clm} \cproof If the claim is not true, then it must be possible to generate in one step a tuple $\wec{c}\in A^n$ whose coordinates are all distinct using other tuples, where fewer than $D(n)$ of these other tuples have the property that their coordinates are all distinct. If the partial operation used is some $F_{\wec{b}}$, $\wec{b}\in A^m$ for some $m$, and the tuples used to generate are $\wec{x}_1,\ldots, \wec{x}_{D(m)}$, then the following row equations must be satisfied. \begin{equation}\label{distinct} \small{ F_{\wec{b}}(\wec{x}_1,\ldots,\wec{x}_{D(m)})= F_{\wec{b}}\left( \left[ \begin{array}{c} x_{1,1}\\ \vdots\\ x_{n,1} \end{array} \right], \ldots, \left[ \begin{array}{c} x_{1,D(m)}\\ \vdots\\ x_{n,D(m)}\\ \end{array} \right] \right) = \left[ \begin{array}{c} c_1\\ \vdots\\ c_n \end{array} \right]=\wec{c}. } \end{equation} Considering the definition of $F_{\wec{b}}$, it is clear that the (distinct!) entries of $\wec{c}$ are among the entries of $\wec{b}$, so $m=|\wec{b}|\geq |\wec{c}|=n$. Moreover, the row equations $F_{\wec{b}}(x_{i,1},\ldots,x_{i,D(m)})=c_i$ can be solved in only one way, namely by using the appropriate row of $M^{(m)}$. This forces all entries of $[x_{i,j}]$ to be distinct. But this means there are $D(m)$ columns, $\wec{x}_j$, whose coordinates are distinct, and we assumed that there were fewer than $D(n)$ such columns. Altogether this yields that $m\geq n$ and $D(m)<D(n)$, contradicting the monotonicity of $D(n)$. The claim is proved. \hfill\rule{1.3mm}{3mm} \medskip The claim shows that $d_{\m a}(n)=D(n)$ when $n>0$ and $D(n)$ is finite and positive, since a subset $G\subseteq A^n$ of size less than $D(n)$ must have fewer than $D(n)$ tuples whose coordinates are distinct. Such a set cannot generate $\m a^n$, since the generated subuniverse $\lb G\rangle$ contains fewer than $D(n)$ tuples whose coordinates are distinct while $A^n$ contains infinitely many such tuples. \end{proof} The construction in this proof may be modified to give some information about $d$-functions of finite algebras. Namely, suppose that $D\colon \{0,1,\ldots,k\}\to \omega$ is (i)~increasing, and satisfies (ii)~$D(0)=0$ or $1$, and (iii)~$D(2)>0$. If one modifies the construction in the proof by omitting the inclusion of the set $X$ in the universe of $\m a$ and then adding only the partial operations that are nullary or of the form $F_{\wec{b}}(M^{(n)})=\wec{b}$ where $n \in \{0, 1, 2, \ldots, k\}$, then the proof shows that there is an algebra of size $|Y\cup Z| = 1+\sum_{j=0}^k j\cdot D(j)$ (finite!) such that $d_{\m a}(n)=D(n)$ for $n \in \{0, 1, 2, \ldots, k\}$. Thus there is no special behavior of $d$-functions of finite algebras on initial segments of $\omega$. \section{Kelly's Completeness Theorem}\label{Kelly_section} In Subsection~\ref{Kelly_subsection} we give a new proof of Kelly's Completeness Theorem for basic identities. The proof involves the construction of a model of a set of basic identities. In Subsection~\ref{V_subsection} we construct a simpler model by modifying the construction from the Completeness Theorem. The simpler model is not adequate for proving the Completeness Theorem, but it is exactly what we need for our investigation of growth rates. \subsection{The Completeness Theorem for basic identities}\label{Kelly_subsection} Let $\mathcal L$ be an algebraic language. Recall that an $\mathcal L$-term is \emph{basic} if it contains at most one nonnullary function symbol. An $\mathcal L$-identity $s\approx t$ is basic if both $s$ and $t$ are basic terms. If $\Sigma\cup\{\varphi\}$ is a set of basic identities, then $\varphi$ is a \emph{consequence} of $\Sigma$, written $\Sigma\models \varphi$, if every model of $\Sigma$ is a model of $\varphi$. Let $C$ be the set of constant symbols of $\mathcal L$ and let $X$ be a set of variables. The \emph{weak closure of $\Sigma$ in the variables $X$} is the smallest set $\overline{\Sigma}$ of basic identities containing $\Sigma$ for which \begin{enumerate} \item[(i)] $(t\approx t)\in \overline{\Sigma}$ for all basic $\mathcal L$-terms $t$ with variables from $X$. \item[(ii)] If $(s\approx t)\in \overline{\Sigma}$, then $(t\approx s)\in \overline{\Sigma}$. \item[(iii)] If $(r\approx s)\in \overline{\Sigma}$ and $(s\approx t)\in \overline{\Sigma}$, then $(r\approx t)\in \overline{\Sigma}$. \item[(iv)] If $(s\approx t)\in \overline{\Sigma}$ and $\gamma\colon X\to X\cup C$ is a function, then $(s[\gamma]\approx t[\gamma])\in \overline{\Sigma}$, where $s[\gamma]$ denotes the basic term obtained from $s$ by replacing each variable $x\in X$ with $\gamma(x)\in X\cup C$. \item[(v)] If $t$ is a basic $\mathcal L$-term and $(c\approx d)\in \overline{\Sigma}$ for $c, d\in C$, then $(t\approx t')\in \overline{\Sigma}$, where $t'$ is the basic term obtained from $t$ by replacing one occurrence of $c$ with $d$. \end{enumerate} These closure conditions may be interpreted as the inference rules of a proof calculus for basic identities. Therefore, write $\Sigma\vdash_X \varphi$ if $\varphi$ belongs to the weak closure of $\Sigma$ in the variables $X$. If the set $X$ is large enough, the relation $\vdash_X$ captures $\models$ for basic identities, as we will prove in Theorem~\ref{kelly}. We define $X$ to be \emph{large enough} if \begin{enumerate} \item[(a)] $X$ contains at least 2 variables, \item[(b)] $|X|\geq {\rm arity}(F)$ for any function symbol $F$ occurring in $\Sigma$, and \item[(c)] $|X|$ is at least as large as the number of distinct variables occurring in any identity in $\Sigma\cup\{\varphi\}$. \end{enumerate} Call $\Sigma$ \emph{inconsistent relative to $X$} if $\Sigma\vdash_X x\approx y$ for distinct $x, y\in X$ and large enough $X$. Otherwise $\Sigma$ is \emph{consistent relative to $X$}. \begin{thm}[David Kelly, \cite{kelly}] \label{kelly} Let $\Sigma\cup \{\varphi\}$ be a set of basic identities and $X$ be a set of variables that is large enough. If $\Sigma$ is consistent relative to $X$, then $\Sigma\vdash_X \varphi$ if and only if $\Sigma\models \varphi$. \end{thm} Kelly's theorem is a natural restriction of Birkhoff's Completeness Theorem for equational logic to the special case of basic identities. However, it is in general undecidable for finite $\Sigma\cup\{\varphi\}$ whether $\Sigma\vdash \varphi$ using Birkhoff's inference rules, while it is decidable for basic identities using Kelly's restricted rules.\footnote{% The reason that $\Sigma\vdash_X \varphi$ is decidable with Kelly's inference rules when $\Sigma\cup \{\varphi\}$ is finite is that deciding $\Sigma\vdash_X \varphi$ amounts to generating $\overline{\Sigma}$. If $\mathcal L$ is the language whose function and constant symbols are those occurring in $\Sigma\cup \{\varphi\}$, $X$ is a minimal (finite) set of variables that is large enough, and $\mathcal T$ is defined to be the set of basic $\mathcal L$-terms in the variables $X$, then generating $\overline{\Sigma}$ amounts to generating an equivalence relation on the finite set $\mathcal T$ using Kelly's inference rules.} In the proof we use a variation of Kelly's Rule (iv): rather than use functions $\gamma\colon X\to X\cup C$ for substitutions we will use functions $\Gamma\colon X\cup C\to X\cup C$ whose restriction to $C$ is the identity. (That is, we replace $\gamma$ with $\Gamma := \gamma\cup {\rm id}|_C$.) \begin{lm}\label{kelly_lm} If $\Sigma\vdash_X x\approx h$ for some basic term $h$ in which $x$ does not occur, then $\Sigma$ is inconsistent relative to any set $X$ containing a variable other than $x$. \end{lm} \begin{proof} Append to a $\Sigma$-proof of $x\approx h$ the formulas $(y\approx h)$ for some $y\in X\setminus \{x\}$ (Rule (iv)); $(h\approx y)$ (Rule (ii)); and $(x\approx y)$ (Rule (iii)). \end{proof} \begin{proof}[Proof of Theorem~\ref{kelly}] Kelly's inference rules are sound, since they are instances of Birkhoff's inference rules for equational logic. Hence $\Sigma\vdash_X \varphi$ implies $\Sigma\models \varphi$ for any $X$. Now assume that $\Sigma{\not{\vdash_X}}\varphi$, where $X$ is large enough and $\Sigma$ is consistent relative to $X$. We construct a model of $\Sigma\cup \{\neg\varphi\}$ to show that $\Sigma{\not{\models}}\varphi$. Let $\mathcal T$ be the set of basic $\mathcal L$-terms in the variables $X$, and let $\equiv$ be the equivalence relation on $\mathcal T$ defined by Kelly provability: i.e., $s\equiv t$ if and only if $\Sigma\vdash_X s\approx t$. Write $[t]$ for the $\equiv$-class of $t$. Now extend $\mathcal T$ to a set ${\mathcal T}_0 = {\mathcal T}\cup \{0\}$ where $0$ is a new symbol, and extend $\equiv$ to this set by taking the equivalence class of $0$ to be $\{0\}$. The universe of the model will be the set $M:={\mathcal T}_0/{\equiv}$ of equivalence classes of ${\mathcal T}_0$ under $\equiv$. We interpret a constant symbol $c$ as the element $c^{\m m}:=[c]\in M$. Now let $F$ be an $m$-ary function symbol for some $m>0$. The natural idea for interpreting $F$ as an $m$-ary operation on this set is to define $F^{\m m}([a_1],\ldots,[a_m]) = [F(a_1,\ldots,a_m)]$. However, this does not work, since $F(a_1,\ldots,a_m)$ will not be a basic term unless all the $a_i$'s belong to $X\cup C$. Nevertheless, we shall follow this idea as far as it takes us, and when we cannot apply it to assign a value to $F^{\m m}([a_1],\ldots,[a_m])$ we shall assign the value $[0]$. Choose and fix a well-order $<$ of the set $C$ of constant symbols of $\mathcal L$. Let $\mathcal I$ be the set of injective partial functions $\imath\colon M\to X\cup C$ that satisfy the following conditions: \begin{enumerate} \item[(1)] If a class $[t]\in M$ in the domain of $\imath$ contains a constant symbol, $c\in C$, then $\imath[t] = d$ where $d\in C$ is the least element in $[t]\cap C$ under $<$. \item[(2)] If a class $[t]$ in the domain of $\imath$ contains a variable, $x\in X$, then $\imath[t] = x$. \item[(3)] If a class $[t]$ in the domain of $\imath$ fails to contain a variable or constant symbol, then $\imath[t]\in X$. \end{enumerate} According to Lemma~\ref{kelly_lm}, the consistency of $\Sigma$ implies that any class $[t]$ contains at most one variable, and if $[t]$ contains a constant symbol, then $[t]$ contains no variable. Hence there is no ambiguity in conditions (1) and (2). If $S\subseteq M$ has size at most $|X|$, then $S$ is the domain of some $\imath\in \mathcal I$. If $S\subseteq M$ and a class $[t]\in S$ contains a variable $x$, then call $x$ a \emph{fixed} variable of $S$. Any other variable is an \emph{unfixed} variable of $S$. Now we define how to interpret an $m$-ary function symbol $F$ as an $m$-ary operation on the set $M$. Choose any $([a_1],\ldots,[a_m])\in M^m$, then choose $\imath\in \mathcal I$ that is defined on $S:=\{[a_1],\ldots,[a_m]\}$. Note that $f:=F(\imath[a_1],\ldots,\imath[a_m])$ is a basic term, since it is a function symbol applied to elements of $X\cup C$. We refer to this term $f$ to define $F^{\m m}([a_1],\ldots,[a_m])$. \begin{enumerate} \item[{\rm Case 1.}] (The class $[f]$ contains a term $h$ whose only variables are among the fixed variables of $S$.) Define $F^{\m m}([a_1],\ldots,[a_m]) = [f]$. \item[{\rm Case 2.}] ($[f]$ contains a variable.) If $x$ is a variable in $[f]$, then $\Sigma\vdash_X f\approx x$. Since $\Sigma$ is consistent, Lemma~\ref{kelly_lm} proves that $x$ must occur in $f$, i.e., $x=\imath[a_k]$ for some $k$. Hence \[ \Sigma\vdash_X F(\imath[a_1],\ldots,\imath[a_m])\approx \imath[a_k] \] for some $k$. In this case we define $F^{\m m}([a_1],\ldots,[a_m])=[a_k]\;(=\imath^{-1}(x)$.) \item[{\rm Case 3.}] (The remaining cases.) Define $F^{\m m}([a_1],\ldots,[a_m]) = [0]$. \end{enumerate} Before proceeding, we point out that there is overlap in Cases 1 and 2, but no conflict in the definition of $F^{\m m}([a_1],\ldots,[a_m])$. If $[f]$ contains a term $h$ whose variables are fixed variables of $S$ and $[f]$ also contains a variable $x$, then $\Sigma\vdash_X f\approx x$ and $\Sigma\vdash_X h\approx x$. The consistency of $\Sigma$ forces $x$ to be a common variable of $f$ and $h$, and (since only fixed variables of $S$ occur in $h$) to be a fixed variable of $S$. In this situation, Case 1 defines $F^{\m m}([a_1],\ldots,[a_m])=[f]=[x]$ while Case 2 defines $F^{\m m}([a_1],\ldots,[a_m])=\imath^{-1}(x)=[x]$. \begin{clm}\label{well-defined} $F^{\m m}\colon M^m\to M$ is a well-defined function. \end{clm} \cproof Choose $([a_1],\ldots,[a_m])\in M^m$ and define $S=\{[a_1],\ldots,[a_m]\}$. There exist elements of $\mathcal I$ defined on $S$, because this set has size $\leq {\rm arity}(F)\leq |X|$. Suppose that $\imath, \jmath\in \mathcal I$ are both defined on this set. Let $f = F(\imath[a_1],\ldots,\imath[a_m])$ and $g = F(\jmath[a_1],\ldots,\jmath[a_m])$. To show that $F^{\m m}([a_1],\ldots,[a_m])$ is uniquely defined it suffices to show that the same value is assigned whether we refer to the term $f$ or the term $g$. In all cases of the definition of $F^{\m m}([a_1],\ldots,[a_m])$, the assigned value depends only on the term $f = F(\imath[a_1],\ldots,\imath[a_m]) = F(\imath|_S[a_1],\ldots,\imath|_S[a_m])$. Thus, to complete the proof of Claim~\ref{well-defined}, we may replace both $\imath$ and $\jmath$ by $\imath|_S$ and $\jmath|_S$ and assume that $\imath$ and $\jmath$ have domain $S$. Now $\imath$ and $\jmath$ are injective functions from $S$ into $X\cup C$, and $\imath[t] = \jmath[t]$ whenever $[t]\in S$ and $[t]$ contains a constant symbol or a fixed variable of $S$. When $\imath[t]\neq \jmath[t]$, then both are unfixed variables of $S$. In this situation, there is a function $\Gamma\colon X\cup C\to X\cup C$ that is the identity on $C$ and on the fixed variables of $S$ for which $\jmath = \Gamma\circ \imath$. Hence $f[\Gamma]=g$ and, if $h$ is a term whose only variables are fixed variables of $S$, then $h[\Gamma] = h$. \begin{enumerate} \item[{\rm Case 1.}] ($[f]$ contains a term $h$ whose only variables are among the fixed variables of $S$.) Here $ \Sigma\vdash_X f=F(\imath[a_1],\ldots,\imath[a_m])\approx h. $ Append to a $\Sigma$-proof of $f\approx h$ the formula $f[\Gamma]\approx h[\Gamma]$ (Rule (iv)). Since $f[\Gamma]=g$ and $h[\Gamma]=h$, this is a proof of $g\approx h$. Next append $h\approx g$ (Rule (ii)) and $f\approx g$ (Rule (iii)). We conclude that $[f]=[g]$, so the value $[f]$ assigned to $F^{\m m}([a_1],\ldots,[a_m])$ using $\imath$ is the same as the value $[g]$ assigned using $\jmath$. \item[{\rm Case 2.}] ($[f]$ contains a variable.) If $x\in X$ is a variable in $[f]$, then $x=\imath[a_k]$ for some $k$ and $ \Sigma\vdash_X F(\imath[a_1],\ldots,\imath[a_m])\approx \imath[a_k] $ for this $k$. Append to a $\Sigma$-proof of $f\approx x$ the formula $f[\Gamma]\approx x[\Gamma]$ (Rule (iv)). Since $f[\Gamma]=g$ and $x[\Gamma]=\jmath[a_k]$, we conclude that $ \Sigma\vdash_X F(\jmath[a_1],\ldots,\jmath[a_m])\approx \jmath[a_k] $ for the same $k$. Whether we use $\imath$ or $\jmath$ we get $F^{\m m}([a_1],\ldots,[a_m])=[a_k]$. \item[{\rm Case 3.}] (The remaining cases.) In Case 1 we showed that $[f]=[g]$ while in Case 2 we showed that if $x$ is a variable in $[f]$, then $x[\Gamma]$ is a variable in $[g]$; together these show that if $[g]$ does not contain a variable nor a term whose only variables are among the fixed variables of $S$, then the same is true of $[f]$. This argument works with $f$ and $g$ interchanged, so the remaining cases are those where both $[f]$ and $[g]$ contain no variables nor terms whose only variables are among the fixed variables of $S$. Whether we use $\imath$ or $\jmath$, we get $F^{\m m}([a_1],\ldots,[a_m]) = [0]$. \end{enumerate} \hfill\rule{1.3mm}{3mm} \medskip $\m m$ is defined. We now argue that $\m m$ is a model of $\Sigma$. Choose an identity $(s\approx t)\in\Sigma$. If $s$ is an $n$-ary function symbol $F$ followed by a sequence $\alpha\colon [n]\to X\cup C$ of length $n$ consisting of variables and constant symbols, then let $F[\alpha]$ be an abbreviation for $s$. If $s$ is a variable or constant symbol, then $s$ determines a function $\alpha\colon [1]\to X\cup C\colon 1\mapsto s$, so abbreviate $s$ by $\diamondsuit[\alpha]$. We will, in fact, write $s$ as $F[\alpha]$ in either case, but will remember that $F$ may equal the artificially introduced symbol $\diamondsuit$. The identity $s\approx t$ takes the form $F[\alpha]\approx G[\beta]$. A valuation in $\m m$ is a function $v\colon X\cup C \to M$ satisfying $v(c)=c^{\m m}=[c]$ for each $c\in C$. To show that $\m m$ satisfies $F[\alpha]\approx G[\beta]$ we must show that $F^{\m m}[v\circ \alpha] = G^{\m m}[v\circ \beta]$ for any valuation $v$. Choose $\imath\in \mathcal I$ that is defined on the set $\im(v\circ \alpha)\cup \im(v\circ \beta)$. This is possible, since we assume that $|X|$ is at least as large as the number of distinct variables in the identity $F[\alpha]\approx G[\beta]\in \Sigma$. The values of $F^{\m m}[v\circ \alpha]$ and $G^{\m m}[v\circ \beta]$ are defined in reference to the terms $f:=F[\imath\circ v\circ \alpha]$ and $g:=G[\imath\circ v\circ \beta]$ respectively. \begin{clm}\label{f=g} $[f]=[g]$. \end{clm} \cproof Observe that $(\imath\circ v)(c) = \imath[c] = d$, where $d\in C$ is the $<$-least constant symbol in the class $[c]$. If $\Gamma\colon X\cup C\to X\cup C$ is a function that agrees with $(\imath\circ v)$ on the variables in $\im(\alpha)\cup \im(\beta)$, but is the identity on $C$, then applications of Rule (v) show that $\Sigma\vdash_X F[\imath\circ v\circ \alpha]\approx F[\Gamma\circ \alpha]$ and $\Sigma\vdash_X G[\imath\circ v\circ \beta]\approx G[\Gamma\circ \beta]$. From Rule~(iv), the fact that $(F[\alpha]\approx G[\beta])\in\Sigma$ implies that $\Sigma\vdash_X F[\Gamma\circ \alpha]\approx G[\Gamma\circ \beta]$. Hence \[ \Sigma\vdash_X f=F[\imath\circ v\circ \alpha]\approx F[\Gamma\circ \alpha]\approx G[\Gamma\circ \beta]\approx G[\imath\circ v\circ \beta]=g, \] from which we get $[f]=[g]$. \hfill\rule{1.3mm}{3mm} \medskip We conclude the argument that $\m m$ satisfies $F[\alpha]\approx G[\beta]$ as follows. \begin{enumerate} \item[{\rm Case 1.}] ($[f]=[g]$ contains a term $h$ whose only variables are among the fixed variables of $S$.) In this case $F^{\m m}[v\circ \alpha]=[f]=[g]=G^{\m m}[v\circ \beta]$. \item[{\rm Case 2.}] ($[f]=[g]$ contains a variable.) If $[f]=[x]=[g]$, then $F^{\m m}[v\circ \alpha]=\imath^{-1}(x)=G^{\m m}[v\circ \beta]$. \item[{\rm Case 3.}] (The remaining cases with $[f]=[g]$.) $F^{\m m}[v\circ \alpha]=[0]=G^{\m m}[v\circ \beta]$. \end{enumerate} To complete the proof of the theorem we must show that $\m m$ does not satisfy $\varphi$. Suppose $\varphi$ has the form $F[\alpha]\approx G[\beta]$. Let $v$ be the canonical valuation \[ X\cup C \to M\colon x\mapsto [x], c\mapsto [c]. \] Choose $\imath\in\mathcal I$ that is defined on $\im(v\circ \alpha)\cup \im(v\circ \beta)$. It follows from the definitions that $\imath\circ v\colon X\cup C\to X\cup C$ fixes every variable in $\im(\alpha)\cup \im(\beta)$, while $(\imath\circ v)(c) = d$ is the $<$-least constant symbol in the class of $c$. If $\Gamma$ is the identity function on $X\cup C$, then just as in the proof of Claim~\ref{f=g}, we obtain $\Sigma\vdash_X f=F[\imath\circ v\circ \alpha]\approx F[\Gamma\circ \alpha] = F[\alpha]$ and $\Sigma\vdash_X g=G[\imath\circ v\circ \beta]\approx G[\Gamma\circ\beta]=G[\beta]$. Now $[f]$ contains a term $h:=F[\alpha]$ whose only variables are among the fixed variables of $S=\im(\alpha)$, so we are in Case 1 of the definition of $F^{\m m}$. Hence $F^{\m m}(v\circ \alpha) = [F[\alpha]]$, and similarly $G^{\m m}(v\circ \beta) = [G[\beta]]$. Part of our assumption about $\varphi=(F[\alpha]\approx G[\beta])$ is that $\Sigma\not\vdash_X \varphi$, so $[F[\alpha]]$ and $[G[\beta]]$ are distinct elements of $M$. Therefore, $v$ witnesses that $\m m$ does not satisfy $\varphi$. \end{proof} Theorem~\ref{kelly} establishes that if $X$ and $Y$ are two sets of variables that are large enough, then $\Sigma\vdash_X\varphi$ holds iff $\Sigma\vdash_Y\varphi$, and hence $\Sigma$ is consistent relative to $X$ if and only if it is consistent relative to $Y$. Now that the theorem is proved, we drop the subscript in $\vdash_X$ and the phrase ``relative to $X$'' when writing about provability. \subsection{The model $\m v$}\label{V_subsection} Later in the paper we prove theorems about finite algebras realizing a set $\Sigma$ of basic identities. For this, we need to be able to construct finite models of $\Sigma$. The model constructed in Theorem~\ref{kelly} may be infinite, so we explain how to produce finite models. \begin{df}\label{V} Let $\Sigma$ be a set of basic identities in a language $\mathcal L$ whose set of constant symbols is $C$. Let $Y$ be a set of variables, $z$ a variable not in $Y$, and $X$ a large enough set of variables containing $Y\cup \{z\}$. Let $V$ be the subset of the model $\m m$ constructed in the proof of Theorem~\ref{kelly} consisting of \[ \{[y]\;|\;y\in Y\}\cup \{[c]\;|\;c\in C\}\cup \{[0]\}. \] Write $[Y]$ for $\{[y]\;|\;y\in Y\}$ and $[C]$ for $\{[c]\;|\;c\in C\}$. As in the proof of Theorem~\ref{kelly}, let $<$ be a well-ordering of $C$. If $F$ is an $m$-ary function symbol of $\mathcal L$ and $([a_1],\ldots,[a_m])\in V^n$, then let \begin{enumerate} \item[(1)] $\imath [a_k] = d$ if $[a_k]\in [C]$ and $d$ is the $<$-least element of $[a_k]\cap C$, \item[(2)] $\imath [a_k] = y$ if $[a_k]=[y]\in [Y]$, and \item[(3)] $\imath [a_k] = z$ if $[a_k]=[0]$. \end{enumerate} Define $ F^{\m v}([a_1],\ldots,[a_m]) = [t] $ if there exists $t\in Y\cup C$ such that \begin{equation}\label{V_def_term} \Sigma\vdash F(\imath [a_1],\ldots,\imath [a_m]) \approx t, \end{equation} and define $F^{\m v}([a_1],\ldots,[a_m]) = [0]$ if there is no such $t$. $\m v$ is the algebra with universe $V$ equipped with all operations of the form $F^{\m v}$. \end{df} \begin{thm}\label{small_models} $\m v$ is a model of $\Sigma$. \end{thm} \begin{proof} Let $F[\alpha]\approx G[\beta]$ be an identity in $\Sigma$, and let $v\colon X\cup C\to V$ be a valuation. We must show that $F^{\m v}(v\circ \alpha) = G^{\m v}(v\circ \beta)$. The function $v$ is also a valuation in $\m m$, because $V\subseteq M$. Since $\m m$ is a model of $\Sigma$, we get $F^{\m m}[v\circ \alpha] = G^{\m m}[v\circ \beta]$. Choose $\imath\in \mathcal I$ defined on the set $\im(v\circ\alpha)\cup \im(v\circ\beta)$ such that $\imath[0]=z$, if $[0]$ is in this set. Let $f = F[\imath\circ v\circ \alpha]$ and $g = G[\imath\circ v\circ \beta]$. As in the proof of Claim~\ref{f=g}, if $\Gamma\colon X\cup C\to X\cup C$ is the identity on $C$ and agrees with $\imath\circ v$ on the variables in $\im(\alpha)\cup \im(\beta)$, then $\Sigma\vdash f \approx F[\Gamma\circ \alpha]\approx G[\Gamma\circ \beta] \approx g$. The term $F(\imath [a_1],\ldots,\imath [a_m])$ from line (\ref{V_def_term}) is none other than $f$. $F^{\m v}[v\circ \alpha] = [t]$ for some $t\in Y\cup C$ if and only if $\Sigma\vdash f = F(\imath [a_1],\ldots,\imath [a_m]) \approx t$. But since $\Sigma\vdash f\approx g$ we also get $G^{\m v}[v\circ \beta]=[t]$. This shows that $F^{\m v}[v\circ \alpha]$ and $G^{\m v}[v\circ \beta]$ are equal when at least one of them is not $[0]$. Of course, they are also equal when both of them equal $[0]$, so $F^{\m v}(v\circ \alpha) = G^{\m v}(v\circ \beta)$. \end{proof} \begin{cor}\label{small_models_cor} If $\Sigma$ is a consistent set of basic identities in a language whose set of constant symbols is $C$, then $\Sigma$ has models of every cardinality strictly exceeding $|C|$. \end{cor} \begin{proof} Vary the size of $Y$ in the definition of $\m v$, and use Theorem~\ref{small_models}. \end{proof} Corollary~\ref{small_models_cor} is close to the best possible result about sizes of models of a set of basic identities, as the next example shows. \begin{exmp}\label{example_constants} Let $C$ be a set of constant symbols and let ${\mathcal B} = \{B_{c,d}\;|\;c, d\in C, c\neq d\}$ be a set of binary function symbols. Let \[ \Sigma = \{B_{c,d}(c,x)\approx x, B_{c,d}(d,x)\approx d\;|\;c, d\in C, c\neq d\}. \] $\Sigma$ is a consistent set of basic identities, since if $A$ is any set containing $C$ we can interpret each $c\in C$ in $A$ as itself and each $B_{c,d}$ on $A$ by letting $B_{c,d}^{\m a}(c,y)=y$ and $B_{c,d}^{\m a}(x,y)=x$ if $x\neq c$. If $\m m$ is any model of $\Sigma$ and $c^{\m m} = d^{\m m}$ for some $c, d\in C$, then the identity function $B^{\m m}_{c,d}(c^{\m m},x)$ equals the constant function $B^{\m m}_{c,d}(d^{\m m},x)$, so $|M|=1$. Thus elements of $C$ must have distinct interpretations in any nontrivial model of $\Sigma$, implying that nontrivial models have size at least $|C|$. \end{exmp} \section{Restrictive $\Sigma$}\label{examples} Call a set $\Sigma$ of basic identities \emph{nonrestrictive} if, whenever $\m a$ is an algebra, there is an algebra $\m b$ realizing $\Sigma$ such that $d_{\m b}(n)=d_{\m a}(n)$. Otherwise $\Sigma$ is \emph{restrictive}. Call $\Sigma$ \emph{nonrestrictive for finite algebras} if, whenever $\m a$ is a finite algebra, there is a \underline{finite} algebra $\m b$ realizing $\Sigma$ such that $d_{\m b}(n)=d_{\m a}(n)$. Otherwise $\Sigma$ is \emph{restrictive for finite algebras}. It is possible for $\Sigma$ to be nonrestrictive, yet restrictive for finite algebras. The set $\Sigma$ from Example~\ref{example_constants} has this property when the set of constants is infinite (cf.\ Remark~\ref{realize}). But the concepts defined in the two preceding paragraphs are close enough that the arguments of this section apply equally well to both of them. We will see that if only finitely many constant symbols appear in $\Sigma$, then $\Sigma$ is restrictive if and only if it is restrictive for finite algebras. Both are equivalent to the property that $\Sigma$ entails the existence of a pointed cube term. Recall from the introduction that an $m$-ary, $p$-pointed, $k$-cube term for the variety axiomatized by $\Sigma$ is an $m$-ary term $F(x_1,\ldots,x_m)$ for which there is a $k\times m$ matrix $M=[y_{i,j}]$ of variables and constant symbols, where every column contains a symbol different from $x$, such that $\Sigma$ proves the identities \[ F(\wec{y}_1,\ldots,\wec{y}_m)= F\left( \left[\begin{array}{c} y_{1,1}\\ \vdots\\ y_{k,1}\\ \end{array}\right], \cdots, \left[\begin{array}{c} y_{1,m}\\ \vdots\\ y_{k,m}\\ \end{array}\right]\right) \approx \left[\begin{array}{c} x\\ \vdots\\ x\\ \end{array}\right]. \] In any nontrivial situation the parameters are constrained by $m, k\geq 2$ and $p\geq 0$. In Subsection~\ref{pointed_subsection} we prove that if $\Sigma$ is restrictive, then it entails the existence of a pointed cube term. The converse is proved in Subsection~\ref{processing}, by showing that an algebra with a pointed cube term whose $d$-function assumes only finite values has growth rate that is bounded above by a polynomial. In particular, it is shown that a finite algebra $\m a$ with a $1$-pointed $k$-cube term satisfies $d_{\m a}(n)\in O(n^{k-1})$. In Subsection~\ref{poly_growth} we give an example of a $3$-element algebra with a $1$-pointed $k$-cube term whose growth rate satisfies $d_{\m a}(n)\in \Theta(n^{k-1})$, showing that the preceding estimate is sharp. In Subsection~\ref{avoid} we show that any function $D\colon \mathbf{Z}^+\to\mathbf{Z}^{\geq 0}$ that occurs as the $d$-function of an algebra with a pointed cube term also occurs as the $d$-function of an algebra that does not have a pointed cube term. In Subsection~\ref{exp} we describe one way of showing that an algebra has an exponential growth rate, and we use it to exhibit a variety containing a chain of finite algebras $\m a_1\leq \m a_2\leq \cdots$, each one a subalgebra of the next, where $\m a_i$ has logarithmic growth when $i$ is odd and exponential growth when $i$ is even. \subsection{Restrictive $\Sigma$ forces a pointed cube term}\label{pointed_subsection} Let $\Sigma$ be a set of basic identities in a language $\mathcal L$ whose set $C$ of constant symbols is finite. Given an algebra $\m a$ in a language disjoint from $\Sigma$ we construct another algebra $\m a_{\Sigma}$ which realizes $\Sigma$, where $\m a_{\Sigma}$ is finite if $\m a$ is. For the first step, let $[C] = \{[c_1],\ldots,[c_p]\}$ be the same set of equivalence classes denoted by $[C]$ in Definition~\ref{V}. These classes represent the different $\Sigma$-provability classes of constant symbols. If there are $p$ such classes, then apply the one-point completion construction of Subsection~\ref{partial_subsection} $p+1$ times to $\m a$ to produce a sequence $\m a$, $\m a_{z_1}$, $\m a_{z_1, z_2}$, \ldots, ending at $\m a_{z_1,\ldots,z_p,0}$. This is an algebra whose universe is the disjoint union of $A$ and $\{z_1,\ldots,z_p, 0\}$. $\m a_{\Sigma}$ will be an expansion of $\m a_{z_1,\ldots,z_p,0}$ obtained by merging the latter algebra with the model $\m v$ introduced in Definition~\ref{V}. Let $Y$ be a set of variables satisfying $|Y|=|A|$, and let $[Y] = \{[y]\;|\;y\in Y\}$ be the set of equivalence classes also denoted by $[Y]$ in Definition~\ref{V}. The universe of $\m v$ is the disjoint union $V = [Y]\cup [C]\cup \{[0]\}$. Let $\varphi\colon [Y]\to A$ be a bijection. Extend this to a bijection from $V=[Y]\cup [C]\cup\{[0]\}$ to $A\cup \{z_1,\ldots,z_p\}\cup\{0\}$ by defining $\varphi([c_i]) = z_i$ and $\varphi([0])=0$. Now $\varphi$ is a bijection from the universe of $\m V$ to the universe of $\m a_{z_1,\ldots,z_p,0}$. Use this bijection to transfer the operations of $\m v$ over to $\m a_{z_1,\ldots,z_p,0}$ to create $\m a_{\Sigma}$. Specifically, the interpretation of the constant symbol $c_i$ in $\m a_{\Sigma}$ will be $z_i$, and if $F$ is an $m$-ary function symbol of $\mathcal L$, then \begin{equation}\label{Ops} F^{\m a_{\Sigma}}(x_1,\ldots,x_m):= \varphi(F^{\m v}(\varphi^{-1}(x_1),\ldots,\varphi^{-1}(x_m))) \end{equation} will be the interpretation of the symbol $F$ in $\m a_{\Sigma}$. $\m a_{\Sigma}$ is the expansion of $\m a_{z_1,\ldots,z_p,0}$ by all constant operations $c_i^{\m a_\Sigma}$ and all operations of the form (\ref{Ops}). Under this definition the function $\varphi$ is an isomorphism from $\m v$ to the $\mathcal L$-reduct of $\m a_{\Sigma}$. \begin{lm}\label{nonrestrictive} Let $\m a$ be an algebra with more than one element and let $\Sigma$ be a set of basic identities involving finitely many constant symbols. Let $\mathcal V$ be the variety axiomatized by $\Sigma$. The following statements about an integer $k\geq 2$ are equivalent. \begin{enumerate} \item[(1)] $\mathcal V$ has a pointed $k$-cube term. \item[(2)] For any $n\geq k$, $A^n_{\Sigma}\setminus A^n$ generates $\m a^n_{\Sigma}$. \item[(3)] For any $n\geq k$, there is a generating set $G(n)$ of $\m a^n_{\Sigma}$ such that $G(n)\cap A^n$ does not generate $\m a^n$. \item[(4)] $\mathcal V$ has a pointed $k$-cube term of the form $F(x_1,\ldots,x_m)$, where $m\geq 2$, $F$ is a function symbol occurring in $\Sigma$, and the variables $x_1,\ldots,x_m$ are distinct. \end{enumerate} \end{lm} \begin{proof} [$(1)\Rightarrow(2)$] Let $F(x_1,\ldots,x_m)$ be a pointed $k$-cube term of the variety axiomatized by $\Sigma$. There is a $k\times m$ matrix $M=[y_{i,j}]$ of variables and $\mathcal L$-constant symbols, where every column contains a symbol different from $x$, such that $\Sigma$ proves the identities \begin{equation}\label{Pidentities} F(\wec{y}_1,\ldots,\wec{y}_m)= F\left( \left[\begin{array}{c} y_{1,1}\\ \vdots\\ y_{k,1}\\ \end{array}\right], \cdots, \left[\begin{array}{c} y_{1,m}\\ \vdots\\ y_{k,m}\\ \end{array}\right]\right) \approx \left[\begin{array}{c} x\\ \vdots\\ x\\ \end{array}\right]. \end{equation} Choose any tuple $\wec{a}\in A^n_{\Sigma}$. Using the row identities of (\ref{Pidentities}), solve the equation $F(\wec{b}_1,\ldots,\wec{b}_m)=\wec{a}$ for the $\wec{b}_i$'s, row by row, according to the following rules. In the $i$-th row, \begin{enumerate} \item[(a)] if $y_{i,j}=x$, then let $b_{i,j}=a_i$. \item[(b)] if $y_{i,j} = c_r$ is a constant symbol, then let $b_{i,j}=z_r$ be its interpretation in $\m a_{\Sigma}$. \item[(c)] if $y_{i,j}$ is a variable different from $x$, then let $b_{i,j}=0$. \end{enumerate} Under these choices, $\wec{b}_i\in A_{\Sigma}^n$ for all $i$ and $F(\wec{b}_1,\ldots,\wec{b}_m)=\wec{a}$. Moreover, since each column $\wec{y}_i$ in (\ref{Pidentities}) has a symbol different from $x$, it follows from (a)--(c) that each $\wec{b}_i$ has a coordinate value that is in the set $\{z_1,\ldots,z_p,0\}$. Hence $\wec{b}_i\in A_{\Sigma}^n\setminus A^n$ for all $i$. This shows that the arbitrarily chosen tuple $\wec{a}\in A^n_{\Sigma}$ lies in the subalgebra $\m a^n_{\Sigma}$ that is generated by $A_{\Sigma}^n\setminus A^n$. [$(2)\Rightarrow(3)$] Let $G(n) = A^n_{\Sigma}\setminus A^n$. [$(3)\Rightarrow(4)$] Let $G(n)$ be the generating set for $\m a^n_{\Sigma}$ that is guaranteed by Item~(3). Since $G(n)\cap A^n$ does not generate $\m a^n$, it follows from Theorem~\ref{partial} that $(A^n_{\Sigma}\setminus A^n)\cup G(n)$ is not a generating set for $\m a^n_{z_1,\ldots,z_p,0}$. Let $S$ be the proper subuniverse of $\m a^n_{z_1,\ldots,z_p,0}$ that is generated by $(A^n_{\Sigma}\setminus A^n)\cup G(n)$. Since $S$ contains $G(n)$, which generates $\m a^n_{\Sigma}$, and contains the interpretations of the $\mathcal L$-constants, it cannot be closed under the interpretations of the function symbols of $\mathcal L$. Hence there is a tuple $\wec{a}\notin S$, an $m$-ary function symbol $F$, and $m$ tuples $\wec{b}_1,\ldots, \wec{b}_m\in S$ such that $F^{\m a_{\Sigma}}(\wec{b}_1,\ldots,\wec{b}_m)=\wec{a}$. Necessarily $\wec{a}\in A^n$. Using the isomorphism $\varphi$ from $\m v$ to the $\mathcal L$-reduct of $\m a_{\Sigma}$, we obtain that there is a tuple $\wec{y} = \varphi^{-1}(\wec{a})\in \varphi^{-1}(A^n)=[Y]^n$ and tuples $\wec{v}_i = \varphi^{-1}(\wec{b}_i)\neq \wec{y}$ such that $ F^{\m v}(\wec{v}_1,\ldots,\wec{v}_m)=\wec{y}. $ Since $\wec{v}_1\neq \wec{y}$, there is a coordinate $\ell$ where these tuples differ. In the $\ell$-th coordinate we have $F^{\m v}([v_{{\ell},1}],\ldots,[v_{{\ell},m}])=[y_{\ell}]$ for some variable $y_{\ell}\in Y$ and some elements $v_{\ell,j}\in Y\cup C\cup \{0\}$ with $[v_{{\ell},1}]\neq [y_{\ell}]$. By the definition of $\m v$, \begin{equation}\label{cube1} \Sigma\vdash F(\imath [v_{\ell,1}],\ldots,\imath [v_{\ell,m}])\approx y_{\ell}, \end{equation} where $\imath [v_{\ell,j}]=v_{\ell,j}$ when $v_{\ell,j}\in Y$, $\imath[v_{\ell,j}]$ is a constant $\Sigma$-provably equivalent to $v_{\ell,j}$ when $v_{\ell,j}\in C$, and $\imath [v_{\ell,j}]=z$ is a variable not in $Y$ when $v_{\ell,j}=0$. Since $[v_{\ell,1}]\neq [y_{\ell}]$, we have $v_{\ell,1}\neq y_{\ell}$. After renaming variables, (\ref{cube1}) can be rewritten as \[ \Sigma\vdash F(m_{1,1},\ldots,m_{1,m})\approx x, \] where each $m_{1,j}$ is a variable or constant and $m_{1,1}\neq x$. Similarly, for each $i$, the fact that $\wec{v}_i\neq \wec{y}$ produces an identity \[ \Sigma\vdash F(m_{i,1},\ldots,m_{i,m})\approx x, \] where each $m_{i,j}$ is a variable or constant and $m_{i,i}\neq x$. Thus, it is a consequence of $\Sigma$ that the row identities of \[ F([m_{i,j}])\approx \left(\begin{matrix} x\\ \vdots\\ x \end{matrix} \right) \] hold. Since the diagonal elements of $[m_{i,j}]$ are not $x$, these identities make $F$ a pointed cube term for $\mathcal V$. [$(4)\Rightarrow(1)$] This is a tautology. \end{proof} One consequence of Lemma~\ref{nonrestrictive} is a procedure to decide if a strong Maltsev condition involving only basic identities implies the existence of a pointed cube term. \begin{cor}\label{decide_cor} A strong Maltsev condition defined by a set $\Sigma$ of basic identities entails the existence of a pointed $k$-cube term if and only if it is possible to prove from $\Sigma$ that some term of the form $F(x_1,\ldots,x_m)$ is a pointed $k$-cube term, where $m\geq 2$, $F$ is a function symbol occurring in $\Sigma$, and the variables $x_1,\ldots,x_m$ are distinct. \end{cor} \begin{proof} A strong Maltsev condition defined by a set $\Sigma$ of identities entails the existence of a pointed $k$-cube term if and only if the variety axiomatized by $\Sigma$ has a pointed $k$-cube term, so the corollary follows from Lemma~\ref{nonrestrictive} (1)$\Leftrightarrow$(3). \end{proof} That the property in the theorem statement can be decided follows from Theorem~\ref{kelly}. The next result is the main one of the subsection. \begin{thm}\label{nonrestrictive_thm} Let $\Sigma$ be a set of basic identities involving finitely many constant symbols. If $\Sigma$ does not entail the existence of a pointed cube term, then $\Sigma$ is nonrestrictive (and also nonrestrictive for finite algebras). \end{thm} \begin{proof} Recall that ``$\Sigma$ is nonrestrictive'' means that for every algebra $\m a$ there is a algebra $\m b$ realizing $\Sigma$ such that $d_{\m b}=d_{\m a}$, ``$\Sigma$ is restrictive'' means the opposite. Assume that $\Sigma$ fails to entail the existence of a pointed cube term. Choose $\m a$ arbitrarily and let $\m b = \m a_{\Sigma}$. $\m b$ realizes $\Sigma$ because $\m v$ is a reduct of $\m b$ and a model of $\Sigma$. We argue that $d_{\m b}=d_{\m a}$. Choose a generating set $G$ for $\m b^n$ such that $|G|=d_{\m b}(n)$. By Lemma~\ref{nonrestrictive} (1)$\Leftrightarrow$(3) we get that $G\cap A^n$ is a generating set for $\m a^n$, so $d_{\m a}(n)\leq |G\cap A^n|\leq d_{\m b}(n)$. Now choose a generating set $H$ for $\m a^n$ such that $|H|=d_{\m a}(n)$. Repeated use of Theorem~\ref{partial}~(1) shows that $H$ generates $\m a^n_{z_1,\ldots,z_p,0}$, hence also generates $\m a^n_{\Sigma}=\m b$. This shows that $d_{\m b}(n)\leq |H|=d_{\m a}(n)$. \end{proof} \begin{remark}\label{realize} In the third paragraph of this section we stated that the set $\Sigma$ from Example~\ref{example_constants} is nonrestrictive, yet restrictive for finite algebras when $\Sigma$ involves infinitely many constants. Here we explain why this remark is true, and also explain to what degree we may remove the assumption of finitely many constants in Theorem~\ref{nonrestrictive_thm}. Let $\Sigma$ be as in Example~\ref{example_constants} with $C$ an infinite set of constants. Let $\m a$ be any finite algebra. There is no finite $\m b$ that realizes $\Sigma$, hence none that realizes $\Sigma$ and satisfies $d_{\m b}=d_{\m a}$, since any nontrivial model of $\Sigma$ has cardinality at least $|C|$. On the other hand, $\Sigma$ does not entail the existence of a pointed cube term. Without attempting to give the full argument for this, we indicate only that if $\Sigma$ entailed the existence of a pointed cube term, then (i)~one would have the form $B_{c,d}(x_1,x_2)$, by Lemma~\ref{nonrestrictive}, and (ii)~it could not be a projection, so we would have to have $\Sigma\vdash B_{c,d}(e,x)\approx x$ and $\Sigma\vdash B_{c,d}(x,f)\approx x$ for some constants $e$ and $f$, and (iii)~$\{B_{c,d}(x,f)\}$ is a singleton class of the weak closure of $\Sigma$, hence we do not have $\Sigma\vdash B_{c,d}(x,f)\approx x$ after all. Finally, we sketch how to modify the proof of Theorem~\ref{nonrestrictive_thm} to eliminate the restriction to finitely many constants in the case where the algebras may be infinite. Recall that we started with an algebra $\m a$, enlarged it to $\m a_{z_1,\ldots,z_p,0}$ by iterating the one-point completion construction, and then merged it with the model $\m v$ of $\Sigma$ to create $\m a_{\Sigma}$, which realized $\Sigma$ and had the same growth rate as $\m a$. In this construction, we used the one-point completion construction $p$ times, where $p$ was the number of equivalence classes of constant symbols under $\Sigma$-provable equivalence. The only thing different here is that we may not have finitely many equivalence classes of constant symbols. However, we may well-order the equivalence classes of constants (say, by stipulating that $[c]<[d]$ if the least constant in class $[c]$ is smaller than the least constant in $[d]$ under the well-order from the proof of Kelly's Theorem). Now, rather than using the one-point completion construction $p$ times, we use the idea of the construction exactly once to adjoin a well-ordered set $\{0\}\cup Z$ to $\m a$ to create $\m a_{Z,0}$. Here the well-order is $0 < z_1 < z_2 < \cdots$, with $0$ the least element, and $\lb Z; <\rangle$ is a well-ordered set for which there is a bijection $\varphi\colon [C]\to Z$ from the set of equivalence classes of constants. The algebra has universe $A_{Z,0}$ equal to the disjoint union of $A$, $Z$ and $0$. If $F$ is a function symbol in the language of $\m a$, then it is defined on $A_{Z,0}$ by \[ F^{\m A_{Z,0}}(\wec{a}) = \begin{cases} F^{\m A}(\wec{a}) & \text{\rm if $\wec{a}\in A^n$};\\ \min\{\{a_1,\ldots,a_n\}\cap (\{0\}\cup Z)\} & \text{\rm else.} \end{cases} \] We also define binary operations corresponding to the operation $x\wedge y$ of the one-point completion, namely $x\wedge_z y$ for $z\in Z\cup \{0\}$. Here \[ x\wedge_z y = \begin{cases} x & \text{\rm if $x=y$};\\ z & \text{\rm if $x\neq y$ and $x, y\in A\cup [z)$};\\ \min\{\{x,y\}\cap (\{0\}\cup Z)\} & \text{\rm else.} \end{cases} \] Arguments similar to those in Theorem~\ref{partial} show that a generating set for $\m a^n$ also generates $\m a_{Z,0}^n$ and any generating set for $\m a_{Z,0}^n$ contains a generating set for $\m a^n$. We can merge $\m a_{Z,0}$ with a model $\m v$ of $\Sigma$ from Definition~\ref{V} to obtain a model $\m a_{\Sigma}$, as we did for the proof of Theorem~\ref{nonrestrictive_thm}. Using the same arguments as before, it can be shown that $\m a_{\Sigma}$ has the same growth rate as $\m a$ unless $\Sigma$ entails the existence of a pointed cube term. \end{remark} \subsection{Pointed cube terms enforce polynomially bounded growth}\label{processing} In the preceding subsection we proved that if $\Sigma$ is restrictive, then $\Sigma$ entails the existence of a pointed cube term. We now prove the converse by showing that if $\m a$ is an algebra with a pointed cube term and sufficiently many of the small powers of $\m a$ are finitely generated, then all finite powers of $\m a$ are finitely generated and $d_{\m a}(n)$ is bounded above by a polynomial. \begin{thm}\label{pointed_polynomial} Let $\m a$ be an algebra with an $m$-ary, $p$-pointed, $k$-cube term, with at least one constant symbol appearing in the cube identities (so $p\geq 1$). If $\m a^{p+k-1}$ is finitely generated, then all finite powers of $\m a$ are finitely generated and $d_{\m a}(n)$ is bounded above by a polynomial of degree at most $\log_w(m)$, where $w = 2k/(2k-1)$. \end{thm} The proof rests on the fact that a cube term, like \begin{equation}\label{R} {\sf P}\left(\begin{matrix} 1&x&2\\ x&2&3\\ \end{matrix} \right) \approx \left(\begin{matrix} x\\ x\\ \end{matrix} \right), \end{equation} may be used to ``factor'' a typical tuple $\wec{a}\in A^n$ into simpler tuples: \[ {\sf P}\left( \left[ \begin{matrix} {\mathbf 1}\\{\mathbf a}_2 \end{matrix} \right], \left[ \begin{matrix} {\mathbf a}_1\\{\mathbf 2} \end{matrix} \right], \left[ \begin{matrix} {\mathbf 2}\\{\mathbf 3} \end{matrix} \right] \right)= \left[ \begin{matrix} {\mathbf a}_1\\{\mathbf a}_2 \end{matrix} \right]. \] Here the $n$-tuple $\wec{a}$ has been split into two blocks of coordinates of roughly equal size, $\wec{a} = \left[ \begin{matrix} {\mathbf a}_1\\{\mathbf a}_2 \end{matrix} \right]$, then factored into $ \left[ \begin{matrix} {\mathbf 1}\\{\mathbf a}_2 \end{matrix}\right], \left[ \begin{matrix} {\mathbf a}_1\\{\mathbf 2} \end{matrix}\right], \left[ \begin{matrix} {\mathbf 2}\\{\mathbf 3} \end{matrix}\right], $ which are simpler than $\wec{a}$ in the sense that some of the coordinates have been replaced by constants. This factorization process can be iterated until the final factors have at most $k-1$ coordinate entries that have not been replaced by elements from the set of constants. The proof of the theorem develops such a factorization scheme under which there are only polynomially many different types of final factors, and the collection of all final factors of a given type lie in a subalgebra of $\m a^n$ isomorphic to $\m a^{j}$ for some $j\leq p+k-1$. The set consisting of the generators of all of these subalgebras is a polynomial-size generating set for $\m a^n$. \begin{proof} Suppose that the fact that ${\sf P}(x_1,\ldots,x_m)$ is a $p$-pointed $k$-cube term (with $p\geq 1$) is witnessed by identities ${\sf P}(M)\approx [x,\ldots,x]^{\mathsf T}$, where $M$ is a $k\times m$ matrix of variables and constant symbols, with at least one constant symbol, where each column of $M$ contains a symbol that is not $x$. Choose a constant symbol $c$ appearing in $M$, replace all instances of variables in $M$ that are not $x$ by $c$. This produces another matrix $R$ with no variables other than $x$ which also witnesses that ${\sf P}$ is a $p$-pointed $k$-cube term. The order of the $k$ rows identities, ${\sf P}(R)\approx [x,\ldots,x]^{\mathsf T}$, is fixed once and for all. We will refer to the function $\lambda\colon [m]\to [k]$ from the column indices to the row indices defined by the property that $\lambda(j)=i$ exactly when $i$ is the least index such that $R$ has a constant symbol in its $i,j$-th position. Such $\lambda$ exists because every column of $R$ contains at least one constant symbol. (For the cube term in the example immediately following the theorem statement $\lambda\colon [3]\to [2]$ is the function $\lambda(1)=\lambda(3) = 1, \lambda(2) = 2$.) The factoring, or ``processing'', of tuples in $A^n$ will make use of an $m$-ary tree which we refer to as the \emph{(processing) template}. We refer to nodes of the template by their \emph{addresses}, which are finite strings in the alphabet $[m] = \{1,\ldots,m\}$. The root node has empty address, and is denoted $\wec{n}_{\emptyset}$. If $\wec{n}_{\sigma}$ is the node at address $\sigma$, then its children are the nodes $\wec{n}_{\sigma 1},\ldots, \wec{n}_{\sigma m}$. Each node $\wec{n}$ of the template is labeled by a subset $\ell(\wec{n})\subseteq [n]$. (Recall that $n$ is the number appearing in the exponent of $A^n$.) To define the labeling function $\ell$ we first specify a fixed method for partitioning some subsets $U\subseteq [n]$. Given a subset $U=\{u_1,\ldots,u_r\}\subseteq [n]$, consider it to be a linearly ordered set $u_1 < \ldots < u_r$ under the order inherited from $[n]$. Define $\pi(U) = (U_1,\ldots,U_k)$ to be the ordered partition of $U$ into $k$ consecutive nonempty intervals that are as equal in size as possible. In more detail, let \[ \pi(U)=(U_1,\ldots,U_k)= (\{u_1,\ldots,u_{i_1}\},\{u_{i_1+1},\ldots,u_{i_2}\},\ldots, \{u_{i_{k-1}+1},\ldots,u_{i_k}=u_r\}), \] where \[ u_1<\cdots<u_{i_1}<u_{i_1+1}<\cdots<u_{i_2}<u_{i_{k-1}+1}<\cdots < u_{i_k}=u_r \] (i.e., the cells of the partition are consecutive nonempty intervals) and \[ |U_1|\geq \cdots \geq |U_k|\geq |U_1|-1 \] (i.e., the cells are as equal sized as possible). The $k$ appearing here as the number of cells of the partition is the same $k$ as the one in the assumption that ${\sf P}$ is a $k$-cube term. In order for $\pi(U)$ to be defined, it is necessary that $|U|\geq k$. As mentioned earlier, the label on node $\wec{n}_{\sigma}$ will be some subset $\ell(\wec{n}_{\sigma})\subseteq [n]$. Recursively define the labels as follows: \begin{enumerate} \item $\ell(\wec{n}_{\emptyset})=\emptyset$. \item If all nodes between $\wec{n}_{\sigma}$ and $\wec{n}_{\emptyset}$ are labeled, $V$ is the union of labels occurring between $\wec{n}_{\sigma}$ and the root $\wec{n}_{\emptyset}$, and $\pi([n]\setminus V) = (U_1,\ldots,U_k)$, then $\ell(\wec{n}_{\sigma i})= U_{\lambda(i)}$. \end{enumerate} In (2), if $[n]\setminus V$ has fewer than $k$ elements, then it is impossible to partition it into $k$ nonempty intervals, in which case there do not exist sufficiently many labels for potential children. In this case, we do not include any descendants of $\wec{n}_{\sigma}$ in the template. Let's illustrate our progress with the example started back at line (\ref{R}). The following picture depicts the processing template in the case $[n]=[5]=\{1,2,3,4,5\}$. (Recall that $\lambda(1)=\lambda(3) = 1, \lambda(2) = 2$.) \bigskip \begin{center} \begin{picture}(300,150) \setlength{\unitlength}{1mm} \put(51,46){\tiny{$\emptyset$}} \put(48,52){$\wec{n}_{\emptyset}$} \put(22,34){\tiny{$\{1, 2, 3\}$}} \put(18,37){$\wec{n}_{1}$} \put(45,37){$\wec{n}_{2}$} \put(51,34){\tiny{$\{4,5\}$}} \put(78.5,37){$\wec{n}_{3}$} \put(82,34){\tiny{$\{1, 2, 3\}$}} \put(-14,22){$\wec{n}_{11}$} \put(-12,17){\tiny{$\{4\}$}} \put(0,22){$\wec{n}_{12}$} \put(3,17){\tiny{$\{5\}$}} \put(14,22){$\wec{n}_{13}$} \put(18,17){\tiny{$\{4\}$}} \put(30,22){$\wec{n}_{21}$} \put(31,17){\tiny{$\{1, 2\}$}} \put(44,22){$\wec{n}_{22}$} \put(52,18){\tiny{$\{3\}$}} \put(65,22){$\wec{n}_{23}$} \put(61,17){\tiny{$\{1, 2\}$}} \put(81,22){$\wec{n}_{31}$} \put(78,17){\tiny{$\{4\}$}} \put(94,22){$\wec{n}_{32}$} \put(93,17){\tiny{$\{5\}$}} \put(109,22){$\wec{n}_{33}$} \put(108,17){\tiny{$\{4\}$}} \put(27,6){$\wec{n}_{221}$} \put(33,1){\tiny{$\{1\}$}} \put(41.5,6){$\wec{n}_{222}$} \put(48,1){\tiny{$\{2\}$}} \put(66,6){$\wec{n}_{223}$} \put(63,1){\tiny{$\{1\}$}} \put(50,50){\line(-2,-1){30}} \put(50,50){\line(0,-1){15}} \put(50,50){\line(2,-1){30}} \put(20,35){\line(-2,-1){30}} \put(20,35){\line(-1,-1){15}} \put(20,35){\line(0,-1){15}} \put(50,35){\line(-1,-1){15}} \put(50,35){\line(0,-3){15}} \put(50,35){\line(1,-1){15}} \put(80,35){\line(2,-1){30}} \put(80,35){\line(1,-1){15}} \put(80,35){\line(0,-1){15}} \put(50,20){\line(-1,-1){15}} \put(50,20){\line(0,-3){15}} \put(50,20){\line(1,-1){15}} \put(50,50){\circle*{1}} \put(20,35){\circle*{1}} \put(50,35){\circle*{1}} \put(80,35){\circle*{1}} \put(-10,20){\circle*{1}} \put(5,20){\circle*{1}} \put(20,20){\circle*{1}} \put(35,20){\circle*{1}} \put(50,20){\circle*{1}} \put(65,20){\circle*{1}} \put(80,20){\circle*{1}} \put(95,20){\circle*{1}} \put(110,20){\circle*{1}} \put(35,5){\circle*{1}} \put(50,5){\circle*{1}} \put(65,5){\circle*{1}} \end{picture} \end{center} Now we define precisely what is meant by processing. Let $P=\{c_1,\ldots,c_p\}$ be the constant symbols appearing in the cube identities for ${\sf P}$. A tuple $\wec{a}\in A^n$ is \emph{processed for node $\wec{n}_{\sigma}$} if there is a constant symbol $c\in P$ such that the $i$-th coordinate of $\wec{a}$ is $c^{\m a}$ for all $i\in \ell(\wec{n}_{\sigma})$. A tuple $\wec{a}$ is \emph{fully processed} if there is a path through the template from the root to a leaf such that $\wec{a}$ is processed for each node in the path. The processing template describes, in reverse order, a particular way to generate tuples in $\m a^n$. Given a tuple $\wec{a}\in A^n$, we assign it to the root $\wec{n}_{\emptyset}$ and denote it $\wec{a}_{\emptyset}$. This tuple $\wec{a} = \wec{a}_{\emptyset}$ is already processed for $\wec{n}_{\emptyset}$, since this is an empty requirement. Now, for each address $\sigma$ of a node in the template, we will construct $\wec{a}_{\sigma 1}, \ldots, \wec{a}_{\sigma m}$ from $\wec{a}_{\sigma}$ so that (i)~${\sf P}^{\m a}(\wec{a}_{\sigma 1}, \ldots, \wec{a}_{\sigma m}) = \wec{a}_{\sigma}$, and (ii)~each $\wec{a}_{\sigma i}$ is processed at all nodes between $\wec{n}_{\sigma i}$ and $\wec{n}_{\emptyset}$. Assign $\wec{a}_{\sigma i}$ to $\wec{n}_{\sigma i}$. The original tuple $\wec{a}$ can be generated via ${\sf P}^{\m a}$ by the fully processed tuples derived from $\wec{a}$ in this way. The following claim is the heart of this argument. \begin{clm} Suppose that $\wec{n}_{\sigma}$ is an internal node of the processing template. Given an arbitrary tuple $\wec{a}\in A^n$, there exist tuples $\wec{b}_{1}, \ldots, \wec{b}_{m}$ such that \begin{enumerate} \item ${\sf P}^{\m a}(\wec{b}_{1}, \ldots, \wec{b}_{m}) = \wec{a}$. \item $\wec{b}_{i}$ is processed for node $\wec{n}_{\sigma i}$ for $i=1,\ldots,m$. \item If $\wec{n}$ is a node between $\wec{n}_{\sigma}$ and $\wec{n}_{\emptyset}$, and $\wec{a}$ is processed for $\wec{n}$, then each $\wec{b}_{i}$ is also processed for $\wec{n}$ for $i=1,\ldots,m$. \end{enumerate} \end{clm} \cproof Let $V$ be the union of labels on nodes between $\wec{n}_{\sigma}$ and $\wec{n}_{\emptyset}$. If $\pi([n]\setminus V) = (U_1,\ldots,U_k)$, then $\{V,U_1,\ldots,U_k\}$ is a partition of $[n]$ (with $V$ possibly empty). For simplicity of expression, reorder coordinates so that $\wec{a}$ and $\wec{b}_i$ can be written $[\wec{a}_V, \wec{a}_{U_1}, \ldots, \wec{a}_{U_k}]^{\mathsf T}$ and $[\wec{b}_{i,V}, \wec{b}_{i,U_1}, \ldots, \wec{b}_{i,U_k}]^{\mathsf T}$, with coordinates from $V$ or $U_j$ grouped together. Given $\wec{a}$, we need to solve for $\wec{b}_{i,V}$ and $\wec{b}_{i,U_j}$ in \begin{equation}\label{ddddddd} \small{ {\sf P}^{\m a}(\wec{b}_1,\ldots,\wec{b}_m)= {\sf P}^{\m a}\left( \left[ \begin{array}{c} \wec{b}_{1,V}\\ \wec{b}_{1,U_1}\\ \vdots\\ \wec{b}_{1,U_k}\\ \end{array} \right], \ldots, \left[ \begin{array}{c} \wec{b}_{m,V}\\ \wec{b}_{m,U_1}\\ \vdots\\ \wec{b}_{m,U_k}\\ \end{array} \right] \right) = \left[ \begin{array}{c} \wec{a}_{V}\\ \wec{a}_{U_1}\\ \vdots\\ \wec{a}_{U_k}\\ \end{array} \right] =\wec{a} } \end{equation} in order to satisfy Item (1) of the claim. We shall do so using the first cube identity in the $V$-coordinates and the $U_1$-coordinates, and the $i$-th cube identity in the $U_i$-coordinates. Whether $W=V$ or $W=U_i$, to solve ${\sf P}^{\m a}(\wec{b}_{1,W},\ldots,\wec{b}_{m,W})=\wec{a}_{W}$ for the $\wec{b}_{i,W}$'s using a particular cube identity, take $\wec{b}_{i,W}=\wec{a}_W$ if there is an $x$ in the $i$-th place of $F$ in the cube identity, and take $\wec{b}_{i,W}=[c^{\m a},\ldots,c^{\m a}]^{\mathsf T}$ if there is a $c$ in the $i$-th place of the cube identity. It is not hard to see that this works, and so (1) holds. The label on node $\wec{n}_{\sigma i}$ is $U_{\lambda(i)}$. The element $\lambda(i)\in [k]$ is the number of the first cube identity that has some constant symbol $c\in P$ in the $i$-th place of $F$. Hence $\wec{b}_{i,U_{\lambda(i)}} = [c^{\m a},\ldots,c^{\m a}]^{\mathsf T}$. Thus $\wec{b}_i$ is processed for node $\wec{n}_{\sigma i}$, establishing (2). If, in the first cube identity, there is an $x$ in the $i$-th place of $F$, then $\wec{b}_{i,V} = \wec{a}_V$. If there is a constant symbol $c\in P$ in the $i$-th place of $F$, then $\wec{b}_{i,V} = [c^{\m a},\ldots,c^{\m a}]^{\mathsf T}$. In the latter case, $\wec{b}$ is processed at all coordinates in $V$, hence at all nodes between $\wec{n}_{\sigma}$ and $\wec{n}_{\emptyset}$. In the former case, $\wec{b}_i$ is processed at any node between $\wec{n}_{\sigma}$ and $\wec{n}_{\emptyset}$ where $\wec{a}$ is processed, since $\wec{b}_{i,V} = \wec{a}_V$. In either case, (3) holds. The claim is proved. \hfill\rule{1.3mm}{3mm} \medskip The claim shows that we can attach any tuple $\wec{a}\in A^n$ to the root node and then process it down the tree using the cube identities until we have attached to the leaves the fully processed tuples associated to $\wec{a}$. Here we indicate the processing of a tuple $\wec{a} \in A^5$ using the example template given earlier. \bigskip \bigskip \begin{center} \begin{picture}(300,180) \setlength{\unitlength}{1.1mm} \put(48,51){\tiny{$\emptyset$}} \put(49,55){${\tiny{\wec{a} = \left[\begin{matrix} a_1\\a_2\\a_3\\a_4\\a_5\end{matrix}\right]}}$} \put(22,34){\tiny{$\{1, 2, 3\}$}} \put(14,43){${\tiny{\left[\begin{matrix} 1\\1\\1\\a_4\\a_5\end{matrix}\right]}}$} \put(42,40){${\tiny{\left[\begin{matrix} a_1\\a_2\\a_3\\2\\2\end{matrix}\right]}}$} \put(51,34){\tiny{$\{4,5\}$}} \put(79,43){${\tiny{\left[\begin{matrix} 2\\2\\2\\3\\3\end{matrix}\right]}}$} \put(82,34){\tiny{$\{1, 2, 3\}$}} \put(-14,10){${\tiny{\left[\begin{matrix} 1\\1\\1\\1\\a_5\end{matrix}\right]}}$} \put(-12,22){\tiny{$\{4\}$}} \put(1,10){${\tiny{\left[\begin{matrix} 1\\1\\1\\a_4\\2\end{matrix}\right]}}$} \put(2,22){\tiny{$\{5\}$}} \put(16,10){${\tiny{\left[\begin{matrix} 2\\2\\2\\2\\3\end{matrix}\right]}}$} \put(16,22){\tiny{$\{4\}$}} \put(27,18){${\tiny{\left[\begin{matrix} 1\\1\\a_3\\1\\1\end{matrix}\right]}}$} \put(33.5,17){\tiny{$\{1, 2\}$}} \put(42,18){${\tiny{\left[\begin{matrix} a_1\\a_2\\2\\2\\2\end{matrix}\right]}}$} \put(52,18){\tiny{$\{3\}$}} \put(65,18){${\tiny{\left[\begin{matrix} 2\\2\\3\\2\\2\end{matrix}\right]}}$} \put(60,17){\tiny{$\{1, 2\}$}} \put(77,10){${\tiny{\left[\begin{matrix} 1\\1\\1\\1\\3\end{matrix}\right]}}$} \put(80,22){\tiny{$\{4\}$}} \put(91,10){${\tiny{\left[\begin{matrix} 2\\2\\2\\3\\2\end{matrix}\right]}}$} \put(93,22){\tiny{$\{5\}$}} \put(106,10){${\tiny{\left[\begin{matrix} 2\\2\\2\\2\\3\end{matrix}\right]}}$} \put(108,22){\tiny{$\{4\}$}} \put(30,-5){${\tiny{\left[\begin{matrix} 1\\a_2\\1\\1\\1\end{matrix}\right]}}$} \put(32,7){\tiny{$\{1\}$}} \put(46,-5){${\tiny{\left[\begin{matrix} a_1\\2\\2\\2\\2\end{matrix}\right]}}$} \put(45,7){\tiny{$\{2\}$}} \put(62,-5){${\tiny{\left[\begin{matrix} 2\\3\\2\\2\\2\end{matrix}\right]}}$} \put(65,7){\tiny{$\{1\}$}} \put(50,50){\line(-2,-1){30}} \put(50,50){\line(0,-1){15}} \put(50,50){\line(2,-1){30}} \put(20,35){\line(-2,-1){30}} \put(20,35){\line(-1,-1){15}} \put(20,35){\line(0,-1){15}} \put(50,35){\line(-1,-1){15}} \put(50,35){\line(0,-3){15}} \put(50,35){\line(1,-1){15}} \put(80,35){\line(2,-1){30}} \put(80,35){\line(1,-1){15}} \put(80,35){\line(0,-1){15}} \put(50,20){\line(-1,-1){15}} \put(50,20){\line(0,-3){15}} \put(50,20){\line(1,-1){15}} \put(50,50){\circle*{1}} \put(20,35){\circle*{1}} \put(50,35){\circle*{1}} \put(80,35){\circle*{1}} \put(-10,20){\circle*{1}} \put(5,20){\circle*{1}} \put(20,20){\circle*{1}} \put(35,20){\circle*{1}} \put(50,20){\circle*{1}} \put(65,20){\circle*{1}} \put(80,20){\circle*{1}} \put(95,20){\circle*{1}} \put(110,20){\circle*{1}} \put(35,5){\circle*{1}} \put(50,5){\circle*{1}} \put(65,5){\circle*{1}} \end{picture} \end{center} \bigskip \bigskip \bigskip \bigskip \bigskip Each leaf of the template determines a \emph{type} of fully processed tuples. Two fully processed tuples $\wec{u}$ and $\wec{v}$ of the same type have the same processed coordinates, and the same constant entries in the processed coordinates. They differ only in the unprocessed coordinates. For any given type there is a partition of the $n$ coordinates into at most $p+k-1$ cells where each unprocessed coordinate is a singleton cell (there are at most $k-1$ of these cells) and all processed coordinates with a given constant entry form a cell (there are at most $p$ of these cells). The collection of all tuples of this type lie in the subalgebra of all tuples constant on these cells, and this subalgebra is isomorphic to $\m a^{j}$ for some $j\leq p+k-1$. The assumption of the theorem is that $\m a^{p+k-1}$ is finitely generated, say by $g$ elements. This paragraph explains why $\m a^n$ has a subalgebra generated by $\leq g$ elements (and isomorphic to $\m a^{j}$ for some $j\leq p+k-1$) which contains all fully processed tuples of a given type. For example, the fully processed tuple ${\tiny{\left[\begin{matrix} 1\\1\\1\\a_4\\2\end{matrix}\right]}}$ from the preceding figure lies in the subalgebra of all tuples of the form ${\tiny{\left[\begin{matrix} x\\x\\x\\y\\z\end{matrix}\right]}}$, which is isomorphic to $\m a^3$. Here $3\leq p+k-1=3+2-1=4$. Now let's count the number of types. Since the template is an $m$-ary tree, and the types are determined by the leaves, the number of types is at most $m^r$ where $r$ is an upper bound on the length of the longest branch in the processing template. We must estimate $r$. Let $V_0 = \ell(\wec{n}_{\emptyset}) = \emptyset$. This represents the set of coordinate positions that have been processed before the processing begins, i.e., no coordinate positions. As we progress down a branch in the template, $\wec{n}_{\emptyset}, \wec{n}_i, \wec{n}_{ij},\ldots, \wec{n}_{\sigma}$, we may construct sets $V_{\sigma i} = V_{\sigma}\cup \ell(\wec{n}_{\sigma i})$, where $V_{\sigma}$ represents the set of coordinate positions that have been processed along this branch from $\wec{n}_{\emptyset}$ to $\wec{n}_{\sigma}$. The unprocessed coordinate positions, $[n]\setminus V_{\sigma}$ are then divided evenly, $\pi([n]\setminus V_{\sigma}) = (U_1,\ldots,U_k)$, to appear as labels of the children of $\wec{n}_{\sigma}$. Thus, $|V_{\emptyset}|=0$ and \begin{equation}\label{WWW} |V_{\sigma i}| = |V_{\sigma}\cup \ell(\wec{n}_{\sigma i})|= |V_{\sigma}|+|\ell(\wec{n}_{\sigma i})|. \end{equation} The useful parameter is the number $u_{\sigma}:= |[n]\setminus V_{\sigma}|=n- |V_{\sigma}|$ of nodes that remain unprocessed after reaching $\wec{n}_{\sigma}$. This parameter satisfies $u_{\emptyset} = |[n]\setminus V_{\emptyset}|=n$ and, from (\ref{WWW}), \begin{equation}\label{XXX} u_{\sigma i} = (n - |V_{\sigma i}|) = (n-|V_{\sigma}|)-|\ell(\wec{n}_{\sigma i})| = u_{\sigma}-|\ell(\wec{n}_{\sigma i})|. \end{equation} Since $\pi([n]\setminus V_{\sigma}) = (U_1,\ldots,U_k)$ is an even division of $[n]\setminus V_{\sigma}$ into $k$ sets, and $\ell(\wec{n}_{\sigma i})=U_{\lambda(i)}$, we get \begin{equation}\label{YYY} |\ell(\wec{n}_{\sigma i})|=|U_{\lambda(i)}|\geq \lfloor(n-|V_{\sigma}|)/k\rfloor = \lfloor u_{\sigma}/k\rfloor. \end{equation} Combining (\ref{XXX}) and (\ref{YYY}) we have \[ u_{\sigma i} \leq u_{\sigma} - \lfloor u_{\sigma}/k\rfloor = \left\lceil\left(\frac{k-1}{k}\right)u_{\sigma}\right\rceil. \] In order to avoid considering truncation error, we use the following fact, whose proof we leave to the reader. \begin{clm} If $u\geq k\geq 1$, then $\left\lceil \left(\frac{k-1}{k}\right)u\right\rceil\leq \left(\frac{2k-1}{2k}\right)u $.\hfill\rule{1.3mm}{3mm} \end{clm} Hence \[ u_{\sigma i} \leq \left(\frac{2k-1}{2k}\right)u_{\sigma} \] for each $\sigma$, and therefore \[ u_{\sigma} \leq \left(\frac{2k-1}{2k}\right)^{|\sigma|}u_{\emptyset} = \left(\frac{2k-1}{2k}\right)^{|\sigma|}n \] for each $\sigma$. If, for some $r$, it happens that $\left(\frac{2k-1}{2k}\right)^{r}n < k$, then there are fewer than $k$ unprocessed nodes at address $\sigma$ for any $\sigma$ satisfying $|\sigma|\geq r$. Such an $r$ is an upper bound on the length of paths through the template. Solving $\left(\frac{2k-1}{2k}\right)^{r}n < k$ for $r$ we obtain that any $r > \log_w(n/k)$, $w = \frac{2k}{2k-1}$, is an upper bound on the length of paths in the template; hence $r=\log_w(n/k)+1$ is such a bound. Hence the number of types of fully processed tuples is no more than \[ m^r= m^{\log_w(n/k)+1} = m^{\log_w(n/k)}m = (n/k)^{\log_w(m)}m \in O(n^{\log_w(m)}). \] Recall that for each type, the set of fully processed tuples lies in a $g$-generated subalgebra of $\m a^n$. Collecting these generators yields a set of size $O(n^{\log_w(m)})$ which generates all fully processed tuples, hence generates $\m a^n$. \end{proof} This theorem deals only with the case $p\geq 1$. We describe next how to refine the estimate in the case $p=1$ and how to derive the result for $p=0$ from the $p=1$ case. \begin{cor}\label{pointed_polynomial_cor} If $\m a^k$ is a finitely generated algebra with a $0$-pointed or $1$-pointed $k$-cube term, then $d_{\m a}(n)\in O(n^{k-1})$. \end{cor} \begin{proof} Suppose that $\m a$ has a $1$-pointed $k$-cube term, and that $c$ is the one constant that appears among the cube identities. Then a fully processed tuple $\wec{a}$ has $c$ in every processed coordinate position, and has at most $k-1$ unprocessed coordinate positions. Hence the set of tuples with a $c$ in all but at most $k-1$ positions contains all the fully processed tuples, and therefore is a generating set for $\m a^n$. Suppose that $\m a^k$ is $g$-generated. If $U\subseteq [n]$ has size $k-1$, then the subalgebra $\m a[U]$ of tuples in $\m a^n$ that are constant off of $U$ is isomorphic to $\m a^k$, and so is also $g$-generated. This subalgebra contains all tuples that have entry $c$ off of $U$. If we collect the $g$ generators for $\m a[U]$ for each $k-1$ element subset $U\subseteq [n]$ we obtain a set of size $ \binom{n}{k-1}g $ which generates $\m a^n$. Therefore $d_{\m a}(n)\leq \binom{n}{k-1}g\in O(n)$. Now suppose that $F(x_1,\ldots,x_m)$ is a $0$-pointed $k$-cube term of $\m a$ and that the cube identities are \begin{equation}\label{F} {\sf P}(M)\approx \left(\begin{array}{c} x\\ \vdots\\x\end{array}\right). \end{equation} Expand $\m a$ to an algebra $\m b$ by adjoining a single constant, say $c$. Replace all variables other than $x$ in (\ref{F}) with $c$ to obtain identities witnessing that $F(x_1,\ldots,x_m)$ is a $1$-pointed $k$-cube term for $\m b$. Hence $d_{\m b}(n)\in O(n^{k-1})$ by the earlier part of the argument. Now $d_{\m a}(n)\in O(n^{k-1})$ by Theorem~\ref{basic_estimates}~(4). \end{proof} In \cite{paper2} we improve this result by showing that finite algebras with a $0$-pointed $k$-cube term have logarithmic or linear growth. Let's combine the results of this subsection with the results of the previous subsection. \begin{thm}\label{summary} The following are equivalent for a set $\Sigma$ of basic identities in which only finitely many constant symbols occur. \begin{enumerate} \item[(1)] $\Sigma$ is restrictive. [That is, the class of $d$-functions of algebras is not equal to the class of $d$-functions of algebras that realize $\Sigma$.] \item[(2)] $\Sigma$ is restrictive for finite algebras. [The class of $d$-functions of finite algebras is not equal to the class of $d$-functions of finite algebras that realize $\Sigma$.] \item[(3)] The variety axiomatized by $\Sigma$ has a pointed cube term. \item[(4)] The variety axiomatized by $\Sigma$ has a pointed cube term of the form $F(x_1,\ldots,x_m)$, where $m\geq 2$, $F$ is a function symbol occurring in $\Sigma$, and the variables $x_1,\ldots,x_m$ are distinct. \item[(5)] If $\m a$ is an algebra realizing $\Sigma$ and $d_{\m a}(n)$ is finite for all $n$, then $d_{\m a}(n)$ is bounded above by a polynomial. \item[(6)] There is no (finite) algebra $\m a$ realizing $\Sigma$ such that $d_{\m a}(n) = 2^n$ for all $n$. \end{enumerate} \end{thm} \begin{proof} [$(1)\Rightarrow(3)$ and $(2)\Rightarrow(3)$] Theorem~\ref{nonrestrictive_thm}. [$(3)\Leftrightarrow(4)$] Lemma~\ref{nonrestrictive}. [$(3)\Rightarrow(5)$] Theorem~\ref{pointed_polynomial} and Corollary~\ref{pointed_polynomial_cor}. [$(5)\Rightarrow(6)$] $d_{\m a}(n) = 2^n$ is not bounded above by a polynomial. [$(6)\Rightarrow(1)$ and $(6)\Rightarrow(2)$] If (1) or (2) fails then $\Sigma$ is nonrestrictive (for finite algebras). Thus, there exists a (finite) algebra $\m A$ realizing $\Sigma$ with the same growth rate $d_{\m A}(n)=2^n$ as the $2$-element set equipped with no operations. Hence (6) fails. \end{proof} \subsection{Finite algebras with polynomial growth}\label{poly_growth} In this subsection we prove that the bound on growth rates for finite algebras with $1$-pointed $k$-cube terms, established in Corollary~\ref{pointed_polynomial_cor}, is sharp. \begin{thm}\label{example} For each $k\geq 2$ there is a finite algebra with a $1$-pointed $k$-cube term whose growth rate satisfies $d_{\m a}(n)\in \Theta(n^{k-1})$. \end{thm} \begin{proof} We shall first construct a partial algebra with the desired growth rate, then modify it slightly to obtain a total algebra satisfying the hypotheses of the theorem. The universe of the partial algebra will be $A = \{a_1,\ldots,a_q,1\}$. We equip this set with a partial $k$-ary operation $F$ which satisfies \[ F^{\m a}(1,x,\ldots,x,x) = F^{\m a}(x,1,\ldots,x,x) = \cdots = F^{\m a}(x,x,\ldots,x,1) = x \] for each $x\in A$, and which is undefined otherwise. Thus, $F^{\m a}$ is a partial near unanimity operation that is defined only on the nearly unanimous tuples where the lone dissenter is $1$ and on the tuple whose entries are unanimously $1$. Set $\m a = \langle A; F\rangle$. We shall prove the exact formula \begin{equation}\label{d_formula} d_{\m a}(n)= \binom{n}{0} + q\binom{n}{1} + q^2\binom{n}{2} + \cdots + q^{k-1}\binom{n}{k-1} \end{equation} for this partial algebra, which is a polynomial in $n$ of degree $k-1$, since $k = {\rm arity}(F)$ and $q=|A|-1$ are fixed. This will show that $\m a$ is a $(q+1)$-element partial algebra with $d_{\m a}(n)\in\Theta(n^{k-1})$. Choose and fix $n$. Define the \emph{support} of a tuple $\wec{a}\in A^n$ to be the subset ${\rm supp}(\wec{a})\subseteq [n]$ consisting of indices $s$ where $a_s\neq 1$. The proof involves showing that the set of all tuples whose support has size at most $k-1$ is the unique minimal generating set for $\m a^n$. To set up language for the argument, call a tuple $\wec{b}\in A^n$ an \emph{essential generator} if it is contained in any generating set for $\m a^n$. \begin{clm}\label{support} If $S\subseteq [n]$ and $G\subseteq A^n$, then let $G_S$ denote the set of tuples in $G$ that have support contained in $S$. If $\wec{a}\in \lb G\rangle$ has support in $S$, then $\wec{a}\in \lb G_S\rangle$. \end{clm} \cproof In $\m a$, we have \[ F^{\m a}(x_1,\ldots,x_k) = 1\Longleftrightarrow x_1 = x_2 = \cdots = x_k = 1. \] Hence, in $\m a^n$, if $F^{\m a^n}(\wec{g}_1,\ldots,\wec{g}_k)$ is defined and equal to $\wec{b}$, then $i\notin {\rm supp}(\wec{b})$ if and only if $i\notin {\rm supp}(\wec{g}_i)$ for any $\wec{g}_i$. Equivalently, \begin{equation}\label{generate} {\rm supp}(F^{\m a^n}(\wec{g}_1,\ldots,\wec{g}_k)) = \bigcup_{i=1}^k {\rm supp}(\wec{g}_i) \end{equation} whenever $F^{\m a^n}(\wec{g}_1,\ldots,\wec{g}_k)$ is defined. Now let $G(0) = G$, $G_S(0)=G_S$, $G(j+1) = G(j)\cup F^{\m a^n}(G(j),\ldots,G(j))$, and $G_S(j+1) = G_S(j)\cup F^{\m a^n}(G_S(j),\ldots,G_S(j))$. By induction on $j$, using (\ref{generate}), it can be shown that any tuple in $G(j)$ that has support in $S$ lies in $G_S(j)$. Since $\lb G\rangle = \bigcup_j G(j)$ and $\lb G_S\rangle = \bigcup_j G_S(j)$, any tuple in $\lb G\rangle$ with support in $S$ lies in $\lb G_S\rangle$. \hfill\rule{1.3mm}{3mm} \medskip \begin{clm}\label{empty} The tuple $\hat{1}=[1,1,\ldots,1]^{\mathsf T}$ of empty support is an essential generator. \end{clm} \cproof This follows immediately from Claim~\ref{support}. \hfill\rule{1.3mm}{3mm} \medskip \begin{clm}\label{supp} Any tuple whose support has size at most $k-1$ is an essential generator of $\m a^n$. \end{clm} \cproof Let $\wec{b}\in A^n$ be a tuple of support $S$ where $1\leq |S|\leq k-1$. Without loss of generality, $S = [\ell] =\{1,\ldots,\ell\}$ for some $1\leq \ell\leq k-1$. In order to obtain a contradiction to the claim, assume that $\wec{b}$ is not an essential generator. Then $\wec{b}$ can be generated by elements different from $\wec{b}$, so the equation $F^{\m a^n}(\wec{x}_1,\ldots,\wec{x}_k) = \wec{b}$ can be solved for the $\wec{x}_i$ in such a way that $\wec{b}\notin \{\wec{x}_1,\ldots,\wec{x}_k\}$. Moreover, by (\ref{generate}), the $\wec{x}_i$'s must be taken from the tuples whose support is contained in $S$. The equation to be solved is therefore: \begin{equation}\label{dividers} \small{ F^{\m a^n}(\wec{x}_1,\ldots,\wec{x}_k)= F^{\m a^n}\left( \left[ \begin{array}{c} x_{1,1}\\ \vdots\\ \underline{\;\;x_{\ell,1}\;\;}\\ 1\\ \vdots\\ 1\\ \end{array} \right], \ldots, \left[ \begin{array}{c} x_{1,k}\\ \vdots\\ \underline{\;\;x_{\ell,k}\;\;}\\ 1\\ \vdots\\ 1\\ \end{array} \right] \right) = \left[ \begin{array}{c} b_1\\ \vdots\\ \underline{\;\;\;b_{\ell}\;\;\;}\\ 1\\ \vdots\\ 1\\ \end{array} \right]=\wec{b}. } \end{equation} We have introduced horizontal segments as dividers separating the coordinates in $S = [\ell]$ from the remaining coordinates in order to make the argument clearer. Since $F^{\m a^n}(\wec{x}_1,\ldots,\wec{x}_k)$ is defined, every row above the dividers is a nearly unanimous row with exactly one $1$. Hence there are exactly $\ell$ $1$'s above the dividers. This means that there are at most $\ell$ columns which contain a $1$ above the dividers. Since there are $k$ such columns, and $k>\ell$, there is a column $\wec{x}_j$ that contains no $1$ above the dividers. Since the $i$-th row above the dividers is nearly unanimous with majority value $b_i$, the column $\wec{x}_j$ which contains no $1$'s above the dividers is exactly $\wec{b}$. This contradicts the assumption that $\wec{b}\notin \{\wec{x}_1,\ldots,\wec{x}_k\}$, showing that $\wec{b}$ is indeed an essential generator. \hfill\rule{1.3mm}{3mm} \medskip \begin{clm} $\m a^n$ is generated by the tuples whose support has size at most $k-1$. \end{clm} \cproof It is enough to show that if $\wec{b}$ has support $S$ of size $\ell \geq k$, then $\wec{b}$ can be generated from tuples whose support is properly contained in $S$. It is enough to prove this in the case where $S = [\ell]$. For this we must explain how to solve \begin{equation}\label{dividers2} \small{ F^{\m a^n}(\wec{x}_1,\ldots,\wec{x}_k)= F^{\m a^n}\left( \left[ \begin{array}{c} x_{1,1}\\ \vdots\\ \underline{\;\;x_{\ell,1}\;\;}\\ 1\\ \vdots\\ 1\\ \end{array} \right], \ldots, \left[ \begin{array}{c} x_{1,k}\\ \vdots\\ \underline{\;\;x_{\ell,k}\;\;}\\ 1\\ \vdots\\ 1\\ \end{array} \right] \right) = \left[ \begin{array}{c} b_1\\ \vdots\\ \underline{\;\;\;b_{\ell}\;\;\;}\\ 1\\ \vdots\\ 1\\ \end{array} \right]=\wec{b} } \end{equation} when $\ell \geq k$ in such a way that every column contains at least one $1$ above the dividers and the $i$-th row above the dividers is nearly unanimously equal to $b_i$. This is easy to do. Set $x_{1,1} = \cdots = x_{k,k} = 1$, then put exactly one $1$ arbitrarily in each of rows $k+1$ to $\ell$, then fill in the remaining entries above the dividers so that the $i$-th row above the dividers is nearly unanimously equal to $b_i$. \hfill\rule{1.3mm}{3mm} \medskip We have established up to this point that the set of tuples of support of size at most $k-1$ is the unique minimal generating set for $\m a^n$. To complete the proof that the partial algebra $\m a$ has the specified growth rate, observe that the number of tuples with support $S$ is $(|A|-1)^{|S|} = q^{|S|}$, so the number of tuples whose support has size $i$ is $q^i\binom{n}{i}$. This yields the formula $d_{\m a}(n)=\sum_{i=0}^{k-1} q^i\binom{n}{i}$. The one-point completion, $\m a_0$, is a total algebra with the same growth rate as $\m a$. Let $\m b$ be the expansion of $\m a_0$ by one constant symbol $1$ whose interpretation is $1^{\m b} = 1$. The operation $F^{\m b}$ still satisfies \[ F^{\m b}(1,x,\ldots,x,x) = F^{\m b}(x,1,\ldots,x,x) = \cdots = F^{\m b}(x,x,\ldots,x,1) = x \] for each $x\in A_0$, so it is a $1$-pointed $k$-cube term for $\m b$. By Theorem~\ref{partial} $\m a^n$ and $\m a_0^n$ have the same unique minimal generating set, $G$, which is the set of all tuples with support at most $k-1$; this set contains $\hat{1}$. The algebra $\m b$ must also have a unique minimal generating set, namely the set obtained from $G$ by deleting $\hat{1}=1^{\m b^n}$. Thus $d_{\m b}(n) = d_{\m a}(n)-1 = \sum_{i=1}^{k-1} q^i\binom{n}{i}\in O(n^{k-1})$. \end{proof} \subsection{Pointed cube polynomials can be avoided}\label{avoid} We have established that if $\m a$ is an algebra whose $d$-function assumes only finite values, and $\m a$ has a pointed cube term (or pointed cube polynomial for that matter), then $d_{\m a}(n)$ is bounded above by a polynomial function of $n$. The same growth rate can be obtained without a pointed cube term (or polynomial), as we show next. \begin{thm}\label{avoid_thm} Let $\m a$ be an algebra with $|A|>1$ whose $d$-function assumes only finite values. There is an algebra $\m b$ such that $d_{\m b}(n) = d_{\m a}(n)$ for all $n$, and \begin{enumerate} \item[(1)] the universe of $\m b$ is $B:=A\cup\{0,z\}$ where $0\neq z$ and $0, z\notin A$, \item[(2)] $\m b$ has a meet semilattice term operation, $\wedge$, with respect to which $\m b$ has height one and least element $0$, and \item[(3)] if $p(x,\wec{y})$ is an $m$-ary polynomial of $\m b$ in which $x$ actually appears and $p(z,\wec{b}) = z$ for some $\wec{b}\in B^{m-1}$, then either $p(x,\wec{y}) \approx x$ or else $p(x,\wec{y}) \approx x\wedge q(\wec{y})$ for some polynomial $q$ in which $x$ does not appear. \end{enumerate} In particular, $\m b$ does not have a pointed cube polynomial. \end{thm} \begin{proof} Let $G(n):=\{\wec{g}_{n,1},\ldots,\wec{g}_{n,d(n)}\}$ be a least size generating set for $\m a^n$. Let $\m a_z$ be the one-point completion of $\m a$ with the element $z\;(\notin A)$ taken to be the new point added. According to Theorem~\ref{partial}, the set $G(n)$ is also a least size generating set for $\m a_z$. Next, copying the idea of the construction in Theorem~\ref{big}, for each $\wec{a}\in (A_z)^n$ introduce a partial operation $F_{\wec{a}}(x_1,\ldots,x_{d(n)})$ on $A_z$ with the properties that (i)~the vector equation \begin{equation}\label{definition_Fb} F_{\wec{a}}(\wec{g}_{n,1},\ldots,\wec{g}_{n,d(n)}) = \wec{a} \end{equation} holds coordinatewise, and (ii)~$F_{\wec{a}}$ is defined only on those tuples required to make this equation hold. Let $\overline{\m A}_z$ be the set $A_z$ equipped with these partial operations. The partial algebra $\overline{\m A}_z$ is a reduct of $\m a_z$, so in passing from $\m a_z$ to $\overline{\m a}_z$ we may have lost but not gained some generating subsets of powers. On the other hand, our choice of the partial operations guarantees that $G(n)$ still generates the $n$-th power of $\overline{\m A}_z$. This implies that $G(n)$ is a least size generating set for $\overline{\m A}_z^n$ for each $n$, and hence that the $d$-functions of $\overline{\m A}_z$ and $\m a_z$ are the same. Finally, let $\m b=(\overline{\m a}_{z})_{0}$ be the one-point completion of $\overline{\m A}_z$ with the element $0\;(\notin A\cup\{z\})$ taken to be the new point added. With this choice the universe of $\m B$ is $B=A\cup\{0,z\}$. Again citing Theorem~\ref{partial}, we see that $G(n)$ is a least size generating set for $\m b^n$. At this point we have that $d_{\m b}(n) = d_{\m a}(n)$ for all $n$, and also, by construction, that Items (1) and (2) hold. (Here the meet operation referred to in Item (3) is the one introduced in the second one-point completion, the one used to construct $\m b$ from $\overline{\m a}_z$.) Let's prove that Item (3) holds. Our argument depends on a Key Fact: $z$ does not appear in any coordinate of any tuple in $G(n)$ for any $n$, hence $z$ does not appear in any tuple in the domain of any partial operation of the form $F_{\wec{a}}$. This implies that any basic operation of $\m b$ of the form $(F_{\wec{a}})_0$ (Definition~\ref{one-point}) assigns the value $0$ to any tuple containing a $z$ (or a $0$). We first prove that if $p(x,\wec{y})$ is an $m$-ary polynomial of $\m b$ in which $x$ appears and $\wec{b}\in B^{m-1}$, then $p(z,\wec{b})\in \{0,z\}$. Arguing by induction on the complexity of $p$, we need to consider the cases were $p$ is a constant, a variable, or of the form \begin{equation}\label{inductive_arg} p(x,\wec{y}) = F(p_1(x,\wec{y}) ,\ldots,p_{\ell}(x,\wec{y})) \end{equation} where $F = (F_{\wec{a}})_0$ or $F = \wedge$. The polynomial $p$ cannot be a constant, since $x$ appears in $p$. If $p$ is a variable, it must be $x$, since $x$ appears in $p$. In this case $p(z,\wec{b})=z\in \{0,z\}$, as claimed. If (\ref{inductive_arg}) holds in the case where $F = (F_{\wec{a}})_0$, then by induction we have $p_i(z,\wec{b})\in \{0,z\}$ for at least one $i$, hence by the Key Fact we obtain that \[ p(z,\wec{b}) = (F_{\wec{a}})_0(p_1(z,\wec{b}) ,\ldots,p_{\ell}(z,\wec{b})) = 0\in\{0,z\}, \] as claimed. If (\ref{inductive_arg}) holds in the case where $F = \wedge$, then by induction we have $p_i(z,\wec{b})\in \{0,z\}$ for at least one $i$, hence $p_i(z,\wec{b})\leq z$. It follows that $p(z,\wec{b}) = p_1(z,\wec{b})\wedge p_2(z,\wec{b})\leq z$, so, since $\langle B;\wedge\rangle$ has height one, we get that $p(z,\wec{b}) \in \{0,z\}$. Now we prove Item (3) by induction on the complexity of $p$. Under the assumptions of Item (3) the polynomial $p$ cannot be a constant, since $x$ appears in $p$. If $p$ is a variable, it must be $x$, since $x$ appears in $p$, in which case $p(x,\wec{y}) = x$ for all $x$ and $\wec{y}$, and Item (3) holds. Now assume that (\ref{inductive_arg}) holds in the case where $F = (F_{\wec{a}})_0$, and fix a tuple $\wec{b}\in B^{m-1}$ satisfying $p(z,\wec{b})=z$ (the existence of such a $\wec{b}$ is assumed in Item~(3)). Since $x$ appears in $p$, by the induction hypothesis we have $p_i(x,\wec{y}) = x$ or $x\wedge q_i(\wec{y})$ for some $i$ and some polynomial $q_i$. In either case, $p_i(z,\wec{b})\in \{0,z\}$ by the result of the preceding paragraph, and this gives us the right hand equality (the only nontrivial equality) in: \[ z = p(z,\wec{b}) = (F_{\wec{a}})_0(p_1(z,\wec{b}) ,\ldots,p_{\ell}(z,\wec{b})) = 0. \] This is a contradiction, which shows that this case cannot occur. Finally, if $ p(x,\wec{y}) = p_1(x,\wec{y}) \wedge p_2(x,\wec{y}) $ and $\wec{b}\in B^{m-1}$ is such that $p(z,\wec{b}) = z$, then $p_i(z,\wec{b})=z$ for $i=1,2$, since $z$ is meet irreducible in $\langle B; \wedge\rangle$. If $x$ appears in both $p_1(x,\wec{y})$ and $p_2(x,\wec{y})$, then by induction both have the form $x$ or $x\wedge q_i(\wec{y})$. Hence $p(x,\wec{y})$ has the form \[ x\wedge x,\quad x\wedge (x\wedge q_2(\wec{y})),\quad (x\wedge q_1(\wec{y}))\wedge x,\;\quad\textrm{or}\quad\; (x\wedge q_1(\wec{y}))\wedge (x\wedge q_2(\wec{y})), \] each of which has the form $x$ or $x\wedge q(\wec{y})$ for some polynomial $q$. A similar conclusion is reached if $x$ appears in one of the polynomials $p_i(x,\wec{y})$ but not the other. Hence Item (3) holds. To complete the proof of the theorem we argue that $\m b$ does not have a pointed cube polynomial. By way of contradiction, assume that $p(x_1,\ldots,x_m)$ is such a polynomial and that $M$ is a $k\times m$ matrix of variables and constants such that $p(M)\approx [x,\ldots,x]^{\mathsf T}$ and every column of $M$ contains at least one entry that is not $x$. In fact, as we have seen before, by substituting constants for the variables different from $x$ we may assume that the entries of $M$ are constants or $x$ and that each column contains at least one constant. We may also assume that $p$ depends on all of its variables, hence that each of $x_1,\ldots, x_m$ appears in $p$. Here are some elementary consequences of our assumptions. \begin{enumerate} \item[(a)] Each row of $M$ must contain at least one $x$, since otherwise we may derive from the associated cube identity that $x\approx y$ holds in $\m b$. By permuting columns of $M$ (hence reordering the variables of $p$), we assume that the first entry of the first row is $x$. \item[(b)] The first column of $M$ contains a constant, which cannot be in the first row. By permuting the later rows of $M$ (hence reordering the cube identities), we assume that the first entry of the second row of $M$ is a constant. There is an $x$ somewhere on the second row, by (a), and permuting the later columns we may assume that it is in the second position of the second row. \end{enumerate} These consequences mean that the first two cube identities look like $p(x,b_2,\wec{b})\approx x$ and $p(c_1,x,\wec{c})\approx x$ where all $b_i, c_j\in B\cup \{x\}$ and $c_1$ is constant. If we substitute $z$ for each $x$ in these equations we get $p(z,b_2',\wec{b}') = z$ and $p(c_1',z,\wec{c}') = z$, where the primes on elements and tuples indicate that the $x$'s in the string have been replaced by $z$'s and constants remain the same. Applying Item (3) of this theorem to these equalities we obtain that \[ p(x_1,x_2,\wec{y}) = x_1\wedge q_1(x_2,\wec{y}) = x_2\wedge q_2(x_1,\wec{y}), \] where $x_i$ does not appear in $q_i$. By meeting $p$ with itself we obtain that \[ p(x_1,x_2,\wec{y}) = (x_1\wedge q_1(x_2,\wec{y}))\wedge(x_2\wedge q_2(x_1,\wec{y})). \] Now the second cube identity may be written \[ x=p(c_1,x,\wec{c})= (c_1\wedge q_1(x,\wec{c}))\wedge(x\wedge q_2(c_1,\wec{c}))\leq c_1. \] This implies $x\leq c_1$ for all $x\in B$, and therefore that the element $c_1\in B$ is the largest element of $\langle B;\wedge\rangle$. But this semilattice has no largest element, since it has at least 4 elements and has height 1. This contradiction proves that $\m b$ has no pointed cube polynomial. \end{proof} \subsection{Exponential growth}\label{exp} If $\m a$ has exponential growth and $\m b$ has arbitrary growth, then $\m a\times \m b$ has exponential growth according to Theorem~\ref{basic_estimates}~(2). Hence it is probably unrealistic to expect any meaningful classification of algebras with exponential growth. This subsection will therefore be limited to identifying one property that forces exponential growth. We will use the property to show that the variety generated by the 2-element implication algebra, $\lb \{0,1\}; \to \rangle$, contains a chain of finite algebras $\m a_1\leq \m a_2\leq \cdots$, each one a subalgebra of the next, where $\m a_i$ has logarithmic growth when $i$ is odd and exponential growth when $i$ is even. We explore a simple idea: Suppose that $\m a$ is finite and $u$ and $v$ are distinct elements of $A$. If every element of $\{u, v\}^n$ is an essential generator of $\m a^n$ for each $n$, then the growth rate of $\m a$ must be at least $2^n$. A way to force some tuple $\wec{t}\in \{u, v\}^n$ to be an essential generator of $\m a^n$ is to arrange that $A^n\setminus\{\wec{t}\}$ is a subuniverse of $\m a^n$. This can be accomplished by imposing an irreducibility condition on each coordinate $t$ of $\wec{t}$, or equivalently by requiring that the complementary set $A\setminus\{t\}$ behaves like an ideal. For this to work it is enough that $A\setminus\{t\}$ behaves like a 1-sided semigroup-theoretic ideal, so we introduce a definition that captures this notion for an arbitrary algebraic signature. \begin{df} Let $\sigma=(F,\alpha)$ be an algebraic signature. I.e., let $F$ be a set (of operation symbols) and let $\alpha\colon F\to \omega$ be a function (assigning arity). Let $F_0\subseteq F$ be the set consisting of those $f\in F$ such that $\alpha(f)>0$. ($F_0$ is the set of nonnullary symbols.) A \emph{selector} for $\sigma$ is a function $\phi\colon F_0\to \omega$ such that $1\leq \phi(f)\leq \alpha(f)$ for each $f\in F_0$. ($\phi$ selects one of the places of the function symbol $f$.) If $\phi$ is a selector for $\sigma$ and $\m a$ is an algebra of signature $\sigma$, then a \emph{$\phi$-irreducible} subset of $\m a$ is a subset $U\subseteq A$ such that whenever $\alpha(f)=n$ and $\phi(f)=i$ one has \[ f^{\m a}(a_1,\ldots,a_n)\in U\Rightarrow a_i\in U. \] The complement of a $\phi$-irreducible subset is called a \emph{$\phi$-ideal}. Explicitly, $I\subseteq A$ is a $\phi$-ideal if whenever $\alpha(f)=n$, $\phi(f)=i$ and $a_i\in I$, then $f^{\m a}(a_1,\ldots,a_n)\in I$. \end{df} In this terminology, a left ideal of a semigroup with multiplication represented by the symbol $m$ would be a $\phi$-ideal for the function $\phi\colon \{m\}\to \{1,2\}\colon m\mapsto 2$, while a right ideal would be a $\phi$-ideal for the function $\phi\colon \{m\}\to \{1,2\}\colon m\mapsto 1$. \begin{thm}\label{ideal} Let $\m a$ be an algebra of signature $\sigma$ and let $\phi$ be a selector for $\sigma$. If $\m a$ is the union of finitely many proper $\phi$-ideals, then $d_{\m a}(n)\geq 2^n$. \end{thm} \begin{proof} The union of $\phi$-ideals is again a $\phi$-ideal, so if $\m a$ is the union of $k\geq 2$ proper $\phi$-ideals then it can be expressed as the union $I\cup J$ of 2 proper $\phi$-ideals. The complements $I':=A\setminus I$ and $J':=A\setminus J$ are disjoint $\phi$-irreducible sets. Any product $T:=X_1\times \cdots \times X_n$, with $X_i=I'$ or $J'$ for all $i$, is a $\phi$-irreducible subset of $A^n$. Each such set must contain at least one element of any generating set, since the $\phi$-irreducibility of $T$ implies that $A^n\setminus T$ is a subuniverse of $\m a^n$. Since there are $2^n$ products of the form $X_1\times \cdots \times X_n$ with $X_i=I'$ or $J'$, and they are pairwise disjoint, any generating set for $\m a^n$ must contain at least $2^n$ elements. \end{proof} \begin{exmp} In this example, $\m 2$ is the 2-element Boolean algebra and ${\m 2}^{\circ} = \lb \{0,1\}; \to\rangle$ is the reduct of $\m 2$ to the operation $x\to y = x'\vee y$. The variety $\vr V$ generated by $\m 2^{\circ}$ is called the variety of implication algebras. This variety is congruence distributive and has $\m 2^{\circ}$ as its unique subdirectly irreducible member. Each finite algebra in $\vr V$ may be viewed as an order filter in a finite Boolean algebra: if $\m a\in {\vr V}_{\textrm{fin}}$, then an irredundant subdirect representation $\m a\leq (\m 2^{\circ})^k$ may be viewed as a representation of $\m a$ as a subset of $\m 2^k$ closed under $\to$; such subsets of $\m 2^k$ are order filters. Considering an algebra $\m a\in {\vr V}_{\textrm{fin}}$ to be an order filter in $\m 2^k$, each order filter contained within $\m a$ is a left ideal in $\m a$ with respect to the operation $\to$. By Theorem~\ref{ideal}, if $\m a$ is the union of its proper order filters, its growth rate is exponential. This case must occur unless $\m a$ itself is a principal order filter in $\m 2^k$. Since we represented $\m a$ irredundantly, $\m a$ is a principal order filter in $\m 2^k$ only when it is the improper filter, i.e., $\m a = (2^{\circ})^k$. In this situation $\m a$ is polynomially equivalent to the Boolean algebra $\m 2^k$. It follows from Theorem~\ref{basic_estimates}~(1) and the fact that $\m 2$ is primal that $\m 2^k$ has logarithmic growth rate. In summary, a finite implication algebra has logarithmic growth rate if it has a least element and has exponential growth rate otherwise. Now, it is easy to produce a chain of implication algebras $\m a_1\leq \m a_2\leq \cdots$, each one a subalgebra of the next, where $\m a_i$ has logarithmic growth when $i$ is odd and exponential growth when $i$ is even. One simply chooses larger and larger Boolean order filters which are principal only when $i$ is odd. The following figure shows how the chain might begin. \bigskip \begin{center} \vbox{ \begin{picture}(300,90) \setlength{\unitlength}{1mm} \put(0,20){\circle*{1}} \put(0,30){\circle*{1}} \put(0,20){\line(0,1){10}} \put(20,20){\circle*{1}} \put(30,20){\circle*{1}} \put(30,30){\circle*{1}} \put(20,20){\line(1,1){10}} \put(30,20){\line(0,1){10}} \put(50,10){\circle*{1}} \put(50,20){\circle*{1}} \put(60,20){\circle*{1}} \put(60,30){\circle*{1}} \put(50,10){\line(1,1){10}} \put(50,10){\line(0,1){10}} \put(50,20){\line(1,1){10}} \put(60,20){\line(0,1){10}} \put(80,10){\circle*{1}} \put(80,20){\circle*{1}} \put(90,20){\circle*{1}} \put(90,30){\circle*{1}} \put(100,20){\circle*{1}} \put(80,10){\line(1,1){10}} \put(80,10){\line(0,1){10}} \put(80,20){\line(1,1){10}} \put(90,20){\line(0,1){10}} \put(100,20){\line(-1,1){10}} \put(-2,2){$\m a_1$} \put(23,2){$\m a_2$} \put(53,2){$\m a_3$} \put(88,2){$\m a_4$} \put(11,2){$\leq$} \put(38,2){$\leq$} \put(71,2){$\leq$} \put(103,2){$\cdots$} \end{picture} \medskip {{\small\sc Figure: A chain of implication algebras.}} } \end{center} \end{exmp} \section{Problems}\label{problems} In this paper, we have filled in one gap in knowledge about the spectrum of possible growth rates of finite algebras by producing examples with superlinear polynomial growth rates. There is an interesting gap in knowledge that remains between logarithmic and linear growth rates. \medskip \noindent {\bf Problem 6.1.} Is there a finite algebra $\m a$ where $d_{\m a}(n)\notin \Omega(n)$ and $d_{\m a}(n)\notin O(\log(n))$? \medskip A special case that might be tractable is the following. \medskip \noindent {\bf Problem 6.2.} Is there a 2-element partial algebra $\m a$ where $d_{\m a}(n)\notin \Omega(n)$ and $d_{\m a}(n)\notin O(\log(n))$? \medskip We know that no finite algebra with a $0$-pointed cube term can have growth rate between logarithmic and linear, but do not know the situation for pointed cube terms. The following seems to be the most interesting special case. \medskip \noindent {\bf Problem 6.3.} Is it true that a finite algebra with a $2$-sided unit for some binary term has logarithmic or linear growth? \medskip There is also a possible gap near the exponential end of the spectrum. \medskip \noindent {\bf Problem 6.4.} Is there a finite algebra $\m a$ where $d_{\m a}(n)\notin 2^{\Omega(n)}$ and $d_{\m a}(n)\notin O(n^k)$ for any $k$? \bibliographystyle{plain}
\subsection{Spectral Mirror Algorithm} We begin by presenting our algorithm for estimating the subspace span $U$. Our algorithm consists of three main steps. First, as pre-processing, we estimate the mean and covariance of the underlying features $X_i$. Second, using these estimates, we identify a vector $\hat{r}$ that concentrates near the convex cone spanned by the profiles $(u_\ell)_{\ell \in [k]}$. We use this vector to perform an operation we call \emph{mirroring}: we `flip' all labels lying in the negative halfspace determined by $\hat{r}$. Finally, we compute a weighted covariance matrix $\hat{Q}$ over all $X_i$, where each point's contribution is weighed by the mirrored labels: the eigenvectors of this matrix, appropriately transformed, yield the span $U$. These operations are summarized in Algorithm~\ref{algo:spectral_mirror}. We discuss each of the main steps in more detail below: \noindent\textbf{Pre-processing.} (Lines 1--2) We split the dataset into two halves. Using the first half (\emph{i.e.}, all $X_i$ with $1\leq i \leq \lfloor \frac{n}{2} \rfloor)$, we construct estimates $\hat{\mu}\in\ensuremath{\mathbb{R}}^d$ and $\hat{\Sigma}\in \ensuremath{\mathbb{R}}^{d\times d}$ of the feature mean and covariance, respectively. Standard Gaussian ({i.e.}, `whitened') versions of features $X_i$ can be constructed as $\hat{\Sigma}^{-1/2}(X_i\!-\!\hat{\mu})$. \noindent{\textbf{Mirroring.}} (Lines 3--4) We compute the vector: \begin{align}\hat{r} = \frac{1}{\lfloor n/2 \rfloor} \textstyle\sum_{i=1}^{\lfloor n/2\rfloor} Y_i \hat{\Sigma}^{-1}(X_i\!-\!\hat{\mu}) \in \ensuremath{\mathbb{R}}^d. \label{eq:mirror} \end{align} We refer to $\hat{r}$ as the \emph{mirroring direction}. In Section~\ref{sec:proof}, we show that $\hat{r}$ concentrates around its population ($n=\infty$) version $r\equiv \ensuremath{\mathbb{E}}[ Y\Sigma^{-1}(X-\mu)]$. Crucially, $r$ lies in the interior of the convex cone spanned by the parameter profiles, i.e., $r=\sum_{\ell=1}^k \alpha_\ell u_\ell$, for some positive $\alpha_\ell>0$, $\ell\in [k]$ (see Lemma~\ref{lemma:quad} and Fig.~\ref{figure:3dplots}(b)). Using this $\hat{r}$, we `mirror' the labels in the second part of the dataset: $$Z_i = Y_i \ensuremath{\mathop{\mathrm{sgn}}}\<\hat{r},X_i\>, \quad \text{for~} \lfloor n/2\rfloor< i \leq n. $$ In words, $Z_i$ equals $Y_i$ for all $i$ in the positive half-space defined by the mirroring direction; instead, all labels for points $i$ in the negative half-space are flipped ({i.e.}, $Z_i=-Y_i$). This is illustrated in Figure~\ref{figure:3dplots}(c). \noindent{\textbf{Spectral Decomposition.}} (Lines 5--8) The mirrored labels are used to compute a weighted covariance matrix over whitened features as follows: \begin{align*} \hat{Q} & =\frac{1}{\lceil \frac{n}{2}\rceil} \sum_{i=\lfloor n/2\rfloor+1}^n Z_i\hat{\Sigma}^{-1/2}(X_i-\hat{\mu})(X_i-\hat{\mu})^T \hat{\Sigma}^{-1/2} \end{align*} The spectrum of $\hat{Q}$ has a specific structure, that reveals the span $U$. In particular, as we will see in Section~\ref{sec:proof}, $\hat{Q}$ converges to a matrix $Q$ that contains an eigenvalue with multiplicity $n-k$; crucially, the eigenvectors corresponding to the remaining $k$ eigenvalues, subject to the linear transform $\hat{\Sigma}^{-1/2}$, span the subspace $U$. As such, the final steps of the algorithm amount to discovering the eigenvalues that `stand out' ({i.e.}, are different from the eigenvalue with multiplicity $n-k$), and rotating the corresponding eigenvectors to obtain $\hat{U}$. More specifically, let $(\lambda_\ell,w_\ell)_{\ell\in [d]}$ be the eigenvalues and eigenvectors of $\hat{Q}$. Recall that $k< n/2$. The algorithm computes the median of all eigenvalues, and identifies the $k$ eigenvalues furthest from this median; these are the `outliers'. The corresponding $k$ eigenvectors, multiplied by $\hat{\Sigma}^{-1/2}$, yield the subspace estimate $\hat{U}$. The algorithm \emph{does not require} knowledge of the classifier response function $f$. Also, while we assume knowledge of $k$, an eigenvalue/eigenvectors statistic (see, e.g., \citet{zelnik2004self}) can be used to estimate $k$, as the number of `outlier' eigenvalues. \subsection{Main Result} Our main result states that \textsc{SpectralMirror}\xspace is a consistent estimator of the subspace spanned by $(u_\ell)_{\ell\in [k]}$. This is true for `most' $\mu\in \ensuremath{\mathbb{R}}^d$. Formally, we say that an event occurs for \emph{generic} $\mu$ if adding an arbitrarily small random perturbation to $\mu$, the event occurs with probability $1$ w.r.t.~this perturbation. \begin{thm} \label{thm:main} Denote by $\hat{U}$ the output of \textsc{SpectralMirror}\xspace, and let $P^\bot_{r} \equiv I -r r^T/\|r\|^2$ be the projector orthogonal to $r$, given by \eqref{eq:RDefinition}. Then, for generic $\mu$, as well as for $\mu=0$, there exists $\epsilon_0>0$ such that, for all $\epsilon\in [0,\epsilon_0)$, \begin{align*} \ensuremath{\mathbf{Pr}}(d_P(P^\bot_{r} U,\hat{U}) > \epsilon) \le C_1\exp(-C_2 \frac{n\epsilon^2}{d}). \end{align*} Here $C_1$ is an absolute constant, and $C_2>0$ depends on $\mu$, $\Sigma$, $f$ and $(u_{\ell})_{\ell\in [k]}$. \end{thm} In other words, $\hat{U}$ provides an accurate estimate of $P^\bot_{r} U$ as soon as $n$ is significantly larger than $d$. This holds for generic $\mu$, but we also prove that it holds for the specific and important case where $\mu=0$; in fact, it also holds for all small-enough $\mu$. Note that this does not guarantee that $\hat{U}$ spans the direction $r\in U$; nevertheless, as shown below, the latter is accurately estimated by $\hat{r}$ (see Lemma~\ref{lemma:hatr_convergence}) and can be added to the span, if necessary. Moreover, our experiments suggest this is rarely the case in practice, as $\hat{U}$ indeed includes the direction $r$ (see Section~\ref{sec:experiments}). \section{A Large-Deviation Lemma} We first prove a Bernstein-type inequality for sub-Gaussian random vectors, that we shall use in our proofs: \begin{lemma} \label{lem:large-deviation-sub-Gaussian-vector} Let $X\in\ensuremath{\mathbb{R}}^d$ be a sub-Gaussian random vector, {i.e.}\ $\langle a,X\rangle$ is sub-Gaussian for any $a\in\ensuremath{\mathbb{R}}^d$. Then there exist universal constants $c_1,c_2$ such that \begin{align*} \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t)& \le\\ c_1\exp\bigg(-\min&\left\{\frac{c_2(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2},\frac{(t^2 - d\norm{\Sigma}_2)^2}{64c_2\norm{X}_{\psi_2}^4}\right\}\bigg). \end{align*} \end{lemma} \begin{proof} By the (exponential) Markov inequality, we have \begin{equation}\begin{split} \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) = \ensuremath{\mathbf{Pr}}(\exp(\lambda\norm{X}_2^2) \ge \exp(\lambda t^2))\\ \le \frac{\Expect[\exp(\lambda\norm{X}_2^2)]}{\exp(\lambda t^2)}.\end{split} \label{eq:large-deviation-sub-Gaussian-vector-1} \end{equation} Let $Z$ be uniformly distributed on the unit sphere $\mathbf{S}^{d-1}$. Then $\sqrt{d}Z$ is isotropic so $d\Expect[\langle Z,a\rangle^2] = \norm{a}_2^2$ for all $a$. We use this fact to bound the m.g.f. of $\norm{X}_2^2$: \begin{align*} \Expect[\exp(\lambda\norm{X}_2^2)] &= \Expect_X[\exp(\lambda d\Expect_{Z}[\langle Z,X\rangle^2])] \\ &\le \Expect_X[\Expect_{Z}[\exp(\lambda d\langle Z,X\rangle^2)]]. \end{align*} We interchange the order of expectation to obtain \begin{align*} \Expect[\exp(\lambda\norm{X}_2^2)] &\le \Expect_Z[\Expect_{X}[\exp(\lambda d\langle X,Z\rangle^2)]] \\ &\le \sup\,\{\Expect_{X}[\exp(\lambda d\langle X,Z\rangle^2)]\mid z\in\mathbf{S}^{d-1}\} . \end{align*} $\langle X, z\rangle$ is sub-Gaussian (for fixed $z$) so $\langle X, z\rangle^2$ is (noncentered) sub-exponential. If $\lambda d < c/\norm{\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]}_{\psi_1}$, then \begin{align} &\Expect[\exp(\lambda d\langle X, z\rangle^2)]\nonumber \\ &\le \exp(\lambda d\Expect[\langle X, z\rangle^2])\Expect[\exp(\lambda d(\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]))] \nonumber\\ &\le \exp(\lambda d\Expect[\langle X, z\rangle^2] + c d^2\lambda^2\norm{\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]}_{\psi_1}^2)\nonumber \\ &\le \exp(\lambda d\Expect[\langle X, z\rangle^2] + 4c d^2\lambda^2\norm{\langle X, z\rangle^2}_{\psi_1}^2) \nonumber \\ &\le \exp(\lambda d\Expect[\langle X, z\rangle^2] + 16c d^2\lambda^2\norm{\langle X, z\rangle}_{\psi_2}^4). \label{eq:mgf-bound} \end{align} We substitute this bound into \eqref{eq:large-deviation-sub-Gaussian-vector-1} to obtain $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le\\ \exp(16c d^2\lambda^2\norm{X}^4_{\psi_2} + \lambda (d\norm{\Sigma}_2 - t^2)), $$ where $\Sigma$ is the covariance matrix of $X$. We optimize over $\lambda$ to obtain $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le e^{-\frac{(t^2 - d\norm{\Sigma}_2)^2}{64c d^2\norm{X}_{\psi_2}^4}}. $$ If the optimum lies outside the region where the m.g.f.\ bound holds \eqref{eq:mgf-bound}, we can always choose $$ \lambda = \frac{c}{4d\norm{X}_{\psi_2}^2} \le \frac{c}{d\norm{\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]}_{\psi_1}} $$ in to obtain the tail bound: $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le e^{c^3 + \frac{c(d\norm{\Sigma}_2 - t^2)}{4d\norm{X}_{\psi_2}^2}}. $$ We combine these two bounds to obtain $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le c_1e^{-\min\left\{\frac{c_2(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2},\frac{(t^2 - d\norm{\Sigma}_2)^2}{64c_2\norm{X}_{\psi_2}^4}\right\}}. $$ Note that the $t^2$ bound always holds. However, for small $t$, the $t^4$ term yields a tighter bound. \end{proof} \section{Proof of Lemma~{\protect{\lowercase{\ref{lemma:hatr_convergence}}}} (Weak Conv.~of $\hat{r}$)}\label{app:hatr} \begin{proof} We expand $\hat{r}_n - r$ (and neglect higher order terms) to obtain \begin{align} \norm{\hat{r}_n - r}_2 &= \Bigl\|\frac{1}{n}\sum_{i=1}^n\hat{\Sigma}^{-1} Y_i(X_i - \hat{\mu}) - \Sigma^{-1}\Expect[s(X)(X-\mu)]\Bigr\|_2\nonumber \\ &\le \norm{\hat{\Sigma}^{-1}}_2\Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \hat{\mu}) - \Expect[s(X)(X-\mu)]\Bigr\|_2 \nonumber\\ &\hspace{1pc}+\|\Expect[s(X)(X-\mu)]\|_2\|\hat{\Sigma}^{-1} - \Sigma^{-1}\|_2 + (o_P(1))^2. \label{eq:r-1} \end{align} The higher order terms generically look like \begin{equation} \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]}\abs{Y - \Expect[Y]} > \epsilon). \label{eq:composite-terms} \end{equation} We apply the union bound to deduce \begin{align*} &\ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]}\abs{Y - \Expect[Y]} > \epsilon) \\ &\hspace{1pc}\le \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \sqrt{\epsilon}) + \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \sqrt{\epsilon}). \end{align*} For any $\epsilon < 1$, $\sqrt{\epsilon} > \epsilon$ and we have $$ \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \sqrt{\epsilon}) \le \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \epsilon). $$ Since terms of the form $\ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \epsilon)$ appear in the upper bounds we derive, we can handle terms like \eqref{eq:composite-terms} with a constant factor (say 2). Since our bounds involve multiplicative constant factors anyways, we neglect these terms to simplify our derivation. We expand the first term to obtain \begin{align*} &\Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \hat{\mu}) -\Expect[Y(X-\mu)]\Bigr\|_2 \\ &\hspace{1pc}\le |\Expect[s(X)]|\norm{\hat{\mu} - \mu}_2 + \norm{\mu}_2\Bigl|\frac{1}{n}\sum_{i=1}^n Y_i - \Expect[s(X)]\Bigr|_2 \\ &\hspace{1pc}\pc+ \Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \mu) - \Expect[s(X)(X-\mu)]\Bigr\|_2 \\ &\hspace{1pc}\pc + (o_P(1))^2. \end{align*} $\hat{\mu} - \mu$ is a sub-Gaussian random variable with sub-Gaussian norm $\frac{\norm{X}_{\psi_2}}{\sqrt{n}}$, so there exist universal $c_1$ and $c_2$ s.t.\ $$ \ensuremath{\mathbf{Pr}}(\norm{\hat{\mu} - \mu}_2 > t) \le c_1\exp\left(-\frac{c_2n(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2}\right). $$ $Y$ is bounded between 1 and -1, so \begin{enumerate} \item We can use Chernoff's inequality to deduce $$ \ensuremath{\mathbf{Pr}}\biggl(\Bigl|\frac{1}{n}\sum_{i=1}^n Y_i - \Expect[s(X)]\Bigr| > t \bigr) \le 2\exp(-nt^2/2). $$ \item $Y_i(X_i - \mu)$ are sub-Gaussian. Thus there exist universal $c_1$ and $c_2$ such that \begin{align*}\begin{split} \ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \mu) - \Expect[s(X)(X - \mu)]\Bigr\|_2 > t \bigr) \le\\ c_1\exp\left(-\frac{c_2n(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2}\right).\end{split} \end{align*} \end{enumerate} We expand the second term in \eqref{eq:r-1} to obtain \begin{align*} \|\hat{\Sigma}^{-1} - \Sigma^{-1}\| = \|\Sigma^{-1/2}\|\|\Sigma^{1/2}\hat{\Sigma}^{-1}\Sigma^{1/2} - I\|\|\Sigma^{-1/2}\|. \end{align*} We expand the middle term to obtain \begin{align*} &\|\Sigma^{-1/2}\hat{\Sigma}\Sigma^{-1/2} - I\| \\ &\hspace{1pc}\le \Bigl\|\Bigl(\frac1n\sum_{i=1}^n \Sigma^{-1/2}(X_i - \mu)(X_i - \mu)^T\Sigma^{-1/2}\Bigr)^{-1} - I\Bigr\|_2 \\ &\hspace{1pc}\pc+2\norm{\mu}_2\norm{\hat{\mu} - \mu}_2 + (o_P(1))^2 \end{align*} We use Theorem 5.39 in \citet{vershynin2010introduction} to bound the first term: $$ \ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n W_iW_i^T - I\Bigr\|_2 > t\Bigr) \le 2\exp(-c_1'(\sqrt{n}t - c_2'\sqrt{d})^2), $$ where $c_1',c_2'$ depend on the sub-Gaussian norm of $W$. We substitute these bounds into our expression for $\|\hat{r} - r\|_2$ to deduce $$\textstyle \ensuremath{\mathbf{Pr}}(\|\hat{r}\! -\! r\|_2 \ge \epsilon) \le Ce^{-\min\left\{\frac{c_1n\epsilon^2}{d},\bigl(c_1'\sqrt{n}\epsilon - c_2'\sqrt{d}\bigr)^2\right\} },$$ where $C$ is an absolute constant and $c_1, c_1',c_2'$ depend on the sub-Gaussian norm of $X$. \end{proof} \section{Proof of Lemma {\protect{\lowercase{\ref{lem:q-large-deviation}}}} (Weak Conv.~of $\hat{Q}$) \label{app:lemmQ} \begin{comment} \begin{lemma}\label{zlema} Let $Z = Y\ensuremath{\mathop{\mathrm{sgn}}}(\gamma,X)$. If $\gamma$ lies in the positive quadrant spanned by $u$ and $v$, then $$\ensuremath{\mathbf{Pr}}[Z=+1\mid X=x] = \begin{cases}1, &\text{if }\ensuremath{\mathop{\mathrm{sgn}}}(\<u,x\>\<v,x\>)>0\\ 0.5, &\text{o.w.} \end{cases}.$$ \end{lemma} \begin{proof} Observe that, by Lemma \ref{lemma:quad}, $\<\gamma,x\>$ is positive when both $\<u,x\>>0$ and $\<v,x\>>0$ and negative when both $\<u,x\><0$ and $\<v,x\><0$. Moreover, by definition, $Y$ is $+1$ and $-1$, respectively, in each of these cases, with probability 1, so for $x$ such that $\ensuremath{\mathop{\mathrm{sgn}}}(\<u,x\>\<v,x\>)>0$ it is indeed true that $Z=+1$. On the other hand, if $\<u,x\>$ and $\<v,x\>$ have different signs, $Y_i=+1$ with probability 0.5. Hence, irrespectively of the sign of $\<\gamma,x\>$, $Z=+1$ with probability 0.5. \end{proof} \end{comment} Let $\tilde{r}$ denote the projection of $r$ onto $U$. \begin{lemma} \label{lem:r-lemma} If $$\|\hat{r} - r\|_2 \le \epsilon_0 = \min\,\{\alpha_1,\ldots,\alpha_k\}\sigma_{\min}(U),$$ then $\tilde{r}$ lies in the interior of the positive cone spanned by $(u_\ell)_{\ell\in [k]}$, where $\sigma_{\min}(U)$ is the smallest nonzero singular value of $U$. \end{lemma} \begin{proof} $r$ lies in interior of the conic hull of $\{u_1,\dots,u_k\}$, so we can express $r$ as $\sum_{i=1}^k\alpha_iu_i$, where $r_i > 0$. If $\tilde{r}$ also lies in the conic hull, then $\tilde{r} = \sum_{i=1}^k\beta_iu_i$ for some $\beta_i > 0$. Then \begin{align*} &\norm{\tilde{r} - r}_2 = \|\sum_{i=1}^k(\alpha_i - \beta_i)u_i\|_2 = \sqrt{(\alpha - \beta)^TUU^T(\alpha - \beta)} \\ &\ge \norm{\alpha - \beta}_2 \sigma_{\min}(U) \ge \norm{\alpha - \beta}_\infty \sigma_{\min}(U). \end{align*} To ensure $\beta$ is component-wise positive, we must have $\norm{\alpha - \beta}_\infty < \min\{\alpha_1,\dots,\alpha_k\}$. A sufficient condition is $\norm{\tilde{r} - r}_2 \leq \epsilon_0 \le \min\{\alpha_1,\dots,\alpha_k\}\sigma_{\min}(U)$. \end{proof} We are now ready to prove Lemma 3. We expand $\|\hat{Q}_n - Q\|_2$ (and neglect higher order terms) to obtain \begin{align*} \|\hat{Q}_n &- Q\|_2 \le\\& \Bigl\|\frac1n\sum_{i=1}^nZ_i\Sigma^{-1/2}(X_i - \mu)(X_i - \mu)^T\Sigma^{-1/2} - Q\Bigr\|_2 \\ &\hspace{1pc} + 2\|\hat{\Sigma}^{-1/2} - \Sigma^{-1/2}\|_2\Expect\bigl[\|(X-\mu)(X-\mu)\Sigma^{-1/2}\|_2\bigr] \\ &\hspace{1pc} + 2\|\Sigma^{-1/2}\|_2\|\hat{\mu} - \mu\|_2\Expect\bigl[\|(X-\mu)\Sigma^{-1/2}\|_2\bigr] \\ &\hspace{1pc} + (o_P(1))^2. \end{align*} The second and third terms can be bounded using the same bounds used in the analysis of how fast $\hat{r}$ converges to $r$. Thus we focus on how fast $$ \sum_{i=1}^n Z_i\Sigma^{-1/2}(X_i - \mu)(X_i - \mu)^T\Sigma^{-1/2} $$ coverges to $Q$. Let $\epsilon'=\min(\epsilon_0,\frac{r}{2})$. First, we note that \begin{align} &\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n Z_iW_iW_i^T - Q\Bigr\|_2 > t \Bigr)\label{bayes} \\ &\hspace{1pc}\le \ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n Z_iW_iW_i^T - Q\Bigr\|_2 > t\mid\hat{r}\in B_{\epsilon'}(r) \Bigr) \ensuremath{\mathbf{Pr}}(\hat{r}\in B_{\epsilon'}(r))\nonumber \\ &\hspace{1pc}\pc+ \ensuremath{\mathbf{Pr}}(\hat{r}\notin B_{\epsilon'}(r)).\nonumber \end{align} Let $\tilde{Z}_i$ denote the ``corrected'' version of the $Z_i$'s, {i.e.}\ the $Z_i$'s we obtain if we use the projection of $\hat{r}$ onto $U$ to flip the labels, and $W_i$ denote $\Sigma^{-1/2}(X_i- \mu)$. We have \begin{align} \label{eq:corrected-z} &\Bigl\|\frac1n\sum_{i=1}^n Z_iW_iW_i^T - Q\Bigr\|_2 \\ &\hspace{1pc}\le \Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 + \Bigl\|\frac1n\sum_{i=1}^n (Z_i-\tilde{Z}_i)W_iW_i^T\Bigr\|_2.\nonumber \end{align} The probability the first term is large is bounded by \begin{align} &\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 > t \mid \hat{r}\in B_{\epsilon'}(r)\Bigr)\label{eq:first-term-large} \\ &\hspace{1pc}\le \sup_{\hat{r}\in B_{\epsilon_0}(r)}\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 > t\mid\hat{r}\in B_{\epsilon'}(r)\Bigr)\nonumber \end{align} The $Z_i$'s are independent of $W_i$'s because the $Z_i$'s were computed using $\hat{r}$ that was in turn computed independently of the $X_i$'s, as the former are computed on a different partition of $[n]$. Thus, the sum in the r.h.s.~of \eqref{eq:corrected-z} is a sum of \emph{i.i.d.}\ r.v. and can be bounded by \begin{align*} &\sup_{\hat{r}\in B_{\epsilon_r}(r)}\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 > t\mid\hat{r}\in B_{\epsilon'}(r)\bigr) \\ &\hspace{1pc}\le 2\exp(-c_1(\sqrt{n}t - c_2\sqrt{d})^2), \end{align*} where $c_1,c_2$ depend on the sub-Gaussian norm of $\tilde{Z}W$. This is a consequence of Remark 5.40 in \citet{vershynin2010introduction}. We now focus on bounding the second term in \eqref{eq:corrected-z}. In what follows, without loss of generality, we will restrict $W$ to the k+1 subspace spanned by $U$ and $\hat{r}$, as remaining components of the $W_i$'s do not contribute to the computation. Let $C_{\hat{r}}$ (for cone) be the ``bad'' region, {i.e.}\ the region where $Z \ne \tilde{Z}$. We have $$ \Bigl\|\frac1n\sum_{i=1}^n (Z_i-\tilde{Z}_i)W_iW_i^T\Bigr\|_2 =\Bigl\|\frac2n\sum_{i=1}^n \mathbf 1_{C_{\hat{r}}}(W_i)W_iW_i^T\Bigr\|_2. $$ By the triangle inequality, we have \begin{align} &\Bigl\|\frac2n\sum_{i=1}^n \mathbf 1_{C_{\hat{r}}}(W_i)W_iW_i^T\Bigr\|_2 \nonumber \\ &\hspace{1pc} \le 2\norm{\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T]}_2 + \Bigl\|\frac2n\sum_{i=1}^n &\mathbf 1_{C_{\hat{r}}}(W_i)W_iW_i^T\nonumber\\&\hspace{1pc}\pc- 2\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T] \Bigr\|_2 \label{eq:triangle-inequality} \end{align} $\mathbf 1_{C_{\hat{r}}}$ is bounded, hence $\mathbf 1_{C_{\hat{r}}}(W_i)W_i$ is sub-Gaussian and \begin{align} \ensuremath{\mathbf{Pr}}\bigg(\Bigl\|\frac2n\sum_{i=1}^n &\mathbf 1_{C_{\hat{r}}}(W_i)W_iW_i^T\nonumber\\&- 2\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T] \Bigr\|_2> t\mid\hat{r}\in B_{\epsilon'}(r)\bigg)\nonumber \\ &\hspace{1pc}\le 2\exp(-c_1(\sqrt{n}t - c_2\sqrt{d})^2), \label{eq:high-prob-C} \end{align} where $c_1,c_2$ depend on the sub-Gaussian norm of $X$. It remains to bound $\norm{\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T]}_2$. We use Jensen's inequality to obtain \begin{align} &\hspace{1pc}\norm{\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T]}_2 \le \Expect[\mathbf 1_{C_{\hat{r}}}(W)\norm{WW^T}_2] \nonumber\\ &\hspace{1pc}\le C\ensuremath{\mathbf{Pr}}(W\in C_{\hat{r}}), \label{eq:expression-C-P} \end{align} where the constant $C$ depends on $k$, as $W$ is restricted to the space spanned by $U$ and $\hat{r}$. Finally, we bound $\ensuremath{\mathbf{Pr}}(W\in C_{\hat{r}})$. The distribution of $W_i$ is spherically symmetric so the probability that $W_i\in C_{\hat{r}}$ is proportional to the surface area of the set $C_{\hat{r}}\cap \mathbf{S}^k$, where $\mathbf{S}^k$ the unit sphere centered at the origin: $$ C_{\hat{r}}\cap \mathbf{S}^k = \{w\in\mathbf{S}^k\mid \hat{r}^Tw \le 0, u_1^Tw, \dots, u_k^Tw \ge 0\}. $$ Lemma~\ref{lemma:quad} implies that this set is contained in the set \[ S = \{w\in\mathbf{S}^k\mid \hat{r}^Tw \le 0, r^Tw \ge 0\}. \] By a symmetry argument, the volume of $S$ (according to the normalized measure on the unit sphere) is simply the angle between $\hat{r}$ and $r$, {i.e.} \[ \ensuremath{\mathbf{Pr}}(W\in C_{\hat{r}}) \le \arccos(\frac{\hat{r}^Tr}{\|\hat{r}\|\|r\|}) \] Let $\epsilon_r\equiv\norm{\hat{r} - r}$. Recall that $\epsilon_r\leq r/2$, by conditioning; it can be verified that this implies $\frac{\hat{r}^Tr}{\|\hat{r}\|\|r\|}\ge 1 - \frac{2\epsilon_r}{\min( \|r\|,\|r\|^2)}$. Thus \[ \ensuremath{\mathbf{Pr}}(W\in C_{\hat{r}}) \le \arccos(1-c_{\|r\|} \epsilon_r) \] where $c_{\|r\|}$ depends on $\|r\|$. We substitute these bounds into \eqref{eq:expression-C-P} to obtain \begin{align} \norm{\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T]}_2 \le C \arccos(1-c_{\|r\|}{\epsilon_r}).\label{crbound} \end{align} We combine \eqref{crbound} with \eqref{eq:high-prob-C} to deduce \begin{align*} &\ensuremath{\mathbf{Pr}}(\Bigl\|\frac1n\sum_{i=1}^n (Z_i-\tilde{Z}_i)W_iW_i^T\Bigr\|_2 > C\arccos(1-c_{\|r\|}\epsilon_r) + t\mid\hat{r}\in B_{\epsilon'}(r)) \\ &\hspace{1pc}\le 2\exp(-c_1(\sqrt{n}t - c_2\sqrt{d})^2) \end{align*} Using this expression and \eqref{eq:first-term-large}, a union bound on Ineq.~\eqref{eq:corrected-z} gives: \begin{align} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} \!- \!Q\|_2 > C\arccos\big(1 \!-\!c_{\|r\|} \epsilon_r\big) + t\mid\hat{r}\in B_{\epsilon'}(r))\nonumber \\ &\hspace{1pc}\le C\exp\left(-\min\left\{\frac{c_1nt^2}{d},\bigl(c_1'\sqrt{n}t - c_2'\sqrt{d}\bigr)^2\right\}\right).\label{weird} \end{align} for appropriate constants $C$,$c_1$,$c_1'$,$c_2'$. Note that \begin{align*} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon\mid\hat{r}\in B_{\epsilon'}(r))\ensuremath{\mathbf{Pr}}(\hat{r}\in B_{\epsilon'}(r)) \\ &\le \ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon \mid C\arccos(1-c_{\|r\|}\epsilon_r)\leq \frac{\epsilon}{2}, \hat{r}\in B_{\epsilon'}(r)) \\ &+ \ensuremath{\mathbf{Pr}}(\norm{\hat{r} - r}_2 > \frac{1}{c_{\|r\|}}(1- \cos(\epsilon/C))), \end{align*} Let $$ f(\epsilon) := \min\left\{\frac{c_1n\epsilon^2}{d},\bigl(c_1'\sqrt{n}\epsilon - c_2'\sqrt{d}\bigr)^2\right\}. $$ The first term is bounded by $C\exp\left(-f(\epsilon/2)\right)$ and the second term is bounded by $C\exp\left(-f( \frac{1}{c_{\|r\|}}(1- \cos(\epsilon/C)) )\right)$. We deduce \begin{align*} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon\mid\hat{r}\in B_{\epsilon'}(r))\ensuremath{\mathbf{Pr}}(\hat{r}\in B_{\epsilon'}(r)) \\ &\hspace{1pc}\le C(\exp(-f( \frac{1}{c_{\|r\|}}(1- \cos(\epsilon/C)))) + \exp(-f(\epsilon/2)))\\ &\hspace{1pc}\le C\exp(-f(\min\{ \frac{1}{c_{\|r\|}}(1- \cos(\epsilon/C)) ,\epsilon/2\})). \end{align*} To complete the proof of Lemma \ref{lem:q-large-deviation}, one can similarly account for the event $\norm{\hat{r} - r}_2 \ge \epsilon'$ in \eqref{bayes}, finaly yielding: \begin{align*} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon) \\ &\hspace{1pc}\le C\exp(-f(\min\{ \frac{1}{c_{\|r\|}}(1- \cos(\epsilon/C)) ,\epsilon/2, \epsilon'\})). \end{align*} The lemma thus follows from the fact that $1-\cos(\epsilon) \approx \frac{1}{2}\epsilon^2$ for small enough $\epsilon>0$. \begin{comment} \section{Proof of Lemma~\ref{lemma:clneg}}\label{app:clneg} Lemma~\ref{lemma:quad} implies that $\tilde{u}_\ell = v_\ell+c \tilde{r}$ for some $c>0$, so, for $\mu=0$, \begin{align*} \tilde{c}_\ell = \mathrm{Cov}[ g(\langle v_\ell, W \rangle + c \langle \tilde{r},W \rangle)\ensuremath{\mathop{\mathrm{sgn}}}(\tilde{r},W) ; \langle W,v_\ell \rangle^2 ]. \end{align*} Let $X = \langle v_\ell, W \rangle $ and $Y =\langle \tilde{r},W\rangle $; then, $X$ and $Y$ are independent Gaussians with zero mean. Moreover, $\tilde{c}_\ell$ can be written as $\tilde{c}_\ell = \mathrm{Cov}[F(X);X^2]$ where $F(x) = \ensuremath{\mathbb{E}}_Y[g(x+cY)\ensuremath{\mathop{\mathrm{sgn}}}(Y)]$. The symmetry assumption~\eqref{fprops} implies that $g$ is antisymmetric, i.e., $g(-x) = -g(x)$. This implies that $F(-x)=\ensuremath{\mathbb{E}}_Y[g(-x+cY)\ensuremath{\mathop{\mathrm{sgn}}}(Y)] \stackrel{Y'\equiv -Y}{=} \ensuremath{\mathbb{E}}_{Y'}[g(-x-cY')\ensuremath{\mathop{\mathrm{sgn}}}(-Y')]= F(x),$ i.e., $F$ is symmetric. Further, \begin{align*} F'(x) &= \ensuremath{\mathbb{E}}_y[g'(x+cY)\ensuremath{\mathop{\mathrm{sgn}}}(Y)]\\ &= \int_0^\infty (g'(x+cy)- g'(x-cy))\frac{e^{-y^2/2\sigma_Y}}{\sigma \sqrt{2\pi}}dy\end{align*} The strict concavity of $g$ in $[0,\infty)$ implies that $g'$ is decreasing in $[0,+\infty)$, and the antisymmetry of $g$ implies that $g'$ is symmetric. Take $x>0$: if $x>cy\geq 0$, $g'(x+cy)> g'(x-cy)$, while if $x\leq cy$, then $g'(x-cy) = g'(cy -x) > g'(cy+x) $, so $F'(x)$ is negative for $x>0$. By the symmetry of $F$, $F'(x)$ is positive for $x<0$. As such, $F(x) =G(x^2)$ for some strictly decreasing $G$, and $\tilde{c}_\ell = \mathrm{Cov}(G(Z);Z)$ for $Z=X^2$; hence, $\tilde{c}_\ell<0$. \qed \end{comment} \section{Principal Hessian Directions}\label{app:phd} In this section, we apply the principal Hessian directions (pHd) \citep{li1992principal} method to our setting, and demontrate its failure to discover the space spanned by parameter profile when $\mu =0$. Recall that pHd considers a setting in which features $X_i\in \ensuremath{\mathbb{R}}^d$ are i.i.d.~and normally distributed with mean $\mu$ and covariance $\Sigma$, while labels $Y_i\in R$ lie, in expectation, in a $k$-dimensional manifold. In particular, some smooth $h:\ensuremath{\mathbb{R}}^k\to\ensuremath{\mathbb{R}}$, $k\ll d$: $$\ensuremath{\mathbb{E}}[Y\mid X= x] = h(\<u_1,x\>,\ldots,\<u_k,x\>)$$ where $u_\ell\in \ensuremath{\mathbb{R}}^d$, $\ell \in [k]$. The method effectively creates an estimate $$\hat{H} = n^{-1}\sum_{i=1}^nY_i\, \Sigma ^{-\frac12}X_iX_i^T\Sigma^{-\frac12}\in\ensuremath{\mathbb{R}}^{d\times d} $$ of the Hessian \begin{align}\begin{split}H=\ensuremath{\mathbb{E}}[\nabla_x^2 h(\<u_1,X\>,\ldots,\<u_k,X\>)] =\\ U^T \ensuremath{\mathbb{E}}[\nabla^2_v h(\<u_1,X\>,\ldots,\<u_k,X\>)] U,\end{split}\label{pHd:hessian} \end{align} where $\nabla_v^2h$ is the Hessian of the mapping $v\mapsto h(v)$, for $v\in \ensuremath{\mathbb{R}}^k$, and $U$ the matrix of profiles. As in our case, the method discovers $\linspan(u_1,\ldots,u_k)$ from the eigenvectors of the eigenvalues that ``stand out'', after appropriate rotation in terms of $\Sigma$. Unfortunately, in the case of linear classifiers, $$h(v) = \sum_{\ell=1}^kp_\ell g(v_\ell) $$ for $g(s) = 2f(s)-1$ is anti-symmetric. As a result, $\nabla^2_v h$ is a diagonal matrix whose $\ell$-th entry in the diagonal is $g''(v_\ell)$. Since $g$ is anti-symmetric, so is $g''$. Hence, if $\mu = 0$ we have that $\ensuremath{\mathbb{E}}[g''(\langle u_\ell,X \rangle)]=0$; hence, the Hessian $H$ given by \eqref{pHd:hessian} will in fact be zero, and the method will fail to discover any signal pertaining to $\linspan(U)$. This calls for an application of the pHd method to a transform of the labels $Y$. This ought to be non-linear, as an affine transform would preserve the above property. Moreover, given that these labels are binary, polynomial transforms do not add any additional signal to $Y$, and are therefore not much help in accomplishing this task. In contrast, the `mirrorring' approach that we propose provides a means of transforming the labels so that their expectation indeed carries sufficient information to extract the span $U$, as evidenced by Theorem~1. \subsection{Technical contribution and related work} Our approach is related to the principal Hessian directions (pHd) method proposed by \citet{li1992principal} and further developed by \citet{cook1998principal} and co-workers. PHd is an approach to dimensionality reduction and data visualization. It generalizes principal component analysis to the regression (discriminative) setting, whereby each data point consists of a feature vector $X_i\in\ensuremath{\mathbb{R}}^d$ and a label $Y_i\in\ensuremath{\mathbb{R}}$. Summarizing, the idea is to form the `Hessian' matrix $\hat{H} = n^{-1}\sum_{i=1}^nY_i\, X_iX_i^T\in\ensuremath{\mathbb{R}}^{d\times d}$. (We assume here, for ease of exposition, that the $X_i$'s have zero mean and unit covariance.) The eigenvectors associated to eigenvalues with largest magnitude are used to identify a subspace in $\ensuremath{\mathbb{R}}^d$ onto which to project the feature vectors $X_i$'s. Unfortunately, the pHd approach fails in general for the mixture models of interest here, namely, mixtures of linear classifiers. For instance, it fails when each component of \eqref{eq:GeneralMixtureRegression} is described by a logistic model $f(y_i=+1| z) = (1+e^{-z})^{-1}$, when features are centered at ${\mathbb E}(X_i)=0$; a proof can be found in \techrep{the extended version of this paper~\citep{full_version}.}{Appendix~\ref{app:phd}.} Our approach overcomes this problem by constructing $\hat{Q} = n^{-1}\sum_{i=1}^nZ_i\, X_iX_i^T\in\ensuremath{\mathbb{R}}^{d\times d}$. The $Z_i$'s are pseudo-labels obtained by applying a `mirroring' transformation to the $Y_i$'s. Unlike with $\hat{H}$, the eigenvector structure of $\hat{Q}$ enables us to estimate the span of $u_1,\dots,u_k$. As an additional technical contribution, we establish non-asymptotic bounds on the estimation error that allow to characterize the trade-off between the data dimension $d$ and the sample size $n$. In contrast, rigorous analysis on pHd is limited to the low-dimensional regime of $d$ fixed as $n\to\infty$. It would be interesting to generalize the analysis developed here to characterize the high-dimensional properties of pHd as well. \subsection{Model} Consider a dataset comprising $n$ i.i.d.~pairs $(X_i,Y_i)\in \ensuremath{\mathbb{R}}^d\times \{-1,+1\}$, $i\in [n]$. We refer to the vectors $X_i\in \ensuremath{\mathbb{R}}^d$ as \emph{features} and to the binary variables as \emph{labels}. We assume that the features $X_i\in\ensuremath{\mathbb{R}}^d$ are sampled from a Gaussian distribution with mean $\mu\in \ensuremath{\mathbb{R}}^d$ and a positive definite covariance $ \Sigma\in \ensuremath{\mathbb{R}}^{d\times d}$. The labels $Y_i\in \{-1,+1\}$ are generated by a \emph{mixture of linear classifiers}, i.e., \begin{align}\label{eq:mix} \ensuremath{\mathbf{Pr}}(Y_i=+1\mid X_i) = \textstyle \sum_{\ell=1}^k p_\ell\, f(\<u_\ell,X_i\>) \, . \end{align} Here, $k\ge 2$ is the number of components in the mixture; $(p_{\ell})_{\ell\in [k]}$ are the weights, satisfying of course $p_{\ell}>0$, $\sum_{\ell=1}^kp_{\ell} = 1$; and $(u_\ell)_{\ell\in [k]}$, $u_{\ell}\in\ensuremath{\mathbb{R}}^d$ are the normals to the planes defining the $k$ linear classifiers. We refer to each normal $u_\ell$ as the \emph{parameter profile} of the $\ell$-th classifier; we assume that the profiles $u_\ell$, $\ell\in[k]$, are linearly independent, and that $k<n/2$. We assume that the function $f:\ensuremath{\mathbb{R}}\to [0,1]$, characterizing the classifier response, is analytic, non-decreasing, strictly concave in $[0,+\infty)$, and satisfies: \begin{align}\label{fprops}\lim_{t\to\infty}\!f(t)\!=\!1, ~~\lim_{t\to -\infty}\!f(t)\!=\!0, \quad 1\!-\!f(t)\!=\!f(-t).\end{align} As an example, it is useful to keep in mind the logistic function $f(t) = (1+e^{-t})^{-1}$. Fig.~\ref{figure:3dplots}(a) illustrates a mixture of $k=2$ classifiers over $d=3$ dimensions. \subsection{Subspace Estimation, Prediction and Clustering} Our main focus is the following task: \begin{quote}\textbf{Subspace Estimation:} After observing $(X_i,Y_i),$ $i\in [n]$, estimate the subspace spanned by the profiles of the $k$ classifiers, {i.e.}, $U \equiv {\rm span}(u_1,\dots,u_k)$. \end{quote} For $\widehat{U}$ an estimate of $U$, we characterize performance via the \emph{principal angle} between the two spaces, namely $$d_P(U,\widehat{U}) = \max_{ x\in U, y \in \widehat{U}} \arccos\left(\tfrac{\<x,y\>}{\|x\|\|y\|}\right).$$ Notice that projecting the features $X_i$ on $U$ entails no loss of information w.r.t.~\eqref{eq:mix}. This can be exploited to improve the performance of several learning tasks through dimensionality reduction, by projecting the features to the estimate of the subspace $U$. Two such tasks are: \begin{quote}\textbf{Prediction}: Given a new feature vector $X_{n+1} $, predict the corresponding label $Y_{n+1}$. \end{quote} \begin{quote}\textbf{Clustering}: Given a new feature vector and label pair $(X_{n+1},Y_{n+1})$, identify the classifier that generated the label. \end{quote} As we will see in Section~\ref{sec:experiments}, our subspace estimate can be used to significantly improve the performance of both prediction and clustering. \subsection{Technical Preliminary} \input{technical} \section{Introduction}\label{sec:intro} \input{intro} \section{Problem Formulation}\label{sec:problem} \input{model} \section{Subspace Estimation}\label{sec:algorithm} \input{algorithm} \section{Proof of Theorem~\ref{thm:main}}\label{sec:proof} \input{proof} \section{Experiments}\label{sec:experiments} \input{experiments} \section{Conclusions}\label{sec:conclusions} \input{conclusions} \section{A Large-Deviation Lemma} We first prove a Bernstein-type inequality for sub-Gaussian random vectors, that we shall use in our proofs: \begin{lemma} \label{lem:large-deviation-sub-Gaussian-vector} Let $X\in\ensuremath{\mathbb{R}}^d$ be a sub-Gaussian random vector, {i.e.}\ $\langle a,X\rangle$ is sub-Gaussian for any $a\in\ensuremath{\mathbb{R}}^d$. Then there exist universal constants $c_1,c_2$ such that \begin{align*} \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t)& \le\\ c_1\exp\bigg(-\min&\left\{\frac{c_2(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2},\frac{(t^2 - d\norm{\Sigma}_2)^2}{64c_2\norm{X}_{\psi_2}^4}\right\}\bigg). \end{align*} \end{lemma} \begin{proof} By the (exponential) Markov inequality, we have \begin{equation}\begin{split} \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) = \ensuremath{\mathbf{Pr}}(\exp(\lambda\norm{X}_2^2) \ge \exp(\lambda t^2))\\ \le \frac{\Expect[\exp(\lambda\norm{X}_2^2)]}{\exp(\lambda t^2)}.\end{split} \label{eq:large-deviation-sub-Gaussian-vector-1} \end{equation} Let $Z$ be uniformly distributed on the unit sphere $\mathbf{S}^{d-1}$. Then $\sqrt{d}Z$ is isotropic so $d\Expect[\langle Z,a\rangle^2] = \norm{a}_2^2$ for all $a$. We use this fact to bound the m.g.f. of $\norm{X}_2^2$: \begin{align*} \Expect[\exp(\lambda\norm{X}_2^2)] &= \Expect_X[\exp(\lambda d\Expect_{Z}[\langle Z,X\rangle^2])] \\ &\le \Expect_X[\Expect_{Z}[\exp(\lambda d\langle Z,X\rangle^2)]]. \end{align*} We interchange the order of expectation to obtain \begin{align*} \Expect[\exp(\lambda\norm{X}_2^2)] &\le \Expect_Z[\Expect_{X}[\exp(\lambda d\langle X,Z\rangle^2)]] \\ &\le \sup\,\{\Expect_{X}[\exp(\lambda d\langle X,Z\rangle^2)]\mid z\in\mathbf{S}^{d-1}\} . \end{align*} $\langle X, z\rangle$ is sub-Gaussian (for fixed $z$) so $\langle X, z\rangle^2$ is (noncentered) sub-exponential. If $\lambda d < c/\norm{\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]}_{\psi_1}$, then \begin{align} &\Expect[\exp(\lambda d\langle X, z\rangle^2)]\nonumber \\ &\hspace{1pc}\le \exp(\lambda d\Expect[\langle X, z\rangle^2])\Expect[\exp(\lambda d(\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]))] \nonumber\\ &\hspace{1pc}\le \exp(\lambda d\Expect[\langle X, z\rangle^2] + c d^2\lambda^2\norm{\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]}_{\psi_1}^2)\nonumber \\ &\hspace{1pc}\le \exp(\lambda d\Expect[\langle X, z\rangle^2] + 4c d^2\lambda^2\norm{\langle X, z\rangle^2}_{\psi_1}^2) \nonumber \\ &\hspace{1pc}\le \exp(\lambda d\Expect[\langle X, z\rangle^2] + 16c d^2\lambda^2\norm{\langle X, z\rangle}_{\psi_2}^4). \label{eq:mgf-bound} \end{align} We substitute this bound into \eqref{eq:large-deviation-sub-Gaussian-vector-1} to obtain $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le \exp(16c d^2\lambda^2\norm{X}^4_{\psi_2} + \lambda (d\norm{\Sigma}_2 - t^2)), $$ where $\Sigma$ is the covariance matrix of $X$. We optimize over $\lambda$ to obtain $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le \exp\left(-\frac{(t^2 - d\norm{\Sigma}_2)^2}{64c d^2\norm{X}_{\psi_2}^4}\right). $$ If the optimum lies outside the region where the m.g.f.\ bound holds \eqref{eq:mgf-bound}, we can always choose $$ \lambda = \frac{c}{4d\norm{X}_{\psi_2}^2} \le \frac{c}{d\norm{\langle X, z\rangle^2 - \Expect[\langle X, z\rangle^2]}_{\psi_1}} $$ in to obtain the tail bound: $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le \exp\left(c^3 + \frac{c(d\norm{\Sigma}_2 - t^2)}{4d\norm{X}_{\psi_2}^2}\right). $$ We combine these two bounds to obtain $$ \ensuremath{\mathbf{Pr}}(\norm{X}_2 \ge t) \le c_1\exp\left(-\min\left\{\frac{c_2(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2},\frac{(t^2 - d\norm{\Sigma}_2)^2}{64c_2\norm{X}_{\psi_2}^4}\right\}\right). $$ Note that the $t^2$ bound always holds. However, for small $t$, the $t^4$ term yields a tighter bound. \end{proof} \section{Proof of Lemma~1 (Weak Convergence of $\hat{r}$)} \begin{proof} We expand $\hat{r}_n - r$ (and neglect higher order terms) to obtain \begin{align} \norm{\hat{r}_n - r}_2 &= \Bigl\|\frac{1}{n}\sum_{i=1}^n\hat{\Sigma}^{-1} Y_i(X_i - \hat{\mu}) - \Sigma^{-1}\Expect[s(X)(X-\mu)]\Bigr\|_2 \\ &\le \norm{\hat{\Sigma}^{-1}}_2\Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \hat{\mu}) - \Expect[s(X)(X-\mu)]\Bigr\|_2 \\ &\hspace{1pc}+\|\Expect[s(X)(X-\mu)]\|_2\|\hat{\Sigma}^{-1} - \Sigma^{-1}\|_2 + (o_P(1))^2. \label{eq:r-1} \end{align} The higher order terms generically look like \begin{equation} \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]}\abs{Y - \Expect[Y]} > \epsilon). \label{eq:composite-terms} \end{equation} We apply the union bound to deduce \begin{align*} &\ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]}\abs{Y - \Expect[Y]} > \epsilon) \\ &\hspace{1pc}\le \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \sqrt{\epsilon}) + \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \sqrt{\epsilon}). \end{align*} For any $\epsilon < 1$, $\sqrt{\epsilon} > \epsilon$ and we have $$ \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \sqrt{\epsilon}) \le \ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \epsilon). $$ Since terms of the form $\ensuremath{\mathbf{Pr}}(\abs{X - \Expect[X]} > \epsilon)$ appear in the upper bounds we derive, we can handle terms like \eqref{eq:composite-terms} with a constant factor (say 2). Since our bounds involve multiplicative constant factors anyways, we neglect these terms to simplify our derivation. We expand the first term to obtain \begin{align*} &\Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \hat{\mu}) -\Expect[Y(X-\mu)]\Bigr\|_2 \\ &\hspace{1pc}\le |\Expect[s(X)]|\norm{\hat{\mu} - \mu}_2 + \norm{\mu}_2\Bigl|\frac{1}{n}\sum_{i=1}^n Y_i - \Expect[s(X)]\Bigr|_2 \\ &\hspace{1pc}\pc+ \Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \mu) - \Expect[s(X)(X-\mu)]\Bigr\|_2 \\ &\hspace{1pc}\pc + (o_P(1))^2. \end{align*} $\hat{\mu} - \mu$ is a sub-Gaussian random variable with sub-Gaussian norm $\frac{\norm{X}_{\psi_2}}{\sqrt{n}}$, so there exist universal $c_1$ and $c_2$ s.t.\ $$ \ensuremath{\mathbf{Pr}}(\norm{\hat{\mu} - \mu}_2 > t) \le c_1\exp\left(-\frac{c_2n(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2}\right). $$ $Y$ is bounded between 1 and -1, so \begin{enumerate} \item We can use Chernoff's inequality to deduce $$ \ensuremath{\mathbf{Pr}}\biggl(\Bigl|\frac{1}{n}\sum_{i=1}^n Y_i - \Expect[s(X)]\Bigr| > t \bigr) \le 2\exp(-nt^2/2). $$ \item $Y_i(X_i - \mu)$ are sub-Gaussian. Thus there exist universal $c_1$ and $c_2$ such that \begin{align*}\begin{split} \ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac{1}{n}\sum_{i=1}^n Y_i(X_i - \mu) - \Expect[s(X)(X - \mu)]\Bigr\|_2 > t \bigr) \le\\ c_1\exp\left(-\frac{c_2n(t^2 - d\norm{\Sigma}_2)}{4d\norm{X}_{\psi_2}^2}\right).\end{split} \end{align*} \end{enumerate} We expand the second term in \eqref{eq:r-1} to obtain \begin{align*} \|\hat{\Sigma}^{-1} - \Sigma^{-1}\| = \|\Sigma^{-1/2}\|\|\Sigma^{1/2}\hat{\Sigma}^{-1}\Sigma^{1/2} - I\|\|\Sigma^{-1/2}\|. \end{align*} We expand the middle term to obtain \begin{align*} &\|\Sigma^{-1/2}\hat{\Sigma}\Sigma^{-1/2} - I\| &\hspace{1pc}\le \Bigl\|\Bigl(\frac1n\sum_{i=1}^n \Sigma^{-1/2}(X_i - \mu)(X_i - \mu)^T\Sigma^{-1/2}\Bigr)^{-1} - I\Bigr\|_2 \\ &\hspace{1pc}\pc+2\norm{\mu}_2\norm{\hat{\mu} - \mu}_2 + (o_P(1))^2 \end{align*} We use Theorem 5.39 in \cite{vershynin2010introduction} to bound the first term: $$ \ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n W_iW_i^T - I\Bigr\|_2 > t\Bigr) \le 2\exp(-c_1'(\sqrt{n}t - c_2'\sqrt{d})^2), $$ where $c_1',c_2'$ depend on the sub-Gaussian norm of $W$. We substitute these bounds into our expression for $\|\hat{r} - r\|_2$ to deduce $$\textstyle \ensuremath{\mathbf{Pr}}(\|\hat{r} - r\|_2 \ge \epsilon) \le C\exp\left(-\min\left\{\frac{c_1n\epsilon^2}{d},\bigl(c_1'\sqrt{n}\epsilon - c_2'\sqrt{d}\bigr)^2\right\}\right), $$ where $C$ is an absolute constant and $c_1, c_1',c_2'$ depend on the sub-Gaussian norm of $X$. \end{proof} \section{Proof of Lemma 3 (Weak Convergence of $\hat{Q}$) \begin{comment} \begin{lemma}\label{zlema} Let $Z = Y\ensuremath{\mathop{\mathrm{sgn}}}(\gamma,X)$. If $\gamma$ lies in the positive quadrant spanned by $u$ and $v$, then $$\ensuremath{\mathbf{Pr}}[Z=+1\mid X=x] = \begin{cases}1, &\text{if }\ensuremath{\mathop{\mathrm{sgn}}}(\<u,x\>\<v,x\>)>0\\ 0.5, &\text{o.w.} \end{cases}.$$ \end{lemma} \begin{proof} Observe that, by Lemma \ref{lemma:quad}, $\<\gamma,x\>$ is positive when both $\<u,x\>>0$ and $\<v,x\>>0$ and negative when both $\<u,x\><0$ and $\<v,x\><0$. Moreover, by definition, $Y$ is $+1$ and $-1$, respectively, in each of these cases, with probability 1, so for $x$ such that $\ensuremath{\mathop{\mathrm{sgn}}}(\<u,x\>\<v,x\>)>0$ it is indeed true that $Z=+1$. On the other hand, if $\<u,x\>$ and $\<v,x\>$ have different signs, $Y_i=+1$ with probability 0.5. Hence, irrespectively of the sign of $\<\gamma,x\>$, $Z=+1$ with probability 0.5. \end{proof} \end{comment} Let $\tilde{r}$ denote the projection of $r$ onto $\linspan\{u,v\}$. \begin{lemma} If $\|\hat{r} - r\|_2 \le \epsilon_0 = \min\,\{\alpha_1,\alpha_2\}\sin(\theta)$, then $\tilde{r}$ also lies in the positive quadrant. \end{lemma} \begin{proof} $r$ lies in interior of the conic hull of $\{u_1,\dots,u_k\}$, so we can express $r$ as $\sum_{i=1}^k\alpha_iu_i$, where $r_i > 0$. If $\tilde{r}$ also lies in the conic hull, then $\tilde{r} = \sum_{i=1}^k\beta_iu_i$ for some $\beta_i > 0$. \begin{align*} \epsilon_0 &= \norm{\tilde{r} - r}_2 = \norm{\sum_{i=1}^k(\alpha_i - \beta_i)u_i}_2 = \sqrt{(\alpha - \beta)UU^T(\alpha - \beta)} \\ &\ge \norm{\alpha - \beta}_2 \sigma_{\min}(U) \ge \norm{\alpha - \beta}_\infty \sigma_{\min}(U). \end{align*} To ensure $\beta$ is component-wise positive, we must have $\norm{\alpha - \beta}_\infty < \min\{\alpha_1,\dots,\alpha_k\}$. A sufficient condition is $\epsilon_0 \le \min\{\alpha_1,\dots,\alpha_k\}\sigma_{\min}(U)$. \end{proof} We are now ready to prove Lemma 3. We expand $\|\hat{Q}_n - Q\|_2$ (and neglect higher order terms) to obtain \begin{align*} \|\hat{Q}_n - Q\|_2 &\le \Bigl\|\frac1n\sum_{i=1}^nZ_i\Sigma^{-1/2}(X_i - \mu)(X_i - \mu)^T\Sigma^{-1/2} - Q\Bigr\|_2 \\ &\hspace{1pc} + 2\|\hat{\Sigma}^{-1/2} - \Sigma^{-1/2}\|_2\Expect\bigl[\|(X-\mu)(X-\mu)\Sigma^{-1/2}\|_2\bigr] \\ &\hspace{1pc} + 2\|\Sigma^{-1/2}\|_2\|\hat{\mu} - \mu\|_2\Expect\bigl[\|(X-\mu)\Sigma^{-1/2}\|_2\bigr] \\ &\hspace{1pc} + (o_P(1))^2. \end{align*} The second and third terms can be bounded using the same bounds used in the analysis of how fast $\hat{r}$ converges to $r$. Thus we focus on how fast $$ \sum_{i=1}^n Z_i\Sigma^{-1/2}(X_i - \mu)(X_i - \mu)^T\Sigma^{-1/2} $$ coverges to $Q$. First, we note that \begin{align*} &\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n Z_iW_iW_i^T - Q\Bigr\|_2 > t \Bigr) \\ &\hspace{1pc}\le \ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n Z_iW_iW_i^T - Q\Bigr\|_2 > t\mid\hat{r}\in B_{\epsilon_0}(r) \Bigr) \\ &\hspace{1pc}\pc+ \ensuremath{\mathbf{Pr}}(\hat{r}\in B_{\epsilon_0}(r)). \end{align*} Let $\tilde{Z}_i$ denote the ``corrected'' version of the $Z_i$'s, {i.e.}\ the $Z_i$'s we obtain if we use the projection of $\hat{r}$ onto the $\linspan\{u,v\}$ to flip the labels, and $W_i$ denote $\Sigma^{-1/2}(X_i- \mu)$. We have \begin{align} &\Bigl\|\frac1n\sum_{i=1}^n Z_iW_iW_i^T - Q\Bigr\|_2 \\ &\hspace{1pc}\le \Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 + \Bigl\|\frac1n\sum_{i=1}^n (Z_i-\tilde{Z}_i)W_iW_i^T\Bigr\|_2. \label{eq:corrected-z} \end{align} The probability the first term is large is bounded by \begin{align*} &\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 > t \mid \hat{r}\in B_{\epsilon_0}(r)\Bigr) \\ &\hspace{1pc}\le \sup_{\hat{r}\in B_{\epsilon_0}(r)}\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 > t\mid\hat{r}\in B_{\epsilon_0}(r)\Bigr) \end{align*} The $Z_i$'s are independent of $W_i$'s because the $Z_i$'s were computed using $\hat{r}$ that was in turn computed independently of the $X_i$'s. Thus this is a sum of \emph{i.i.d.}\ r.v. and we can bound it by \begin{align*} &\sup_{\hat{r}\in B_{\epsilon_r}(r)}\ensuremath{\mathbf{Pr}}\biggl(\Bigl\|\frac1n\sum_{i=1}^n \tilde{Z}_iW_iW_i^T - Q\Bigr\|_2 > t\mid\hat{r}\in B_{\epsilon_0}(r)\bigr) \\ &\hspace{1pc}\le 2\exp(-c_1(\sqrt{n}t - c_2\sqrt{d})^2), \end{align*} where $c_1,c_2$ depend on the sub-Gaussian norm of $\tilde{Z}W$. This is a consequence of Remark 5.40 in \cite{vershynin2010introduction}. We now focus on bounding the second term in \eqref{eq:corrected-z}. Let $W_i$ denote the spherically symmetric (whitened) version of $X$. We restrict ourselves to the 3d subspace spanned by $u,v,\hat{r}$. Let $C_{\hat{r}}$ (for cone) be the ``bad'' region, {i.e.}\ the region where $Z \ne \tilde{Z}$. We have $$ \Bigl\|\frac1n\sum_{i=1}^n (Z_i-\tilde{Z}_i)W_iW_i^T\Bigr\|_2 =\Bigl\|\frac2n\sum_{i=1}^n \mathbf 1_{C_{\hat{r}}}(W_i)W_iW_i^T\Bigr\|_2. $$ $\mathbf 1_{C_{\hat{r}}}$ is bounded, hence $\mathbf 1_{C_{\hat{r}}}(W_i)W_i$ is sub-Gaussian and \begin{align*} &\ensuremath{\mathbf{Pr}}\bigg(\Bigl\|\frac2n\sum_{i=1}^n \mathbf 1_{C_{\hat{r}}}(W_i)W_iW_i^T- 2\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T] \Bigr\|_2> t\mid\hat{r}\in B_{\epsilon_0}(r)\big) \\ &\hspace{1pc}\le 2\exp(-c_1(\sqrt{n}t - c_2\sqrt{d})^2), \end{align*} where $c_1,c_2$ depend on the sub-Gaussian norm of $X$. It remains to bound $\norm{\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T]}_2$. We use Jensen's inequality to obtain \begin{align*} &\hspace{1pc}\norm{\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T]}_2 \le \Expect[\mathbf 1_{C_{\hat{r}}}(W)\norm{WW^T}_2] \\ &\hspace{1pc}\le C\ensuremath{\mathbf{Pr}}(W\in C_{\hat{r}}), \end{align*} where the constant $C$ depends on the number of mixture components. Finally, we bound $\ensuremath{\mathbf{Pr}}(W\in C_{\hat{r}})$. The distribution of $W_i$ is spherically symmetric so the probability that $W_i\in C_{\hat{r}}$ is proportional to the surface area of the set $$ S = \{w\in\mathbf{S}^{k+1}\mid \hat{r}^Tw \le 0, u_1^Tw, \dots, u_k^Tw \ge 0\}. $$ This set is contained in the set $$ \bar{S} = \{w\in \mathbf{S}^{k+1}\mid \norm{w - u_1}_2 \le \norm{\hat{r} - r}_2\} $$ This set is nothing but the spherical cap of radius $\norm{\hat{r} - r}_2$ in $k+1$ dimensions. An upper bound is $$ \gamma(\bar{S}) \le \exp(-(k+1)\cos(\frac{2-r^2}{2ab})^2), $$ where $\gamma$ denotes the uniform measure on the unit sphere. We substitute these bounds into our expression for the second term to obtain \begin{align*} &\hspace{1pc}\Expect[\mathbf 1_{C_{\hat{r}}}(W)WW^T] \\ &\hspace{1pc}\le C\arccos\left(\frac{2 - \epsilon_r^2}{2}\right)(1 - \cos\angle(u,v))(1 + \cos\angle(u,v) + \\ &\hspace{1pc}\pc\cos^2\angle(u,v)) \\ &\textstyle\hspace{1pc}\le C\arccos\left(1 - \frac{\epsilon_r^2}{2}\right). \end{align*} We combines these bounds to deduce \begin{align*} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > C\arccos\left(1 - \frac{\epsilon_r^2}{2}\right) + \epsilon\mid\hat{r}\in B_{\epsilon_0}(r)) \\ &\hspace{1pc}\le C\exp\left(-\min\left\{\frac{c_1n\epsilon^2}{d},\bigl(c_1'\sqrt{n}\epsilon - c_2'\sqrt{d}\bigr)^2\right\}\right). \end{align*} Finally, we have \begin{align*} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon\mid\hat{r}\in B_{\epsilon_0}(r)) \\ &\hspace{1pc}\le \ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon \mid \norm{\hat{r} - r}_2 \le \sqrt{1 - \cos(\epsilon/2)}, \hat{r}\in B_{\epsilon_0}(r)) \\ &\hspace{1pc}\pc+ \ensuremath{\mathbf{Pr}}(\norm{\hat{r} - r}_2 \le \sqrt{1 - \cos(\epsilon/2)}). \end{align*} Let $$ f(\epsilon) := \min\left\{\frac{c_1n\epsilon^2}{d},\bigl(c_1'\sqrt{n}\epsilon - c_2'\sqrt{d}\bigr)^2\right\}. $$ The first term is bounded by $C\exp\left(-f(\epsilon/2)\right)$ and the second term is bounded by $C\exp\left(-f(\sqrt{1 - \cos(\epsilon/2)})\right)$. We deduce \begin{align*} &\ensuremath{\mathbf{Pr}}(\|\hat{Q} - Q\|_2 > \epsilon\mid\norm{\hat{r} - r}_2 \le \epsilon_0) \\ &\hspace{1pc}\le C(\exp(-f(\sqrt{1-\cos(\epsilon/2)})) + \exp(-f(\epsilon/2)))\\ &\hspace{1pc}\le C\exp(-f(\min\{\sqrt{1-\cos(\epsilon/2)},\epsilon/2\})). \end{align*} \section{Principal Hessian Directions} In this section, we apply the principal Hessian directions (pHd) \cite{li1992principal} method to our setting, and demontrate its failure to discover the space spanned by parameter profile when $\mu =0$. Recall that pHd considers a setting in which features $X_i\in \ensuremath{\mathbb{R}}^d$ are i.i.d.~and normally distributed with mean $\mu$ and covariance $\Sigma$, while labels $Y_i\in R$ lie, in expectation, in a $k$-dimensional manifold. In particular, some smooth $h:\ensuremath{\mathbb{R}}^k\to\ensuremath{\mathbb{R}}$, $k\ll d$: $$\ensuremath{\mathbb{E}}[Y\mid X= x] = h(\<u_1,x\>,\ldots,\<u_k,x\>)$$ where $u_\ell\in \ensuremath{\mathbb{R}}^d$, $\ell \in [k]$. The method effectively creates an estimate $$\hat{H} = n^{-1}\sum_{i=1}^nY_i\, \Sigma ^{-\frac12}X_iX_i^T\Sigma^{-\frac12}\in\ensuremath{\mathbb{R}}^{d\times d} $$ of the Hessian \begin{align}\begin{split}H=\ensuremath{\mathbb{E}}[\nabla_x^2 h(\<u_1,X\>,\ldots,\<u_k,X\>)] =\\ U^T \ensuremath{\mathbb{E}}[\nabla^2_v h(\<u_1,X\>,\ldots,\<u_k,X\>)] U,\end{split}\label{pHd:hessian} \end{align} where $\nabla_v^2h$ is the Hessian of the mapping $v\mapsto h(v)$, for $v\in \ensuremath{\mathbb{R}}^k$, and $U$ the matrix of profiles. As in our case, the method discovers $\linspan(u_1,\ldots,u_k)$ from the eigenvectors of the eigenvalues that ``stand out'', after appropriate rotation in terms of $\Sigma$. Unfortunately, in the case of linear classifiers, $$h(v) = \sum_{\ell=1}^kp_\ell g(v_\ell) $$ for $g(s) = 2f(s)-1$ is anti-symmetric. As a result, $\nabla^2_v h$ is a diagonal matrix whose $\ell$-th entry in the diagonal is $g''(v_\ell)$. Since $g$ is anti-symmetric, so is $g''$. Hence, if $\mu = 0$ we have that $\ensuremath{\mathbb{E}}[g''(\langle u_\ell,X \rangle)]=0$; hence, the Hessian $H$ given by \eqref{pHd:hessian} will in fact be zero, and the method will fail to discover any signal pertaining to $\linspan(U)$. This calls for an application of the pHd method to a transform of the labels $Y$. This ought to be non-linear, as an affine transform would preserve the above property. Moreover, given that these labels are binary, polynomial transforms do not add any additional signal to $Y$, and are therefore not much help in accomplishing this task. In contrast, the `mirrorring' approach that we propose provides a means of transforming the labels so that their expectation indeed carries sufficient information to extract the span $U$, as evidenced by Theorem~1.
\section{Introduction} \label{Intro} In a recent work Papamikos and Robnik \cite{PapRob2012} have studied time dependent nonlinear Hamiltonian oscillators from the point of view of their statistical properties, in order to generalize a series of studies on the time dependent linear oscillator by Robnik and Romanovski \cite{RR2006a,RR2006b,RRS2006,KR2007,RR2008}, where the rigorous WKB method has been employed \cite{RR2000}. We are interested in the time evolution of a microcanonical ensemble of initial conditions. If the evolution is ideal adiabatic (i.e. infinitely slow), then the adiabatic invariant, which is also the action of the system, or the area inside the contour of constant energy in the phase space (divided by $2\pi$), is conserved, and this is precisely the adiabatic theorem on one-dimensional Hamiltonian systems \cite{Arnold}, provided we do not cross a separatrix during the adiabatic process. What happens if the changing of the system parameter is not adiabatic? For the linear oscillator with an arbitrary time dependence it has been shown rigorously in the above mentioned papers (see in particular the review \cite{RR2008}) that the value of the adiabatic invariant at the average value of the energy during the evolution is always increasing. Since the adiabatic invariant is proportional to the number of states, this implies an irreversibility in the mean, because the entropy is the logarithm of the number of states, in the sense of statistical mechanics, as explained in section 2. This finding was a motivation to analyze the {\em nonlinear} oscillators from this point of view. In \cite{PapRob2012} it was shown using the numerical techniques (highly accurate symplectic integrators of 8th order, \cite{M1995,McL2002,HLW2006,LeimReich,SenzSernaCalvo, SY1996,Y1990,Y1993,P2011}) that for slow but {\em not} adiabatic changes in a quartic oscillator the adiabatic invariant at the mean energy can {\em decrease}, just due to the nonlinearity and nonisochronicity. However, for sufficiently fast changes the property is restored, especially in the extreme case of a {\em parametric kick}, when the system's parameter jumps discontinuously. This led us to demonstrate analytically by a rigorous calculation for the case of homogeneous power law potentials (of which the quartic oscillator and the harmonic oscillators are two special cases), that the adiabatic invariant at the average final energy indeed increases under a parametric kick. Therefore, we \cite{PapRob2012} have put forward a Conjecture (henceforth called {\em PR Conjecture}) that the adiabatic invariant for an initial microcanonical ensemble at the mean energy always increases under a parametric kick. The purpose of the present work is to investigate the validity and the conditions under which it is true. We show by a series of case studies, that the PR Conjecture is largely true, except if either the potential is not smooth enough, or if the energy is too close to a stationary point of the potential (close to a separatrix in the phase space). The latter complications are not unexpected, because existence of a separatrix in the phase space always complicates matters, e.g. implies violation of the adiabatic theorem, because the averaging method does not work there, since the period of oscillation is infinite. Also, breaking the smoothness properties of the potential obviously can break "communication" between the different parts of the potential well. A very recent review of the main ideas has been published in \cite{Rob2013}. In a more general context, time dependent Hamiltonian systems are very interesting and important dynamical models, where many interesting questions about their statistical behaviour can be studied \cite{Arnold, Lochak, Zaslavsky, Ott}. The time dependence of the Hamilton function describes, or models, the interaction of the system with the environment. Whilst the energy of the system is not conserved, the Liouville theorem of course still applies and thus the phase space volume is preserved by the flow. One of the major questions is the time evolution of the energy of certain ensembles of initial conditions. The microcanonical ensemble is the most fundamental, like in statistical mechanics, and we investigate how it develops in time. In the ideal adiabatic processes, which are infinitely slow, the adiabatic invariant is conserved, and using this conservation law we can calculate the sharply defined energy changing in time. If the process is non-adiabatic, having a finite speed of changing the Hamilton function, the energy will be spread around its mean value. In the linear oscillator \cite{RR2006a}-\cite{RR2008} the distribution function of the energy is universal, independent of the driving law of the frequency as a function of time, and is given by the arcsine distribution. In nonlinear systems this universality is lost, and the evolution of the energy distribution can exhibit a rich variety of behaviour. Once we have the energy distribution, we can calculate distribution of other dynamical quantities, in particular of the adiabatic invariant. In particular in time periodic (Floquet) systems we can find a very rich behaviour, from integrability to full chaoticity (ergodicity), and also the scenario in between, namely the case of a mixed phase space, even in 1D systems. One example is the kicked rotator (standard map) and many other time periodic systems \cite{Chirikov,Zaslavsky}. More recent works included the periodic parametric kicking of the quartic oscillator \cite{PapRob2012} and the periodic linear driving (sawtooth driving law) of the quartic oscillator \cite{PapSowRob2012}. Time periodic systems are interesting also from the point of view of the Fermi acceleration, including the quantum mechanical counterparts, that is unlimited growth of the energy, in 1D systems and higher dimensions. For some recent works see \cite{BatRob, BatRob2012, Sch2009, Schmelcher} and references therein. Lots of interesting empirical material has accumulated, including the power law behaviours with universal scaling properties \cite{LMS2004}. In this paper we study the parametrically kicked 1D Hamiltonian systems, trying to work out conditions under which the PR Conjecture holds true, showing, as mentioned above, that it is largely satisfied except when the initial energy (of the microcanonical ensemble) is too close to a stationary point of the potential (separatrix in the phase space), or if the potential is not analytic and not sufficiently smooth. For these reasons we shall speak of the PR property, namely that a certain potential behaves in agreement with the PR Conjecture, and thus possesses the PR property, but possibly only for a certain range of energies, or entirely not. The paper is organized as follows. In section 2 we explain the connection to the statistical mechanics in the sense of Gibbs, especially for small number of degrees of freedom, explaining why the Gibbs entropy is fundamental and correct (in contradistinction to the Boltzmann entropy), as emphasized very recently also by Dunkel and Hilbert \cite{DunHil2014}, corroborating the views of Gibbs \cite{Gibbs1902}, Einstein \cite{Ein1911} and Hertz \cite{Her1910}. In section 3 we present the general theory of parametric kicks, in section 4 we analyze the examples where the PR Conjecture is entirely satisfied, in section 5 we analyze the counter examples, where the PR Conjecture is violated or partially violated (that is, its validity applies only to a certain energy range of the potential). In section 6 we discuss the results and conclude. In Appendix A as an overview we summarize the list of potentials with valid or broken PR property. \section{The PR-property and its connection to the statistics in the sense of Gibbs} In statistical mechanics of classical mechanical systems the most fundamental ensemble to calculate the entropy, and thus all other equilibrium properties, is the Gibbs microcanonical ensemble, based on the number of states $\Omega(E)$ {\em inside the (closed) energy surface} of energy $E$, calculated as the phase space volume \begin{equation} \label{Omega} \Omega (E) = \int_{H(q,p)\le E} d^fq\;d^fp. \end{equation} Here $f$ is the number of degrees of freedom. Following Gibbs the entropy, called {\em Gibbs entropy}, in order to distinguish it from other definitions like, e.g. the Boltzmann entropy, is defined as follows \begin{equation} \label{Gibbsentr} S_G (E) = k_B\ln \Omega (E), \end{equation} where $k_B$ is the Boltzmann constant. $\Omega$ has the dimension of the $2f$-dim phase space volume, which is a technical nuisance when taking the logarithm of it. This difficulty can be removed by dividing $\Omega$ by a constant $a$ with the same physical dimension. However, so long as we are interested only in differences of the entropy, which is the case in the classical statistical mechanics, such a constant $a$ drops out from all calculations and thus has no physical significance. Usually, however, the natural choice is $a=(2\pi\hbar)^f$, thus making the definition of $S_G$ compatible with the quantum version, where the entropy is well defined in absolute terms. In this paper we deal only with classical mechanics. The fundamental role of $\Omega$ has been established by Gibbs himself \cite{Gibbs1902}, later discussed by Hertz \cite{Her1910} and Einstein \cite{Ein1911}, and recently corroborated in a critical analysis by Dunkel and Hilbert \cite{DunHil2014}, showing that Gibbs entropy is at variance with the Boltzmann entropy. The latter is defined as the logarithm of the number of states inside an {\em energy shell} around the energy $E$, and differs from the Gibbs entropy, especially in systems with a small number of degrees of freedom $f$ (small systems), such as treated in this paper. It is precisely the Gibbs entropy which gives the right answers and results in small systems. For example, in the case of an ideal monoatomic gas it was shown in \cite{DunHil2014}, regarding e.g. the calculation of the equipartition, that Gibbs definition is the right one, whilst the Boltzmann entropy differs at small $f$, but of course nevertheless agrees with the Gibbs entropy for large systems (large $f$). It has been realized by Gibbs \cite{Gibbs1902}, Hertz \cite{Her1910} and Einstein \cite{Ein1911}, that the fundamental quantity of classical statistical mechanics is $\Omega(E)$, defined in (\ref{Omega}). It is precisely the adiabatic invariant of the system, which is conserved under adiabatic infinitely slow changes, as proven by Paul Hertz \cite{Her1910} for ergodic systems. Also quantum mechanically, $N(E) = \Omega/a$ is precisely the number of quantum eigenstates of a bound system below the energy $E$, which is also the adiabatic invariant in quantum mechanics. In one degree of freedom systems, $f=1$, we have of course $\Omega (E) = 2\pi I (E)$, where $I(E)$ is the classical Hamilton action of the system at energy $E$. Hence the importance of $I(E)$, which we also call adiabatic invariant. On the other hand, if the system (its Hamilton function) depends on time nonadiabatically, meaning having finite speed of changing the parameter, the energy of the system still changes, but now it has a distribution around its mean value $\bar{E}$, and $\Omega(E)$ also depends on time, having a certain distribution, but the most interesting and important question is then under what conditions $\Omega (\bar{E})$ will increase or decrease in time, implying increasing or decreasing Gibbs entropy (at the mean energy), respectively. What is the relevance of 1D Hamiltonian systems in this context? If we have an ensemble, even macroscopic ensemble, of identical 1D {\em noninteracting} systems, the behaviour of the macroscopic ensemble will be obviously very much determined by the behaviour of a single system. One example is the ideal gas. Enclosing particles in a 1D box of length $L$, the behaviour of the 1D gas is governed by the behaviour of one particle interacting with the moving walls, simply due to the absence of interactions between the particles. By calculating $\Omega = 2Lp = 2L \sqrt{2mE}$ for a single particle, we find immediately $E= \frac{1}{2}k_B T$, if $1/T$ is interpreted as $dS_G/dE$. This is precisely the equipartition law. Since $E$ and $S_G$ are additive, it follows immediately $E=nk_BT/2$ for $n$ noninteracting particles in such 1D box. The "temperature" $T$ here can be understood also as the time average of the particle's kinetic energy over sufficiently many oscillation periods, divided by $k_B/2$. All equilibrium properties of the 1D ideal gas can be determined in that way, using $S_G$. Moreover, if $L(t)$ is a function of time, we can calculate the evolution of the energy $E$, for an ensemble of initial conditions, and $\Omega$. This picture can be generalized to the 3D ideal gas whose thermodynamic equation of state can be derived. Similar approach is applicable to the general time dependent 1D nonlinear Hamiltonian oscillators. Of course, the most fundamental is the {\em microcanonical ensemble} of initial conditions, all located on the initial energy contour with sharp energy $E_0$, but distributed uniformly on the torus w.r.t. the canonical angle (the phase). Then, $E$ changing in time is a function of the initial condition, and spreads around its mean value $\bar{E}$, also varying in time. Calculating the $\Omega (\bar{E}) = 2\pi I(\bar{E})$ at the mean energy then enables us to calculate the Gibbs entropy $S_G = S_G(\bar{E})$ as a function of time, and with it thus all the "thermodynamic" properties. By the additivity of $S_G$ such a picture can be generalized to arbitrary 1D time dependent and {\em noninteracting} Hamiltonian systems. It has been shown by Robnik and Romanovski \cite{RR2006a,RR2006b,RRS2006,RR2008} that in the case of the 1D linear oscillator with {\em arbitrary} parametric driving (the frequency $\omega(t)$ is an arbitrary function of time, including a discontinuous jump (=parametric kick)) the $\Omega (\bar{E})$ always increases, and so does the Gibbs entropy $S_G(\bar{E})$, except in adiabatic, infinitely slow processes, where it is exactly conserved. In nonlinear 1D systems this property is in general lost, just due to the nonlinearity. Thus $\Omega (\bar{E})$ and $S_G(\bar{E})$ can decrease. For the adiabatic processes $\Omega(\bar{E})$ is of course exactly conserved, by the theorem on adiabatic invariants \cite{Arnold}, as proven by Paul Hertz \cite{Her1910} for the general ergodic systems of any degrees of freedom, thus including $f=1$. But for slow although not adiabatic processes $\Omega(\bar{E})$ can decrease, just due to the nonlinearity, which was demonstrated by Papamikos and Robnik, as mentioned above. This is an important observation, because it indicates that the nonlinear interactions can lead to the decrease of the entropy of parametrically driven systems, which is not possible in the linear oscillators. Nevertheless, for sufficiently fast parametric driving, we intuitively expect that the PR property holds true and thus the law of the increasing entropy is restored. This is precisely what we observe, and the extreme case is the case of parametric kicking, being the fastest possible change. This is the main motivation of the present work, where we systematically explore by a number of case studies when the PR property is valid or not. One should bear in mind that the parametric kicking is a good, leading, approximation of systems which undergo very fast changes of the parameter, on time scales of less than one oscillation period, as has been also demonstrated by Papamikos and Robnik mentioned above. Of course, there is a number of open questions, in fact a whole programme of research in this direction: What can we say about general parametric driving of nonlinear 1D Hamiltonian systems? How can we treat {\em collectively} an ensemble of noninteracting identical 1D nonlinear driven oscillators, by calculating $\Omega$ as a function of time? Furthermore, how can we generalize all results for driven higher dimensional oscillators, first for a single system, and then collectively for an ensemble of identical noninteracting systems? It turns out that already the first step in this programme, namely the case of the parametrically kicked 1D systems, is difficult enough, dealt with in this paper. \section{General theory of parametric kicks in 1D Hamiltonian systems} \label{General} We consider Hamiltonian systems with one degree of freedom in the form quadratic kinetic energy (except for the subsection \ref{hompot}, where the kinetic energy is more general) plus potential, where the potential depends linearly on the parameter $A$, which is the system's control parameter, \begin{equation} \label{generalhamiltonian} H(p,q) = \frac{p^{2}}{2} - Af(q),\quad A>0. \end{equation} Sometimes we shall use also the notation for the potential $V(q)=-Af(q)$. The time dependence that we study is an instantaneous jump of $A$, from $A_0$ to $A_1$. Thus, the initial and the final Hamilton functions are \begin{eqnarray} \label{inifinHam} H_0= H(q,p,A_0) &=& \frac{p^{2}}{2} - A_0 f(q), \nonumber\\ H_1= H(q,p,A_1) &=& \frac{p^{2}}{2} - A_1 f(q). \end{eqnarray} We assume the {\em microcanonical ensemble of initial conditions}, which means that the initial energy $E_0$ is sharply defined, and the distribution of the initial conditions is uniform with respect to the canonical angle variable $\theta$, which means that the density of points on an infinitesimal interval on the energy contour $H_0=E_0$ is proportional to the length of time spent in that interval under the dynamics of the initial but frozen Hamiltonian $H_0$. Thus, for an observable $F(q,p)$, the final $(q,p)$ are functions of $E_0$, $\theta$ and time, and the average at time $t$ is defined as follows \begin{equation} \label{average} \langle F \rangle (E_0,t) = \frac{1}{2\pi} \int_0^{2\pi} F\left(q(\theta,t),p(\theta,t)\right) d\theta. \end{equation} This can be also written as (suppressing the arguments) \begin{equation} \label{timeaverage} \langle F \rangle = \frac{1}{T_0} \oint F(\tau)\, d\tau = \frac{\oint F(q,p)\, dq_0/p_0}{\oint dq_0/p_0}, \end{equation} where $T_0=\oint dq_0/p_0$ is the period of the oscillation of the initial frozen system, $\tau$ is its time, and $q,p$ are regarded as functions of the initial point $q_0,p_0$ on the energy contour $E_0$, and of time $t$. When the jump $A_0\rightarrow A_1$ takes place, the coordinate in the configuration space $q$ and the canonical conjugate momentum $p$ remain continuous, because $q$ is by definition a continuous variable, whilst $p$ is continous because there is no external kick (Dirac delta function peaked force) acting on the system. These statements are not trivial, because we cannot choose for $q,p$ the action-angle variables, simply because they are not defined in time dependent systems. Indeed, if $p,q$ were action-angle variables $(I,\theta)$, not changing at all, nothing at all would happen in the system, because $H(I,\theta)$ would not change at all. Thus, it is important to realize that we must work in the ordinary phase space $(q,p)$. Also, we must remark that the phase flow in such case is only $C^0$. For the final energy $E_1$, after the jump, we can write \begin{eqnarray} \label{finE1} E_1 &=& \frac{p^{2}}{2} - \frac{A_1}{A_0} A_0 f(q) \nonumber\\ &=& \frac{p^{2}}{2} + \frac{A_1}{A_0} (E_0 - \frac{p^{2}}{2}) \nonumber\\ &=& \frac{p^{2}}{2}(1-x) + E_0 x , \end{eqnarray} where we shall use the notation $x=A_1/A_0$ throughout this paper. So the final energy $E_1$ has some distribution implied by the nature of the initial microcanonical ensemble and the change of the geometry of the phase space, and is now only a function of initial and final $p$ at fixed $x$. We can thus immediately calculate the average value of $E_1$, denoted by $\langle E_1 \rangle$, namely \begin{eqnarray} \label{aveE1-0} \langle E_1 \rangle &=& (1-x)\frac{\langle p^{2} \rangle}{2} + E_0 x = E_0 x + \frac{(1-x)}{2T_0} \oint p^{2} \; dt \nonumber \\ &=& E_0 x + \frac{(x-1)}{2T_0} \oint p \; dq = E_0 x + \frac{(1-x)\pi}{T_0}I_0, \end{eqnarray} where the integration $\oint$ is taken over the entire oscillation cycle, that means from the smaller turning point usually denoted by $q_1$ to the larger one $q_2$, and back. $T_0 = T(E_0,A_0)$ is the period at the initial energy $E_0$ and $I_0 = I(E_0,A_0) $ is the action as a function of the energy $E_0$ and parameter $A_0$. They are generally defined as \begin{equation} \label{Idef} I(E,A) = \frac{1}{2\pi} \oint p\,dq = \frac{1}{\pi}\int_{q_1}^{q_2} dq\, \sqrt{2(E+A f(q))}, \end{equation} and the period of oscillation $T(E,A)$ as \begin{eqnarray} \label{Tdef} T(E,A) &=& \oint dt = \oint \frac{dq}{p} = 2\int_{q_1}^{q_2} \frac{dq}{\sqrt{2(E+A f(q))}} \nonumber \\&=& 2\pi \frac{\partial I(E,A)}{\partial E}. \end{eqnarray} In the following we denote also $I_1=I(E_1,A_1)$ and $T_1=T(E_1,A_1)$. We can also calculate $\langle E_1 \rangle$ in terms of the new action $I_1$ and period $T_1$ evaluated at $E=\langle E_1 \rangle$ and $A_1$, namely as follows \begin{equation} \label{aveE1-1} \langle E_1 \rangle = \frac{\langle p^2 \rangle}{2} + \langle -A_1f(q) \rangle_1 = \pi \frac{I_1}{T_1} -A_1 \langle f(q) \rangle_1 \end{equation} where the averaging is now over the contour $E_1=H(q,p,A_1)$, and thus \begin{equation} \label{avefq} \langle f(q)\rangle_1 = \frac{1}{T_1} \oint_{\langle E_1 \rangle} f(q)\,dt. \end{equation} Using the two expressions for $\langle E_1 \rangle$ we arrive at the expression for the final action $I_1$ at the average final energy \begin{equation} \label{finI} I_1 = I\left(\langle E_1 \rangle, A_1\right)= \frac{T_1}{T_0}I_0 (1-x) + \frac{T_1 E_0 x}{\pi} + \frac{T_1 A_1\langle f(q) \rangle_1}{\pi}, \end{equation} and the average $\langle f(q) \rangle_1$ in (\ref{avefq}) can be further expressed \begin{equation} \label{avefq1} \langle f(q) \rangle_1 = \frac{1}{T_1} \oint_{\langle E_1 \rangle} \frac{f(q) \; dq}{\sqrt{2\left(E + Af(q)\right)}} = \langle f(q) \rangle_1 = \frac{2\pi}{T_1}\frac{\partial I_1}{\partial A_1}. \end{equation} Therefore the final action at the final average energy from (\ref{finI}) is \begin{equation} \label{finIE} I_1 = \frac{T_1}{T_0}I_0 (1-x) + \frac{T_1 E_0 x}{\pi} + 2 A_1 \frac{\partial I_1}{\partial A_1}\vert_{\langle E_1 \rangle}. \end{equation} We can also express the action $I_1$ as, \begin{equation} \label{finI1C} I_1 = \frac{T_1 \langle E_1 \rangle}{\pi} + 2 A_1 \frac{\partial I_1}{\partial A_1}\vert_{\langle E_1 \rangle}. \end{equation} {\bf The Papamikos-Robnik Conjecture} ({\bf PR Conjecture}) is now formulated as $I_1 \ge I_0$, where $I_1$ is given either in (\ref{finI1C}), or in (\ref{finIE}), or in (\ref{finI}), or more explicitly \begin{eqnarray} \label{PRC} &&I_1 = I\left( \langle E_1 \rangle, A_1 \right) = I\left( \frac{\pi}{T_0}I_0 (1-x) + E_0 x,\; xA_0 \right) \nonumber \\ &&I\left( \langle E_1 \rangle, A_1 \right)\geq I\left( E_0,A_0\right), \end{eqnarray} which must be satisfied for all values of $E_0$, $A_0\ge 0$ and $x=A_1/A_0\in[0,\infty)$. \\\\ If the PR property (\ref{PRC}) is satisfied only for some energy range $E_0$ of the potential, we shall say that the PR Conjecture is partially sastisfied, and if is entirely violated, we shall say that the potential does not possess the PR property. Since $x$ is any positive real number, and since in the case $x=1$ nothing happens at all (no parametric kick at all), (\ref{PRC}) must be equality $I_1 = I_0$ for $x=1$, and strict inequality for all other $x$. This Conjecture is difficult or almost impossible to prove in general with available techniques, especially as it is valid under restricted conditions. Nevertheless, we shall prove it rigorously by direct calculations in a very large class of specific systems (potentials), treated in section \ref{PRCY}. Nevertheless, we can do much more in the local analysis, in the sense of investigating when (\ref{PRC}) has a minimum at $x=1$. We define $x= 1 + \epsilon$, where $0< \vert \epsilon \vert \ll 1$, and look at the Taylor expansion in $\epsilon$, \begin{equation} \label{Taylor} I(E,A) = I_0 + L\; \epsilon + Q \; \epsilon^2 + O(\epsilon^3), \end{equation} where the linear term $L$ is, using $\frac{\partial I}{\partial E} =T_0/(2\pi)$, \begin{equation} \label{L} L = \frac{E_0T_0}{2\pi} -\frac{I_0}{2} + A_0 \frac{\partial I}{\partial A}\vert_{(E_0,A_0)}, \end{equation} and the quadratic term $Q$ is equal to: \begin{eqnarray} \label{Q} Q&=&\frac{1}{2}\left\{\frac{\partial^{2}I}{\partial E^{2}}\left(E_{0} - \frac{\pi I_{0}}{T_{0}}\right)^{2} \right. \nonumber\\ &+& \left. 2A_{0} \frac{\partial^{2}I}{\partial E \partial A}\left(E_{0} - \frac{\pi I_{0}}{T_{0}}\right) + A_{0}^{2} \frac{\partial^{2}I}{\partial A^{2}}\right\}, \end{eqnarray} where all partial derivatives must be taken at $E=E_0$ and $A=A_0$. Now we evaluate $L$ and $Q$. We prove that $L=0$, meaning that $x=1$ is a stationary point. The calculation is straightforward as follows \begin{eqnarray} \label{L=0} A \frac{\partial I}{\partial A} &=& \frac{1}{2\pi}\oint \frac{A f(q)}{\sqrt{2E + 2Af(q)}}\; dq = \frac{1}{2\pi}\oint \frac{-E + \frac{p^{2}}{2}}{\sqrt{2E + 2Af(q)}}\; dq \nonumber\\ &=& -\frac{ET}{2\pi} + \frac{1}{4\pi}\oint p\; dq = -\frac{ET}{2\pi} + \frac{I}{2}. \end{eqnarray} Now we calculate $Q$ in (\ref{Q}). First we calculate the second partial derivatives \begin{eqnarray} \label{2ndder} \frac{\partial^{2}I}{\partial E^{2}} &=& \frac{1}{2\pi} \frac{\partial T}{\partial E} ,\\ \frac{\partial^{2}I}{\partial E \partial A} &=& -\frac{T}{4\pi A} -\frac{E}{2\pi A}\frac{\partial T}{\partial E},\\ \frac{\partial^{2}I}{\partial A^{2}} &=& \frac{T E}{2\pi A^{2}} - \frac{I}{4 A^{2}} + \frac{E^{2}}{2\pi A^2} \frac{\partial T}{\partial E} . \end{eqnarray} The first equation above is obvious. The second one follows immediately by using the equality (\ref{L=0}). The third one follows in a similar manner, using (\ref{L=0}) three times and $\partial T/\partial A=2\pi \partial^2 I/\partial A\partial E=2\pi\partial/\partial E (\partial I/\partial A)$. Substituting the above results in (\ref{Q}) we find in a straightforward manner the final expression for $Q$, \begin{equation} \label{Qfin} Q= \frac{I\left(2\pi I\frac{\partial T}{\partial E} + T^{2}\right)}{8 T^{2}}. \end{equation} If we have a minimum of (\ref{Taylor}) then we must have $Q>0$, or from (\ref{Qfin}) \begin{equation} \label{mincond1} \frac{1}{2} + \frac{\pi I}{T^{2}}\frac{\partial T}{\partial E} \geq 0. \end{equation} Using the definition $\omega = 2\pi/T$, and the fact $T = 2\pi \partial I/\partial E$, we find that the above condition is equivalent to \begin{equation} \label{mincond2} \frac{\partial \omega}{\partial E}\leq\frac{1}{I}, \end{equation} or in final form, with a simple geometrical interpretation, \begin{equation} \label{mincond3} \frac{\partial^{2}\left(I^{2}\right)}{\partial E^{2}} \geq 0. \end{equation} Namely, the last inequality (\ref{mincond3}) means that $I^2(E)$ is a positive, monotonically increasing and convex function of $E$ for all $E$ in the range of the definition of $I(E)$. It could be that it is satisfied only for a certain energy range of $E$, in which case we shall say that the system (potential) has the PR property on the underlying/relevant energy range. In the next section \ref{PRCY} we shall study many examples of potentials $V(q)=-Af(q)$, all of them entirely satisfying the PR Conjecture. Potentials having an escape energy will be always defined (without loss of generality) in such a way, by an energy shift, that the escape energy will be equal to zero. The question arises at what value of the parametric kick strength $x=A_1/A_0$ the final average energy $\langle E_1\rangle\ge 0$ leads to the escape. From equation (\ref{aveE1-0}), and reminding that in this case $E_0<0$, we see immediately that escape takes place if $x\le x_{esc}$, \begin{equation} \label{escape} x \le x_{esc} = \frac{1}{1 - \frac{T_0 E_0}{\pi I_0}}. \end{equation} The potentials with a single minimum satisfying the PR Conjecture will be treated in the next section \ref{PRCY}, whilst more complicated case studies will be presented in section \ref{PRCN}. \begin{figure} \center \includegraphics[width=3.5cm]{potentials3.eps} \includegraphics[width=3.5cm]{action3pot.eps} \caption{Sketch of the potentials and the corresponding squared action $I^2(E)$, which is a positive, monotonically increasing convex function, satisfying the PR property.} \label{sketchofpot} \end{figure} In figure \ref{sketchofpot} we sketch the behaviour of various potentials and the associated squared action $I^2(E)$ as a function of the energy $E$, for which the PR property is satisfied for all $E$. \section{Examples of validity of the PR Conjecture} \label{PRCY} In this section we give examples of Hamiltonian systems in which the PR Conjecture (\ref{PRC}) is explicitly and rigorously satisfied for the entire energy range of the potential $V(q)$. \subsection{Family of Hamiltonians with homogeneous potential and homogeneous kinetic energy} \label{hompot} The Hamiltonian is \begin{equation}\label{homohamiltonian} H(p,q)=\frac{p^{2n}}{2n} + \frac{Aq^{2m}}{2m},\quad A>0,\; m>0,\; n>0, \end{equation} where $m$ and $n$ are positive integers. The action is \begin{eqnarray} \label{homoaction} I\left(E\right)&=&\frac{1}{2\pi}\oint p \; dq = \frac{1}{\pi}\int_{-q_{0}}^{q_{0}} \sqrt[2n]{2n\left( E - \frac{Aq^{2m}}{2m}\right)} \; dq = \nonumber \\ &=& \frac{2^{\frac{2nm+m+n}{2mn}}n^{\frac{1}{2n}}m^{\frac{1}{2m}}}{\pi A^{\frac{1}{2m}}2m}B \left(\frac{1}{2m},\frac{2n+1}{2n}\right)E^{\frac{n+m}{2mn}}, \end{eqnarray} where $-q_0$ and $q_0$ are the turning points, and $B$ is the Beta function, \begin{equation} \label{betaf} B \left( x,y \right) = \int_0^{1} t^{x-1}\left( 1-t \right)^{y-1} \; dt. \nonumber \end{equation} The period $T_{n,m}$ is related to the frequency $\omega_{n,m}$, as $T_{n,m}=\frac{2\pi}{\omega_{n,m}}$. The frequency is \begin{eqnarray} \label{freqmn} &&\omega_{n,m}\left( E\right)=\frac{\mathrm{d}E}{\mathrm{d}I}=\left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}, \nonumber \\ &&\frac{dI}{dE}= a(m,n)\;B\left(\frac{1}{2m},\frac{2n+1}{2n}\right)\,E^{\frac{n+m-2mn}{2mn}},\\ &&a(m,n)=\frac{2^{\frac{m+n-2mn}{2mn}}n^{\frac{1}{2n}}m^{\frac{1-4m}{2m}}\left( n+m \right)}{\pi A^{\frac{1}{2m}}}\nonumber \end{eqnarray} Now we calculate the average energy after the kick, denoting the kinetic energy as $K=\frac{p^{2n}}{2n}$, \begin{eqnarray} \langle E_1 \rangle &=& \langle K\rangle + \langle V_1 = \rangle \rangle \nonumber \\ &=& \langle K \rangle + \langle V_0 \rangle - \langle V_{0} \rangle + \langle V_1\rangle\nonumber\\ &=& E_0 + \frac{A_1-A_0}{A_0} \langle V_0 \rangle. \end{eqnarray} We use the virial theorem \cite{LL} for this system, rather than equation (\ref{aveE1-0}), because the kinetic energy $K = \frac{p^{2n}}{2n}$, $n\ge 1$, is more general than quadratic $n=1$ (and this is the only system in this paper in which we have general nonquadratic kinetic energy), and obtain \begin{equation} \langle K \rangle =\frac{m}{n} \langle V_0 \rangle \Rightarrow \langle V_0 \rangle =\frac{n} {m+n} E_0 . \end{equation} The ratio of the actions is \begin{equation}\label{homoactionratio} \frac{I\left(\langle E_1 \rangle\right)}{I(E_0)}=\frac{\left(1+ \left(x-1\right)\frac{n}{n+m}\right)^{\frac{m+n}{2m n}}}{x^{\frac{1}{2m}}}=F_{n,m}(x). \end{equation} The function $F_{n,m}(x)$ has only one minimum at $x=1$ and the value of the function is equal to $F_{n,m}(1)=1$, as stated by the PR Conjecture. \subsection{Pendulum} The Hamiltonian is \begin{equation}\label{pendulumham} H(q,p)= \frac{p^{2}}{2} - \Omega^{2} \cos q. \end{equation} The action is \begin{eqnarray} I\left(E\right)=\frac{1}{2 \pi} \oint p \; dq = \frac{1}{ \pi} \int_{-q_0}^{q_0} \sqrt{2\left(E + \Omega^{2} \cos q\right)} \; dq, \end{eqnarray} where $-q_0$ and $+q_0$ are the two turning points in the case of libration (oscillation), defined by $E+\Omega^2\cos q_0=0$. The action depends on the region in the phase space that we consider. There are three regions of energy $E$ and $q_0$ as follows: \begin{enumerate} \item outside the separatrix, $E>\Omega^{2}$, $q_0=\pi$, \item on the separatrix, $E=\Omega^{2}$, $q_0=\pi$ \item inside the separatrix, $E<\Omega^{2}$, $q_0<\pi$. \end{enumerate} \noindent We denote $k=\sqrt{\frac{E+\Omega^{2}}{2\Omega^{2}}}$, and obtain \begin{equation} I\left(E\right)=\frac{4}{\pi}\sqrt{2(E+\Omega^{2})}\int_{0}^{\frac{q_0}{2}} \sqrt{1-\frac{\sin^{2}\phi}{k^{2}}} \; d\phi. \end{equation} For the first case, $k \ge 1$, we have \begin{equation}\label{pendulumactionout} I\left(E\right)=\frac{8\Omega k}{\pi} \rm{E} \left( \frac{1}{k} \right), \end{equation} where $\rm{E}$ is complete elliptic integral of the second kind, which is defined as \begin{equation} {\rm{E}} \left( k \right)=\int_0^{\frac{\pi}{2}} \sqrt{1-k^{2}\sin^{2}\theta} \; d\theta. \end{equation} For the second case, $k=1$, we have, $I\left(E\right)=\frac{8\Omega}{\pi}$. For the last case, $k \le 1$, we have \begin{equation}\label{pendulumactionin} I\left(E\right)=\frac{8\Omega}{\pi} \left[ {\rm{E}} \left(k\right) -(1-k^{2}) {\rm{K}} \left(k\right) \right], \end{equation} where $\rm{E}$ is the same complete elliptic integral of the second kind and the $\rm{K}$ is elliptic integral of the first kind, which is defined as \begin{displaymath} {\rm{K}}\left(k\right)=\int_0^{\frac{\pi}{2}} \frac{d\theta}{\sqrt{1-k^{2}\sin^{2}\theta}}. \end{displaymath} The period $T(E)$ is related to the frequency $\omega(E)$ as $T=\frac{2\pi}{\omega}$, and the frequency is \begin{displaymath} \omega\left( E\right)=\frac{\mathrm{d}E}{\mathrm{d}I}=\left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}= \frac{\pi \Omega}{2} \times \left\{ \begin{array}{lr} \frac{k}{{\rm{K}}\left(\frac{1}{k}\right)}, & \mbox{} k\in(1,\infty)\\ \frac{1}{{\rm{K}}\left(k\right)}, & \mbox{} k\in[0,1] \\ \end{array} \right. \end{displaymath} For the average energy we also have two cases. For outside the separatrix, $E_0 \geq \Omega_0^{2}$, we get \begin{eqnarray} \langle E_1 \rangle &= E_0 x + \frac{2\Omega_0^2\; k_0\;{\rm{E}}\left(\frac{1}{k_0}\right)}{{\rm{K}}\left(\frac{1}{k_0}\right)}(1-x), \end{eqnarray} whilst for inside the separatrix, $ E_0\le \Omega_0^{2}$, we have \begin{equation} \langle E_1 \rangle = E_0 x + 2\Omega_0^2\left[ \frac{\rm{E}(k_0)}{\rm{K}(k_0)} -\left(1-k_0^{2}\right)\right](1-x), \end{equation} where $k_0=\sqrt{\frac{E_0^{2} + \Omega_0^{2}}{2\Omega_0^{2}}}$. We have calculated the average energy $\langle E_1 \rangle$ from the previous equations, but also checked it numerically. We find that the difference is of the order of about $10^{-13}$. In figure \ref{pendactionrat} we show the action ratio for the pendulum in the libration regime. \begin{figure} \center \includegraphics[width=3.5cm]{actionratioin.eps} \includegraphics[width=3.5cm]{actionratioin2.eps} \caption{Ratio of initial and final actions for the pendulum, inside the separatrix, with $E_0$ is the initial energy and $n$ is the number of the points of initial contour. The bullet is the point $(x_{sep},I(x_{sep}))$, which corresponds to the crossing the separatrix after the kick.} \label{pendactionrat} \end{figure} \noindent For the case of libration the action $I(E)$ as a power series of $\epsilon = E + \Omega^{2}$ reads as \begin{eqnarray} I(E) &= \frac{\epsilon}{4194304 \Omega^{11}} ( 1323 \; \epsilon^{5} + 3920 \Omega^{2}\; \epsilon^{4} + 12800 \Omega^{4}\; \epsilon^{3} \nonumber \\ &+ 49152 \Omega^{6}\; \epsilon^{2} + 262144 \Omega^{8}\; \epsilon + 4194304 \Omega^{10} + h.o.t. ). \nonumber \\ \end{eqnarray} and the function $I^2(E)$ is plotted in figure \ref{I2Epend}. The PR property is satisfied. \begin{figure} \center \includegraphics[width=4.5cm]{actpendexp.eps} \caption{The pendulum: The second derivative $d^2I^{2}(E)/dE^2$, with $\omega=1$. } \label{I2Epend} \end{figure} \subsection{Radial Kepler problem with zero angular momentum} \label{Keplerzero} The Hamiltonian is \begin{equation}\label{keplerpotential} H(r,p)= \frac{p^{2}}{2} - \frac{A}{r},\quad A>0. \end{equation} The action is \begin{eqnarray} \label{kepleraction} I\left(E\right)&=&\frac{1}{2 \pi} \oint p \; dr = \nonumber \\ &=& \frac{1}{\pi} \int_{r_{1}}^{r_{2}} \sqrt{2\left(E + \frac{A}{r}\right)} \; dr = \frac{A}{\sqrt{2\vert E\vert}}, \end{eqnarray} where $r_1$ and $r_2$ are the two turning points. In fact, $r_1=0$. The period $T$ is related to the frequency $\omega$, as $T=\frac{2\pi}{\omega}$, \begin{equation} \omega\left( E\right)=\frac{\mathrm{d}E}{\mathrm{d}I}=\left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}=\frac{2\sqrt{2}\;\vert E \vert ^{\frac{3}{2}}}{A}. \end{equation} The average energy is \begin{equation} \langle E_1 \rangle = \langle K \rangle + \langle V_1 \rangle = E_0 + \frac{A_1 - A_0}{A_0} \langle V_0 \rangle. \nonumber \end{equation} We can use the virial theorem \cite{LL} to calculate the average potential, \begin{equation} \langle E_0 \rangle = \frac{\langle V_0 \rangle}{2} = E_0. \end{equation} Finally, we get the ratio of the actions $ I_1 $ and $ I_0 $ \begin{equation}\label{kepleractionratio} \frac{I\left(\langle E_1 \rangle\right)}{I(E_0)} = \frac{x}{\sqrt{2x - 1}}=F(x), \end{equation} where $ x = A_1/A_0$. For this Kepler problem we have positive average energy $\langle E_1\rangle$ if $A_1 \le A_0/2$. Thus the ratio of the actions $I_1 \geq I_0$, for no-escape orbits, as the function $F(x)$ has the minimum at $x=1$ and the value of the function is $F(1)=1$ in accordance with the PR Conjecture. \subsection{Radial Kepler problem with nonzero angular momentum} The Hamiltonian with the angular momentum $M$ is \begin{equation}\label{keplereffectivepotential} H(r,p)= \frac{p^{2}}{2} - A \left(\frac{a}{r} - \frac{M^{2}}{2r^{2}}\right),\quad A>0,\; a>0,\;M^{2}>0. \end{equation} The action is \begin{eqnarray} \label{keplereffectiveaction} I\left(E\right)&=&\frac{1}{2 \pi} \oint p \; dq =\frac{1}{\pi} \int_{r_{1}}^{r_{2}} \sqrt{2\left(E + A\left(\frac{a}{r} - \frac{M^{2}}{2r^{2}}\right)\right)} \; dr \Rightarrow \nonumber \\ I\left(E\right)&=& \frac{A \; a}{\sqrt{2\vert E\vert}} - \sqrt{A}\; M. \end{eqnarray} Here $r_1$ and $r_2$ are the two turning points. The frequency is \begin{equation} \omega\left( E\right)=\frac{\mathrm{d}E}{\mathrm{d}I}=\left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}= \frac{2\sqrt{2} \; \vert E \vert ^{\frac{3}{2}}}{A\; a}. \end{equation} The average energy is, using equation (\ref{aveE1-0}), \begin{equation} \langle E_1 \rangle = E_0 x + \frac{\sqrt{2}\;\vert E_0 \vert ^{\frac{3}{2}}}{A_0 \;a} \left(\frac{A_0 \; a}{\sqrt{2\;\vert E_0 \vert}} - \sqrt{A_0}\;M\right)(1-x), \end{equation} where $x=A_1/A_0$. \\ We have positive average energy after the kick, if \begin{displaymath} x \leq x_{esc}= \frac{A_0 \; a - \sqrt{2 A_0}\;M\vert E_0 \vert^{\frac{1}{2}}}{2 A_0 \; a - \sqrt{2 A_0}\;M\vert E_0 \vert^{\frac{1}{2}}}. \end{displaymath} For the action ratio we have finally \begin{equation} \frac{I\left(\langle E_1 \rangle\right)}{I\left(E_0\right)} = \frac{x \sqrt{\frac{1}{2x - 1 + (1-x)c}} - c \sqrt{x}}{1 - c}=F(x), \end{equation} where $c=\frac{M\sqrt{2 \vert E_0 \vert}}{a \sqrt{A_0}}$. Since the minimal energy $E_{min} \le 0$ is determined by the condition $I(E_{min})=0$, we find from (\ref{keplereffectiveaction}) $|E_{min}| = a^2A/(2M^2)$, and thus due to $|E_0| \le E_{min}$ we have $c \le 1$. If $M=0$ and consequently $c=0$, we recover the result for the Kepler problem with zero angular momentum of the previous subsection \ref{Keplerzero}. The derivative of $F(x)$ is \begin{eqnarray} \frac{\partial F(x)}{\partial x} = \frac{-\frac{x}{2}(2-c)\left(\frac{1}{B(x)}\right)^{\frac{3}{2}} + \left(\frac{1}{B(x)}\right)^{\frac{1}{2}} - \frac{c}{2\sqrt{x}}}{1-c}, \end{eqnarray} where $B(x) = (2-c)x + c-1$. There is only one root for $\frac{\partial F(x)}{\partial x} = 0$, namely precisely $x = 1$. For the second derivative of $F(x)$ we find \begin{eqnarray} \frac{\partial^{2} F(x)}{\partial x^{2}} = \frac{\frac{3}{4}(c-2)^{2}x\left(\frac{1}{B(x)}\right)^{\frac{5}{2}} + (c-2)\left(\frac{1}{B(x)}\right)^{\frac{3}{2}} + \frac{c}{4 x^{3/2}}}{1-c}, \end{eqnarray} and at $x = 1$ we have \begin{eqnarray} \frac{\partial^{2} F(x)}{\partial x^{2}}\vert_{x=1} = 1 - \frac{3}{4}c > 0 \end{eqnarray} for all $c \le 1$. Thus there is only one minimum for the function $F(x)$ at $x=1$, with $F(1)=1$, exactly in accordance with the PR Conjecture. In figure \ref{actratKepler} we show the action ratio as a function of $x$ for $c=1/2$. \begin{figure} \center \includegraphics[width=3.5cm]{kepler1act.eps} \caption{Action ratio for the Kepler potential with nonzero angular momentum, with $A_0=1$ and $c=\frac{1}{2}$. The $x$ for escape orbit is $x\le x_{esc}=\frac{1}{3}$, with action ratio $I_1/I_0 \rightarrow \infty$ as $x\rightarrow x_{esc}$.} \label{actratKepler} \end{figure} \subsection{Morse potential} \label{Morse} The Hamiltonian is \begin{equation}\label{morsepotential} H(q,p)= \frac{p^{2}}{2} + A(e^{-2 \lambda r} - e^{- \lambda r}),\quad A>0,\quad \lambda>0. \end{equation} The minimum of the potential is equal to $-A/4$. The action is \begin{eqnarray}\label{morseaction} I\left(E\right) = \frac{1}{2 \pi} \oint p \; dr = \frac{1}{\pi} \int_{r_1}^{r_2} \sqrt{2\left(E - A(e^{-2 \lambda r} - e^{- \lambda r})\right)} \; dr, \nonumber \end{eqnarray} where $r_1$ and $r_2$ are the two turning points, and we get \begin{equation} I\left(E\right) = -\frac{\sqrt{2\vert E \vert}}{\lambda} + \frac{\sqrt{A}}{\lambda \sqrt{2}}. \end{equation} The frequency $\omega$ is \begin{equation} \omega\left( E \right)=\frac{\mathrm{d}E}{\mathrm{d}I}= \left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}=\lambda \sqrt{2 \vert E \vert}. \end{equation} For the average energy, we have \begin{equation} \langle E_1 \rangle = \langle K \rangle + \langle V_1 \rangle = E_0 + \frac{A_1-A_0}{A_0}\langle V_0 \rangle. \nonumber \end{equation} In contradistinction to some other systems, here we can find the $r$ as a function of time $t$ in an explicit closed form \begin{eqnarray} \label{morseactionratio} &&\frac{dr}{dt}=\sqrt{2(E-A(e^{-\lambda r} - e^{-2\lambda r}))}\Rightarrow \nonumber \\ && t= \int \frac{dr}{\sqrt{2(E-A(e^{-\lambda r} - e^{-2\lambda r}))}}. \end{eqnarray} Using $u=e^{-\lambda r}$ and after some calculations we obtain \begin{equation} t = -\frac{\arcsin\left( \frac{2E+Au}{u\sqrt{4AE+A^{2}}}\right) }{\lambda\sqrt{2\vert E \vert}}, \end{equation} or, \begin{equation} \label{Morseroft} r(t) = -\frac{1}{\lambda}\ln \left[ \frac{2\vert E \vert}{\sqrt{4EA+A^{2}}\; \sin \left(\sqrt{2\vert E \vert}\;\lambda t\right) +A}\right]. \end{equation} Finally the average energy is, either using the equation (\ref{aveE1-0}) or doing the direct averaging over time using (\ref{Morseroft}), \begin{equation} \label{Morseaverageenergy} \langle E_1 \rangle = E_0 + \frac{\sqrt{A_0 \vert E_0 \vert}}{2}(1-x), \end{equation} where $x=A_1/A_0$. We have positive average energy after the kick, at $x \leq x_{esc}= 1 - 2\vert E_0 \vert/A_0$. For the action ratio we have \begin{equation} \frac{I\left(\langle E_1 \rangle\right)}{I\left(E_0\right)} = \frac{ \sqrt{x} -2 \sqrt{ \vert \frac{E_0}{A_0} + \frac{1}{2}\sqrt{\frac{\vert E_0 \vert}{A_0}}\; (1 - x)\vert}} {1 - 2\sqrt{\frac{\vert E_0 \vert}{A_0}}}=F(x). \end{equation} We define $E_0 = -A_0/4 + \epsilon$, $0\leq \epsilon \leq A_0/4$, and $c=\epsilon/A_0$, so that $0\le c \le 1/4$, and simplify \begin{eqnarray} F(x) = \frac{I\left(\langle E_1 \rangle\right)}{I\left(E_0 \right)} = \frac{\sqrt{x} -2 \sqrt{\vert -c + \frac{1}{4} + \frac{\sqrt{-c + \frac{1}{4}}}{2}\; (1 - x)\vert}}{1 - 2\sqrt{\vert \frac{1}{4}-c \vert}}. \end{eqnarray} The function $F(x)$ has a single minimum at $x=1$ and $F(1)=1$, in agreement with the PR Conjecture. In figure \ref{morsefig} we plot the action ratio as a function of $A_1$ at $\lambda=1$ and $A_0=1$ \begin{figure} \center \includegraphics[width=3.5cm]{morseact.eps} \caption{Action ratio for the Morse potential, with $\lambda=1$. $A_0=1$ and $n$ is the number of points in initial contour. $x$ for escape orbits is $x \le x_{esc}=0.31803$, with the action ratio $I_1/I_0=1.18103$, marked by a bullet.} \label{morsefig} \end{figure} \subsection{P\"oschl-Teller I potential} \label{PTI} The Hamiltonian is \begin{equation} \label{pt1hamiltonian} H(q,p) = \frac{p^{2}}{2} + \frac{A}{\cos^{2} \left( \lambda q \right)},\quad \lambda>0,\; A>0. \end{equation} The action is \begin{eqnarray} \label{pt1action} I\left(E\right) &= \frac{1}{2 \pi} \oint p \; dq = \frac{1}{\pi} \int_{q_1}^{q_2} \sqrt{2\left(E - \frac{A}{\cos^{2} \lambda q}\right)} \; dq, \nonumber \end{eqnarray} where $q_1$ and $q_2$ are the two turning points. Using transformation, $ y=\tan \left(\lambda q\right) $ and after calculation, we get finally, \begin{equation} I\left(E\right) = \frac{\sqrt{2}}{\lambda} (\sqrt{E} - \sqrt{A}). \end{equation} The frequency is \begin{equation} \omega\left( E\right)=\frac{\mathrm{d}E}{\mathrm{d}I}= \left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}=\lambda\sqrt{2E}. \end{equation} Now we calculate the final average energy either using (\ref{aveE1-0}) or by direct averaging \begin{equation} \langle E_1 \rangle = E_0 + \frac{A_1 - A_0}{A_0} \langle V_0 \rangle. \nonumber \end{equation} where \begin{equation} \langle V_0 \rangle = \frac{1}{T} \int_{0}^{T} \frac{A_0}{\cos^{2} \left( \lambda q(t) \right)} \; dt. \end{equation} Here again we can express $q$ as function of time $ t $, namely, after some straightforward calculation, we obtain \begin{equation} q(t) = \frac{1}{\lambda} \arcsin \left( \sqrt{\frac{E-A}{E}} \sin \left( \lambda t\sqrt{2E} \right)\right). \end{equation} Further, using the transformation $ y= \lambda \sqrt{2E_0} t $, we find \begin{eqnarray} \langle V_0 \rangle &=& \frac{1}{T} \int_{0}^{T} \frac{A_0}{\cos^{2} \left( \lambda q(t) \right)} \; dt \nonumber\\ &=& \frac{A_0}{\lambda T \sqrt{2E_0}} \int_{0}^{2\pi} \frac{dy}{1-\frac{E_0 - A_0}{E_0} \sin^{2} y} \Rightarrow \nonumber \\ \langle V_0 \rangle &=& \sqrt{A_0E_0}. \end{eqnarray} The average energy, with $x=A_1/A_0$, is thus \begin{equation} \langle E_1 \rangle = E_0 +(x-1) \sqrt{E_0A_0}. \end{equation} The minimum energy is at $E_0=A_0$, so we write $E_0=A_0+\epsilon$, where $\epsilon\geq 0$. Then, setting $c=\epsilon/A_0$, we obtain the action ratio $F(x)$ as \begin{equation} \frac{I\left(\langle E_1 \rangle\right)}{I\left(E_0\right)}=F(x)= \frac{\sqrt{ c+1 +\sqrt{ c+1 }(x-1)}-\sqrt{x}}{\sqrt{ c+1 }-1}. \end{equation} This function $F(x)$ has a single minimum at $x=1$ and $F(1)=1$, in agreement with the PR Conjecture. \subsection{P\"oschl-Teller II potential} \label{PTII} The Hamiltonian is \begin{equation}\label{pt2hamiltonian} H(q,p) = \frac{p^{2}}{2} - \frac{A}{\cosh^{2} \left(\lambda q\right)},\quad \lambda>0,\; A>0. \end{equation} The action is \begin{eqnarray} \label{pt2action} I\left(E\right) &= \frac{1}{2 \pi} \oint p \; dq = \frac{\sqrt{2}}{\lambda}\left(\sqrt{A}-\sqrt{\vert E \vert}\right). \end{eqnarray} The frequency is \begin{equation} \omega\left( E\right)=\frac{\mathrm{d}E}{\mathrm{d}I}=\left(\frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1} =\lambda\sqrt{2 \vert E \vert}. \end{equation} The final average energy is \begin{eqnarray} \langle E_1 \rangle &= \langle K \rangle + \langle V_1 \rangle = \langle K \rangle + \langle V_0 \rangle - \langle V_0 \rangle + \langle V_1 \rangle \Leftrightarrow \nonumber \\ \langle E_1 \rangle &= E_0 + \frac{A_1 - A_0}{A_0} \langle V_0 \rangle, \nonumber \end{eqnarray} where we need \begin{equation} \label{aveVPTII} \langle V_0 \rangle = \frac{1}{T} \int_{0}^{T} \frac{A_0}{\cosh^{2} \left(\lambda q(t)\right)} \; dt. \end{equation} In this case again we can find the explicit solution of $q(t)$ in a closed form, namely \begin{equation} q(t) = \frac{1}{\lambda} \rm{arcsinh} \left( \sqrt{\frac{E+A}{\vert E \vert}} \sin \left( \lambda t \sqrt{2\vert E \vert}\right)\right). \end{equation} The average potential is \begin{eqnarray} \langle V_0 \rangle &=& -\frac{1}{T} \int_{0}^{T} \frac{A_0}{\cosh^{2}\left( \lambda q(t)\right)} \; dt \nonumber \\ &=& -\frac{A_0}{\lambda T \sqrt{2\vert E_0 \vert}} \int_{0}^{2\pi} \frac{dy}{1+\frac{E_0 + A_0}{\vert E_0 \vert}\sinh^{2} y} \Rightarrow \nonumber \\ \langle V_0 \rangle &=& -\sqrt{A_0 \vert E_0 \vert}, \end{eqnarray} The minimum of the potential is $-A_0$, so we introduce $\epsilon$, such that $E_0=-A_0 +\epsilon$. Further, by defining $c=\epsilon/A_0$, we can calculate the ratio of the actions after the kick, as follows \begin{equation} \frac{I\left(\langle E_1 \rangle\right)}{I\left(E_0 \right)}= F(x)=\frac{\sqrt{\vert c-1 +\sqrt{\vert c-1 \vert}(x-1)\vert} -\sqrt{x}}{\sqrt{\vert c-1 \vert}-1}. \end{equation} We have positive average energy after the kick if $x\le x_{esc} = 1-\sqrt{\vert E_0 \vert/A_0}$. For $x\ge x_{esc}$ the function $F(x)$ has a single minimum at $x=1$ and its value is $F(1)=1$, in complete agreement with the PR Conjecture. \subsection{Cosh potential} \label{Coshpotential} The Hamiltonian is \begin{equation}\label{coshhamiltonian} H(q,p) = \frac{p^{2}}{2} + A \cosh \lambda q,\quad A>0. \end{equation} The action is \begin{eqnarray} I(E) &=& -\frac{i\,4\sqrt{2(E-A)}}{\pi \lambda} \nonumber\\ &&\quad\,{\rm E} \left(\frac{i}{2} \log \left( \frac{E+\sqrt{E^2+A^2}}{A} \right), \frac{2A}{A-E} \right), \end{eqnarray} where ${\rm E}$ is the elliptic integral of the second kind, with $k^2=2A/(A-E)$, and the definition \begin{equation} {\rm E}(\phi,k) = \int_0^\phi dt \sqrt{1 - k^2\sin^2 t}. \end{equation} For large energies $E \gg A$ we have the asymptotics $I(E) \approx \sqrt{E}\log E$, and the second derivative of $I^{2}$ is approximately \begin{eqnarray} \frac{\partial ^{2} I^{2}(E)}{\partial E^{2}} \propto \frac{2\log E}{E}, \end{eqnarray} which means again that the PR Conjecture is satisfied. In figure \ref{actioncosh} we plot the function $I^2(E)$ as a function of $E$ for the case $A=0$ and $\lambda=0$. \begin{figure} \center \includegraphics[width=4.5cm]{actioncosh.eps} \caption{Cosh potential: $I^2(E)$ as a function of $E$, for $A=1$ and $\lambda=1$, showing that it is a convex function, satisfying the PR property. } \label{actioncosh} \end{figure} \section{Studies of breaking the PR Conjecture} \label{PRCN} In this section we shall investigate the potentials where the PR Conjecture is violated, either because of the nonanalyticity (nonsmoothness) or due to the closeness to a stationary point (separatrix in the phase space). Since the PR Conjecture in full generality is not true, we will better speak of the PR property rather than PR Conjecture. We shall show the breaking of the PR property in case of $C^0$ potential in the subsection \ref{box}, and then demonstrate in the two next subsections that PR property is restored if the first derivative is continuous, and even more so if the second derivative of the potential $V(q)$ is continuous. After that we shall study what happens if the potential is not a single well potential, but has other stationary points, implying the existence of a separatrix in the phase space. In those cases the PR property can be broken, typically for a certain energy range around the separatrix. \subsection{Harmonic oscillator in a box} \label{box} The first example of a potential (system) which violates the PR Conjecture is a nonanalytic potential, namely the harmonic oscillator in a box. The Hamiltonian is \begin{equation}\label{boxhamiltonian} H(q,p) = \frac{p^{2}}{2} + V(q), \end{equation} where \begin{displaymath} V(q)=\left\{ \begin{array}{lr} \frac{\omega^{2} q^{2}}{2}, & \mbox{} q\in[-q_{0},q_{0}] \\ q = \infty , & \mbox{} |q| \ge q_0 \end{array} \right\}, \end{displaymath} and the kick parameter is $A=-\omega^2$. The potential is continuous, but has discontinuous first derivative at $q=q_0$, so it is $C^0$ only. For the energy $E$ smaller than $E_c = \omega^2q_0^2/2$ the system behaves just as the ordinary linear oscillator treated in subsection \ref{hompot}, equation (\ref{homohamiltonian}), and $d^2I^2/dE^2=1/\omega^2\ge 0$, whilst at $E\ge E_c$ the box potential plays a role. For a pure box potential without the harmonic oscillator we have $I\propto \sqrt{E}$ and consequently $d^2I^2/dE^2=0$. As we will see, for the combined potential (\ref{boxhamiltonian}) at $E\ge E_c$ we find that $d^2I^2/dE^2<0$. The action at $E\ge E_c$ is \begin{eqnarray} \label{boxaction} I\left(E\right) &= \frac{1}{2 \pi} \oint p \; dq = \frac{1}{\pi} \int_{-q_0}^{q_0} \sqrt{2\left(E + \frac{1}{2} \omega^{2} q^{2}\right)} \; dq. \nonumber \\ I\left(E\right)&=\frac{2 E}{\pi \omega}\left({\rm{arcsin}} \left( \frac{q_{0}\omega}{\sqrt{2E}} \right) + \frac{q_{0} \omega}{\sqrt{2E}}\sqrt{1 - \frac{q_{0}^{2} \omega^{2}}{2E}} \;\right). \end{eqnarray} The action in series expansion is \begin{eqnarray} I(E) = \frac{4 E x_{0}}{\pi \omega} \left( 1 - \frac{x_{0}^{2}}{6} + \frac{3 x_{0}^{4}}{20} - \frac{x_{0}^{6}}{112} + h.o.t. \right) \end{eqnarray} where $x_0 = \frac{q_{0}\omega}{\sqrt{2 E}}$. Finally, we have \begin{equation} \frac{\partial ^{2} I^{2}(E)}{\partial E^{2}} = - \frac{16 x_{0}^{6}}{45 \pi^{2} \omega^{2}}\left( 2 + \frac{18}{7}x_{0}^{2} + h.o.t. \right) \leq 0 \end{equation} Thus we see that due to the harmonic potential inside the box $d^2I^2/dE^2$ is no longer zero, but negative. Therefore, the curve $I^2(E)$ is not convex for the energy range $E \ge E_c=\omega^2q_0^2/2$ and implies that $Q$ defined in equation (\ref{Taylor}) and given in the closed form in (\ref{Qfin}), is negative and therefore $F(x)$ has a maximum $F(1)=1$ at $x=1$ instead of a minimum, which is a violation of the PR property at energy range $E\ge E_c$. \subsection{Quadratic-linear potential} \label{qlpot} The previous nonanalytic potential does not possess the PR property, due to nonanalyticity. In fac, $V(q)$ is only a continuous function, the first derivative is discontinuous. Therefore, let us look at an example where the first derivative is smooth, whilst the second is not. Thus we construct a potential which is quadratic (harmonic) up to $|q| \le q_0$, and linear outside the interval $(-q_0,q_0)$, such that the first derivative $dV/dq$ at $|q|=q_0$ is continuous. The Hamiltonian is \begin{equation}\label{qlhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0. \end{equation} where \begin{displaymath} f(q)=\left\{ \begin{array}{lr} -\frac{\omega^{2} q^{2}}{2}, & \mbox{} q\in[-q_{0},q_{0}] \\ - \omega^2 q_0|q| + \frac{\omega^2q_0^2}{2}, & \mbox{} q \not\in [-q_{0},q_{0}]. \end{array} \right. \end{displaymath} In figure \ref{I2Eql} we show the plot of the $I^{2}(E)$, obtained numerically, and it is obvious that it is a convex function for all $E$, meaning that the PR property is satisfied. We might conjecture that the $C^1$ smoothness of a single minimum potential $V(q)=-Af(q)$ is enough for the PR property to hold. \begin{figure} \center \includegraphics[width=4.5cm]{actql.eps} \caption{Quadratic-linear potential. The $I^{2}(E)$ with $A=1$, $\omega = 1$ and $q_0=1$.} \label{I2Eql} \end{figure} \subsection{Quadratic-quartic potential} \label{qqpot} Let us increase the degree of smoothness and consider the $V(q)$ functions of class $C^2$, i.e. having a continuous second derivative only. Of course, we expect the PR property to hold, and this is indeed observed. The Hamiltonian is \begin{equation}\label{qqhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where \begin{displaymath} f(q)=\left\{ \begin{array}{lr} -\frac{\omega^{2} q^{2}}{2}, & \mbox{} q\in[-q_{0},q_{0}] \\ -\frac{\omega^{2}(q + 2 q_{0}\; {\rm sign}(q))^{4}}{108\;q_{0}^{2}} + \frac{\omega^{2} q_{0}^2}{4} , & \mbox{} q \not\in [-q_{0},q_{0}]. \end{array} \right. \end{displaymath} We have calculated $I(E)$ numerically, and show the plot of the function $I^{2}(E)$ in figure \ref{I2Eqq} and it is obvious that it is a convex function of $E$ for all $E$. Thus, the PR property is satisfied. \begin{figure} \center \includegraphics[width=4.5cm]{actqq.eps} \caption{Quadratic-quartic potential: $I^{2}(E)$ with $A=1$, $\omega = 1$ and $q_0=1$.} \label{I2Eqq} \end{figure} \subsection{Sextic potential} \label{sexticpot} The Hamiltonian is \begin{equation} \label{qlhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where the potential $V(q)=-Af(q)$ is given by \begin{equation} f(q) = -(cq + (q-1)^{3})(cq + (q+1)^{3}), \; c\ge 0. \end{equation} This model cannot be analyzed analytically. If $c=0$, there are two stationary points at $q=\pm 1$, namely inflection points, at the energy (potential level) $E=V=0$. By numerical techniques one can convince himself that the $dI^2/dE^2$ can be negative if the energy is close to the value of the stationary points. This is still true if $c$ is nonzero, but small. We choose $c=0.001$. The potential is plotted in figure \ref{sexpotVq}, and it has only one minimum. \begin{figure} \center \includegraphics[width=4.5cm]{potsextic.eps} \caption{Sextic potential, with $A=1$, $c=0.001$, for $-1.5 \leq q \leq 1.5$.} \label{sexpotVq} \end{figure} Now let us calculate directly the action ratio $I_1/I_0 =F(x)$ for the parameters $A_0=1$ as a function of $x=A_1$. We observe three cases in figure \ref{sexpotF}. At high energy $E=1.9584$ there is a minimum of $F(x)$ at $x=1$, which means that the PR property is satisfied. At small energy $E=0.032004$ we have a maximum at $x=1$, meaning the violation of the PR property, whilst at some intermediate energy $E=0.10509$ we have a very flat maximum, very close to an inflection point. \begin{figure} \center \includegraphics[width=3.5cm]{actionratiostr1.eps} \includegraphics[width=3.5cm]{actionratiostr3.eps} \includegraphics[width=3.5cm]{actionratiostr2.eps} \caption{Sextic potential. The action ratio $I_{1}/I_{0}$ with $A_{0}=1$, $c=0.001$ and initial energy $E_0$. $n$ is the number of the points in the initial ensemble. Clockwise: For (a) we find a maximum, (b) inflection point (or very shallow maximum) and in (c) we have a minimum, at $x=A_1/A_0=1$. Only in the last case (c) PR property is satisfied.} \label{sexpotF} \end{figure} \subsection{Quartic double well potential} \label{qdwpot} The Hamiltonian is \begin{equation}\label{qlhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where the quartic double well potential is \begin{equation} V(q) = -Af(q) = A(q-1)^{2}(q+1)^{2}. \end{equation} We plot it in figure \ref{qdwpotfig}. \begin{figure} \center \includegraphics[width=4.5cm]{potquartic.eps} \caption{Quartic double well potential, with $A=1$, for $-2 \leq q \leq 2$.} \label{qdwpotfig} \end{figure} We start above the potential maximum $E_0 > A_0$, and make a kick $A_0\rightarrow A_1$. The critical value $x_c$ of $x=A_1/A_0$ for the energy level $\langle E_1\rangle$ to stay outside the separatrix at $E=A_1$, namely $x\le x_c$, is \begin{equation} x_c = \frac{\pi I_0}{\pi I_0+ A_1T_0-E_0T_0}. \end{equation} We have calculated the action ratio $I_{1}/I_{0}$, assuming $A_0=1$, as a function of $x=A_1$ using only numerical methods, and we find three different cases for the point $x=1$ in figure \ref{IEquarpotF}. For the energy $E=1.0323$ close to the potential local maximum having level $V=1$ at $q=0$ we find a maximum of $F(x)$ at $x=1$, meaning the broken PR property, whilst at high energy $E=5.0226$ we observe a local minimum at $x=1$, meaning that the PR Property is satisfied. At intermediate energy $E=1.2205$ we are close to an inflection point. \begin{figure} \center \includegraphics[width=3.5cm]{actionquartic1.eps} \includegraphics[width=3.5cm]{actionquartic2.eps} \includegraphics[width=3.5cm]{actionquartic3.eps} \caption{Quartic double well potential: The action ratio $I_{1}/I_{0}$ with $A_{0}=1$ and initial energy $E_0$ as a function of $x=A_1$. $n$ is the number of points in the initial ensemble and the bullet denotes the critical point at which we cross the separatrix becoming trapped in one of the two wells. Observe the behaviour at $A_1=1$, clockwise: For (a) we find a minimum, in (b) we have a maximum, and in (c) the inflection point.} \label{IEquarpotF} \end{figure} The examples of this section demonstrate that the PR property can be broken if the potential has a single minimum but is not sufficiently smooth (if it is just $C^0$), whilst the class $C^1$ or higher is sufficient for the PR property to hold. On the other hand, if the potential is analytic but is not a single minimum potential, thus having a separatrix in the phase space, the PR property can be broken for energies in the range near to the separatrix energy level. Moreover, the potential can be analytic, with a single minimum, but has a region where there is almost an inflection point, where the breaking of PR property is observed for the energies in the range close to the almost-inflection point. \section{Discussion and conclusions} \label{Conclusion} In this work we have analyzed the statistical properties of one degree of freedom parametrically kicked Hamiltonian systems, which is the extreme case of fast time dependence, being the opposite extreme to an adiabatic, infinitely slow, changing. As such, the parametric kick behaviour is a very good approximation for the behaviour of the systems under very fast changes in the system parameters, within a time scale of less than one period of oscillation, as has been demonstrated already by Papamikos and Robnik \cite{PapRob2012}. The most natural ensemble, and the most important one is the microcanonical ensemble, because if we have a large ensemble of identical systems with the same ("prepared") energy, and we do not have any further information about them, the uniform distribution with respect to the canonical angle ("the phases") is the most appropriate one. Our main interest is in the value of the final average energy, just after the parametric kick, and the value of the action, identical to the adiabatic invariant, or identical to the area $\Omega$ inside the energy contour (divided by $2\pi$) in the phase space. If the value of the adiabatic invariant at the average final energy increases after the kick, we say that the system has the PR property, following the conjecture put forward by Papamikos and Robnik \cite{PapRob2012}. As discussed in section 2 this implies increasing Gibbs entropy $S_G$ at the mean energy. It turns out that the PR property is satisfied in a vast variety of potentials in which we have proven the validity of the conjecture by direct and rigorous calculations. However, the PR Conjecture is not always satisfied. We have explored exceptions and found that the PR property can be broken if the potential is not sufficiently smooth (if it is $C^0$ only), or if it has several local minima and maxima, implying existence of a separatrix, or of several separatrices, in the phase space. Of course, these complications are not unexpected, because the existence of a separatrix in the phase space always plays and important role, e.g. crossing a separatrix breaks the validity of the adiabatic theorem \cite{Arnold}. In energy ranges close to the separatrix, i.e. stationary points of the potential, the PR property can be broken. Further study of the PR properties in Hamiltonian systems with one and also with more degrees of freedom seems to be very challenging and important. Our results suggest that the following proposition might be true: A strictly convex $C^2$ potential (thus with a single minimum) has the PR property. A proof of this statement is lacking and left for the future. We do believe that this research and the results are important in the statistical mechanics of few body systems, and also for the large, macroscopical, ensembles of identical {\em noninteracting} parametrically driven nonlinear oscillators, as discussed in section 2. Further theoretical research is in progress. Moreover, the research of the general statistical behaviour of the energy distribution (and of other dynamical variables), in the regimes between the ideal adiabatic variation and the parametric kicking, is of great interest and importance. \section*{Acknowledgements} Financial support of the Slovenian Research Agency ARRS under the grant P1-0306 is gratefully acknowledged. We also thank the referees for constructive critical remarks and very useful suggestions. \section*{Appendix A: Summary of the cases with or without the PR property} {\bf Examples of PR property valid for all energies} \begin{enumerate} \item Homogeneous power law kinetic energy and potential \begin{equation}\label{homohamiltonian} H(p,q)=\frac{p^{2n}}{2n} + \frac{Aq^{2m}}{2m},\quad A>0,\; m>0,\; n>0. \end{equation} \item Pendulum (for the average energy inside the separatrix - librations) \begin{equation}\label{pendulumham} H(q,p)= \frac{p^{2}}{2} - \Omega^{2} \cos q. \end{equation} \item Radial Kepler problem with zero angular momentum \begin{equation}\label{keplerpotential} H(r,p)= \frac{p^{2}}{2} - \frac{A}{r},\quad A>0. \end{equation} \item Radial Kepler problem with nonzero angular momentum \begin{equation}\label{keplereffectivepotential} H(r,p)= \frac{p^{2}}{2} - A \left(\frac{a}{r} - \frac{M^{2}}{2r^{2}}\right),\quad A>0,\; a>0,\;M^{2}>0. \end{equation} \item Morse potential \begin{equation}\label{morsepotential} H(q,p)= \frac{p^{2}}{2} + A(e^{-2 \lambda r} - e^{- \lambda r}),\quad A>0,\quad \lambda>0. \end{equation} \item P\"oschl-Teller I potential \begin{equation} \label{pt1hamiltonian} H(q,p) = \frac{p^{2}}{2} + \frac{A}{\cos^{2} \left( \lambda q \right)},\quad \lambda>0,\; A>0. \end{equation} \item P\"oschl-Teller II potential \begin{equation}\label{pt2hamiltonian} H(q,p) = \frac{p^{2}}{2} - \frac{A}{\cosh^{2} \left(\lambda q\right)},\quad \lambda>0,\; A>0. \end{equation} \item Cosh potential \begin{equation}\label{coshhamiltonian} H(q,p) = \frac{p^{2}}{2} + A \cosh \lambda q,\quad A>0. \end{equation} \item Quadratic-linear potential $C^1$ \begin{equation}\label{qlhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where \begin{displaymath} f(q)=\left\{ \begin{array}{lr} -\frac{\omega^{2} q^{2}}{2}, & \mbox{} q\in[-q_{0},q_{0}] \\ - \omega^2 q_0|q| + \frac{\omega^2q_0^2}{2}, & \mbox{} q \not\in [-q_{0},q_{0}]. \end{array} \right. \end{displaymath} \item Quadratic-quartic potential $C^2$ \begin{equation}\label{qqhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where \begin{displaymath} f(q)=\left\{ \begin{array}{lr} -\frac{\omega^{2} q^{2}}{2}, & \mbox{} q\in[-q_{0},q_{0}] \\ -\frac{\omega^{2}(q + 2 q_{0}\; {\rm sign}(q))^{4}}{108\;q_{0}^{2}} + \frac{\omega^{2} q_{0}^2}{4} , & \mbox{} q \not\in [-q_{0},q_{0}]. \end{array} \right. \end{displaymath} \end{enumerate} \noindent {\bf Examples of violated PR property} \begin{enumerate} \item Harmonic oscillator in a box ($C^0$, entirely convex, violation at $E\ge V(q_0)$ ) \begin{equation}\label{boxhamiltonian} H(q,p) = \frac{p^{2}}{2} + V(q), \end{equation} where \begin{displaymath} V(q)=\left\{ \begin{array}{lr} \frac{\omega^{2} q^{2}}{2}, & \mbox{} q\in[-q_{0},q_{0}] \\ q = \infty , & \mbox{} |q| \ge q_0 \end{array} \right. . \end{displaymath} \item Sextic potential (single minimum potential) violation around $E=0$, due to the flatness near $|q|\approx 1$ (close to a stationary point) \begin{equation} \label{qlhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where the potential $V(q)=-Af(q)$ is given by \begin{displaymath} f(q) = -(cq + (q-1)^{3})(cq + (q+1)^{3}), \; c\ge 0. \end{displaymath} \item Quartic double well potential PR property broken near the local maximum $E=A$. \begin{equation}\label{qlhamiltonian} H(q,p) = \frac{p^{2}}{2} - A \; f(q),\quad A>0, \end{equation} where the quartic double well potential is \begin{displaymath} V(q) = -Af(q) = A(q-1)^{2}(q+1)^{2}. \end{displaymath} \end{enumerate}
\section{Introduction} K\"ahler-Ricci flow is the natural version of Ricci flow, introduced by Hamilton in \cite{ham82}, in the context of complex manifold. The study of it was led to the case of Fano manifold, i.e. closed complex manfold with positive first Chern class, after the work \cite{cao} by H. D. Cao. Major works have been carried out in this direction, for example, \cite{chen-tian, tian-zhu}, and some most recent ones including \cite{tian-zzl, jiang} appearing after the breakthrough regarding the sufficient condition on the existence of K\"ahler-Einstein metric by G. Tian and and also Chen-Sun-Donaldson. On the other hand, in recent years, K\"ahler-Ricci flow in the most general setting has also attracted extensive attentions, especially with the program proposed by Tian in \cite{tian02} as the geometric analysis version of the renowned Minimal Model Program in algebraic geometry. A lot of progress has been made on this topic, for example, in \cite{t-znote, song-tian}. In this work, we consider the strong convergence for these two main versions of K\"ahler-Ricci flows. In the Fano case, we orginally made the assumptions on the volume form and the Ricci curvature form, while the one on Ricci form can be removed by properly applying the result in \cite{ses-tian}. For the flow in the most general setting, the assumptions are on the global volume and the Ricci curvature form. Let's state the main results below. \begin{theorem} \label{thm:fano-1} Consider the K\"ahler-Ricci flow over a Fano manifold, (\ref{eq:krf}). If uniformly for all time, the volume form is bounded away from $0$, then the K\"ahler-Einstein metric exists over this manifold and the flow metric converges to it at time infinity. \end{theorem} \begin{theorem} \label{thm:general-conv} Consider the K\"ahler-Ricci flow, (\ref{eq:gkrf}). If uniformly in the maximal time interval of existence, Ricci curvature form is bounded from below by a fixed real smooth $(1, 1)$-form and the global volume is bounded away from $0$, then the flow exists forever. Furthermore, the manifold has negative first Chern class and the flow metric converges to the K\"ahler-Einstein metric at time infinity. \end{theorem} \noindent{\bf Notation:} in the following, $C$ stands for a positive constant which might be different at places. We also use other letters to indicate constants for better illustration of the idea. \begin{acknowledgment} The author would like to thank Professor G. Tian for introducing him to this interesting field of research and constant encouragement. We also appreciate the interest by people including Kwok-Kun Kwong, Yalong Shi and Ben Weinkove. The discussion with Yalong Shi stimulated the improvement on the result for the Fano case. \end{acknowledgment} \section{Fano K\"ahler-Ricci Flow} The traditional K\"ahler-Ricci flow over a Fano manifold is one case of the more general K\"ahler-Ricci flow. Of course, it is fairly justified to focus on this specific flow because of its natural relation to the search of the necessary and sufficient condition of the existence of K\"ahler-Einstein metrics over Fano manifolds. Over a Fano manifold $X$ of complex dimension $n\geqslant 2$, i.e. the first Chern class $c_1(X)$ being K\"ahler, we consider the following geometric evolution: \begin{equation} \label{eq:krf} \frac{\p \widetilde\omega_t}{\p t}=-\Ric(\widetilde\omega_t)+\widetilde\omega_t, ~~~~\widetilde\omega_0=\omega \end{equation} with the initial metric $\omega$ representing $c_1(X)$, i.e. $[\omega]=[\Ric(\omega)]=c_1(X)\in H^2(X;\mathbb{R})\cap H^{1, 1}(X;\mathbb{C})$. This flow is known to exist forever and $[\widetilde\omega_t]=c_1(X)$ for all time, and so we set $\widetilde\omega_t=\omega+\sqrt{-1}\p\bar\p u$. There is a unique smooth volume for $\Omega$ such that $\omega=\Ric(\Omega)$ and $\int_X\Omega=[\omega]^n$. Now (\ref{eq:krf}) can be transformed to the following equivalent scalar evolution equation for the metric potential $u$: \begin{equation} \label{eq:skrf} \frac{\p u}{\p t}=\log\frac{(\omega+\sqrt{-1}\p\bar\p u)^n}{\Omega}+u, ~~~~u|_{t=0}=0. \end{equation} Take $t$-derivative to get $$\frac{\p}{\p t}\(\frac{\p u}{\p t}\)=\Delta\(\frac{\p u}{\p t}\)+\frac{\p u}{\p t},$$ which can be transformed to $$\frac{\p}{\p t}\(\frac{\p u}{\p t}-u\)=\Delta\(\frac{\p u}{\p t}-u\)+n-\<\widetilde\omega_t, \omega\>,$$ where $\<\widetilde\omega_t, \omega\>$ means taking trace of $\omega$ with respect to $\widetilde\omega_t$. Meanwhile, (\ref{eq:skrf}) can be reformulated in the following way which is more of a complex Monge-Ampere look: \begin{equation} \label{eq:cma} (\omega+\sqrt{-1}\p\bar\p u)^n=e^{\frac{\p u}{\p t}-u}\Omega. \end{equation} In \cite{ses-tian}, together with other higher order controls, it is shown that $$\osc_{X\times \{t\}} \frac{\p u}{\p t}\leqslant C$$ uniformly for all time, which is an essential building block in the recent study of this flow. \subsection{Volume Form Bounded from Above} If we assume for some constant $p>1$, $$\int_X e^{p\(\frac{\p u}{\p t}-u\)}\Omega\leqslant C$$ uniformly for all time, applying the original $L^\infty$-estimate by Kolodziej in \cite{koj98} (or \cite{kojnotes}) to (\ref{eq:cma}), we know that uniformly for all time, $$\osc_{X\times \{t\}}u\leqslant C.$$ \begin{remark} The assumption above can certainly be weakened using Kolodziej's original result, and we leave it like this for the simplification of terminology. Of course, it would be the case if $\frac{\p u}{\p t}-u\leqslant C$, i.e. the volume form has a uniform upper bound. \end{remark} Then as $\int_X e^{\frac{\p u}{\p t}-u}\Omega=[\widetilde\omega_t]^n=c_1(X)^n$, we know $$\vline\frac{\p u}{\p t}-u\vline\leqslant C.$$ Now recall the following inequality for Parabolic Schwarz Lemma: $$\(\frac{\p}{\p t}-\Delta\)\log\<\widetilde\omega_t, \omega\>\leqslant C\<\widetilde\omega_t, \omega\>+C.$$ Combining it with $$\(\frac{\p}{\p t}-\Delta\)\(\frac{\p u}{\p t}-u\)=n-\<\widetilde\omega_t, \omega\>,$$ we arrive at $$\(\frac{\p}{\p t}-\Delta\)\(\log\<\widetilde\omega_t, \omega\>+E\(\frac{\p u}{\p t}-u\)\)\leqslant -\<\widetilde\omega_t, \omega\>+C$$ for some large enough positive constant $E$. Applying Maximum Principle in the standard way, together with the uniform bound of $\frac{\p u}{\p t}-u$, we conclude $$\<\widetilde\omega_t, \omega\>\leqslant C.$$ Hence we could have the following uniform bound of $\widetilde\omega_t$ as metric for all time: $$\frac{1}{C}\omega\leqslant \widetilde\omega_t\leqslant C\omega.$$ It is now classic to obtain uniform $C^k$ bounds for the normalized $u$, and one can then apply the K-energy argument to achieve a sequential convergence with the limit being the K\"ahler-Einstein metric. By the result in \cite{tian-zhu}, we actually have the flow convergence as $t\to\infty$. We summarize all this in the following proposition. \begin{prop} \label{prop:v-u-b} If for some $p>1$, $\int_X e^{p\(\frac{\p u}{\p t}-u\)}\Omega\leqslant C$ uniformly for all time, then the K\"ahler-Eintein metric exists over $X$ and the flow metric of (\ref{eq:krf}) converges to it as $t\to\infty$. \end{prop} This result is by no means new and has appeared essentially in the works of Tian and Zhu, for example. We merely use it to collect some known techniques in this business. \subsection{Volume Form Bounded from Below} Now we move on to the case when the volume form has a uniform positive lower bound, i.e. for all time, $$\frac{\p u}{\p t}-u\geqslant -C.$$ In this case, we also assume that for all time, $$\Ric(\widetilde\omega_t)\geqslant \alpha$$ for a fixed real smooth $(1, 1)$-form $\alpha$ over $X$. By (\ref{eq:krf}), we then have $$\alpha\leqslant \Ric(\widetilde\omega_t)=-\frac{\p \widetilde\omega_t}{\p t}+\widetilde\omega_t=\omega+\sqrt{-1}\p\bar\p\(-\frac{\p u}{\p t}+u\),$$ and so for some large enough $F$, $$F\omega+\sqrt{-1}\p\bar\p\(-\frac{\p u}{\p t}+u\)>0.$$ Now we need the solution of Berndtsson in the very recent work \cite{bo} on the Openness Conjecture proposed by Demailly and Kollar. The result is as follows. \begin{theorem} \label{thm:openness} Let $u$ be a plurisubharmonic function in the unit ball $B\subset\mathbb{C}^n$ and $u\leqslant D$. Using the standard Euclidean measure $d\lambda$, assume that $$\int_Be^{-u}d\lambda<A<\infty.$$ Then we have for the half unit ball $B/2$ and $\epsilon=\frac{\delta(n)}{e^DA}$ with some universial dimensional constant $\delta(n)$, $$\int_{B/2}e^{-(1+\epsilon)u}d\lambda<C(n)\cdot\(e^{-\epsilon D}A+e^{-(1+\epsilon)D}\),$$ where $C(n)$ is a positive constant only depending on $n$. \end{theorem} The above statement is more quantitative comparing with the orginal statement in \cite{bo} and certainly a direct consequence of the argument there. In application, we look at the function $-\frac{\p u}{\p t}+u\in PSH_{F\omega}(X)$ and $-\frac{\p u}{\p t}+u\leqslant C$. The integral bound is clear from global cohomology information: $$\int_X e^{-(-\frac{\p u}{\p t}+u)}\Omega=c_1(X)^n<\infty.$$ So the above result by Berndtsson tells us for some $\epsilon>0$ $$\int_X e^{(1+\epsilon)(\frac{\p u}{\p t}-u)}\Omega=\int_X e^{-(1+\epsilon)(-\frac{\p u}{\p t}+u)}\Omega\leqslant C$$ uniformly for all time. Again, by Kolodziej's $L^\infty$-estimate, we know from (\ref{eq:cma}): $$\osc_{X\times\{t\}}u\leqslant C$$ uniformly for all time. Now the argument in Subsection 2.1 can be recycled to get the existence of K\"ahler-Einstein metric and the convergence of flow metric to it. Let's summarize all this in the following theorem. \begin{theorem} \label{th:v-ric-conv} If for all time, $\frac{\p u}{\p t}-u\geqslant -C$ and $\Ric(\widetilde\omega_t)\geqslant\alpha$ for a fixed real smooth $(1, 1)$-form $\alpha$ over $X$, then the K\"ahler-Einstein metric exists over $X$ and the flow metric of (\ref{eq:krf}) converges to it as $t\to\infty$. \end{theorem} The assumption on Ricci form can be removed by observing that the result in \cite{ses-tian} quoted before allows us to replace $\frac{\p u}{\p t}$ by a constant (the average for example) up to a control term, which can be absorbed by the volume form. Then the plurisubharmonicity of the metric potential $u$ is enough for the application of Theorem \ref{thm:openness}, together with the positive lower volume form bound. So we have proved Theorem \ref{thm:fano-1}. \begin{remark} As pointed out by Yalong Shi, such implication of the flow convergence and existence of K\"ahler-Einstein metric by the one-sided volume form bound can be viewed as a reminisence of the classic result of Tian and Siu on the more direct continuity method setting to solve the K\"ahler-Einstein equation on a Fano manifold. \end{remark} \section{General K\"ahler-Ricci Flow} In this section, we turn to the study of the following K\"ahler-Ricci flow over a closed K\"ahler manifold $X$ of complex dimension $n\geqslant 2$: \begin{equation} \label{eq:gkrf} \frac{\p\widetilde\omega_t}{\p t}=-\Ric(\widetilde\omega_t)-\widetilde\omega_t, ~~~~\widetilde\omega_0=\omega_0, \end{equation} where the initial metric $\omega_0$ is any K\"ahker metric over $X$. The K\"ahler class of the flow metric is changing in general along the flow, and we prefer this version because the cohomology is automatically controlled under it. To study this evolution equation, we set $$\omega_t=-\Ric(\omega_0)+e^{-t}\(\omega_0+\Ric(\omega_0)\)$$ and $\widetilde\omega_t=\omega_t+\sqrt{-1}\p\bar\p u$. The following scalar evolution equation of the metric potential is equivalent to (\ref{eq:gkrf}) \begin{equation} \label{eq:sgkrf} \frac{\p u}{\p t}=\log\frac{(\omega_t+\sqrt{-1}\p\bar\p u)^n}{\omega^n_0}-u, ~~~~u|_{t=0}=0. \end{equation} In \cite{t-znote}, the {\it Optimal Existence Result} is obtained, i.e. the flow exists as long as the class $[\omega_t]$ remains to be K\"ahler. We denote $[0, T)$ as the maximal time interval of existence, where $T\in (0, \infty]$, depending on the cohomology information. The main result of this section is the following. \begin{theorem} \label{th:gkrf-conv} If uniformly for all $t\in [0, T)$, $\Ric(\widetilde\omega_t)\geqslant\alpha$ for a fixed real smooth $(1, 1)$-form $\alpha$ and $[\omega_t]^n>\epsilon>0$ for some $\epsilon$, then $T=\infty$ and actually $c_1(X)<0$ (implying $X$ being projective) and the flow converges to the K\"ahler-Einstein metric as $t\to\infty$. \end{theorem} This is just Theorem \ref{thm:general-conv}. Let's clarify that the above $\alpha$ and $\epsilon$ are independent of the singularity time $T$. \begin{proof} In light of $\Ric(\widetilde\omega_t)\geqslant\alpha$ and applying a simple ODE calculation, we see $\widetilde\omega_t\leqslant C\omega_0$ uniformly for $t\in [0, T)$. It is also clear that $$\alpha\leqslant \Ric(\widetilde\omega_t)=-\frac{\p \widetilde\omega_t}{\p t}-\widetilde\omega_t=\Ric(\omega_0)+\sqrt{-1}\p\bar\p\(-\frac{\p u}{\p t}-u\),$$ and so for some large enough $F$, $$F\omega_0+\sqrt{-1}\p\bar\p\(-\frac{\p u}{\p t}-u\)>0.$$ Applying a result in \cite{tian-87} which is the manifold version of a classic result by H\"ormander in \cite{hor}, we have for any $t\in [0, T)$, a constant $a>0$ and $C$ such that $$\int_X \exp\(a\(\sup_{X\times \{t\}}\(-\frac{\p u}{\p t}-u\)+\frac{\p u}{\p t}+u\)\)\omega^n_0\leqslant C,$$ where obviously we could make sure $a\in (0, 1]$. Now we have \begin{equation} \begin{split} \exp\(a\cdot\inf_{X\times \{t\}}\(\frac{\p u}{\p t}+u\)\) &\geqslant C\int_X \exp\(a\(\frac{\p u}{\p t}+u\)\)\omega^n_0 \\ &\geqslant C\int_X \exp\(a\(\frac{\p u}{\p t}+u-C\)\)\omega^n_0 \\ &\geqslant C\int_X \exp\(\frac{\p u}{\p t}+u-C\)\omega^n_0 \\ &= C[\widetilde\omega_t]^n=C[\omega_t]^n>\epsilon>0, \end{split} \nonumber \end{equation} where for the third $\geqslant$, we make use of $\frac{\p u}{\p t}+u\leqslant C$ which are known by standard Maximum Principle argument (as in \cite{r-blow-up} for example). Hence we conclude $$\frac{\p u}{\p t}+u\geqslant -C,$$ which gives the uniform volume form lower bound. Together with $\widetilde\omega_t\leqslant C\omega_0$ pointed out at the beginning, we conclude that uniformly for all $t\in [0, T)$, $$\frac{1}{C}\omega_0\leqslant \widetilde\omega_t\leqslant C\omega_0,$$ where $C$ does not depend on $T$. As in \cite{r-blow-up}, we can then see that there is no finite time singularity and $[\omega_t]$ is K\"ahler also for $t=\infty$, i.e. $-c_1(X)$ is K\"ahler. In this case, the convergence of flow metric to the K\"ahler-Einstein metric is known even when the initial K\"ahler class is not $-c_1(X)$, for example, in \cite{thesis}. The proof is finished. \end{proof} \begin{remark} The above argument is similar to and also extends that in \cite{ricci-lower}, since for finite time singularity case $\Ric(\widetilde\omega_t)\geqslant -C\widetilde\omega_t$ implies $\Ric((\widetilde\omega_t)\geqslant \alpha$ for some fixed $\alpha$, but not the other way, at least a priori. \end{remark} \section{Final Remarks} In spite of the essential difference in the existing methods and techniques for the study of these two versions of K\"ahler-Ricci flows, there is similarity illustrated by these results and arguments. It would be interesting to create further combination of the ideas, especially through making use of the recent progress in pluripotential theory.
\section{Main Body} {\bfseries The advent of quantum optical techniques based on superconducting circuits\cite{Wallraff:2004} has opened new regimes in the study of the non-linear interaction of light with matter. Of particular interest has been the creation of non-classical states of light,\cite{Hofheinz:2009,Sayrin:2011,Kirchmair:2013,Hoi:2012,Brooks:2012} which are essential for continuous-variable quantum information processing,\cite{Braunstein:2005} and could enable quantum-enhanced measurement sensitivity.\cite{Giovannetti:2004} Here we demonstrate a device consisting of a superconducting artificial atom, the Cooper pair transistor,\cite{Joyez:1994} embedded in a superconducting microwave cavity that may offer a path toward simple, continual production of non-classical photons.\cite{Armour:2013} By applying a dc voltage to the atom, we use the ac Josephson effect to inject photons into the cavity. The backaction of the photons on single-Cooper-pair tunneling events results in a new regime of simultaneous quantum coherent transport of Cooper pairs and microwave photons. This single-pair Josephson laser offers great potential for the production of amplitude-squeezed photon states and a rich environment for the study of the quantum dynamics of nonlinear systems.\cite{Habib:1998,Chaudhury:2009,Blencowe:2012} } \begin{figure}[h] \includegraphics[width=9cm]{Fig1.pdf} \caption{\label{fig1} {\bfseries Single-pair Josephon laser. a,} Schematic illustration of the circuit. {\bfseries b,} Electron micrograph of the CPT location. Proximitized \amount{30}{nm} thick Au/Ti contact pads are used to allow electrical contact between the CPT and the cavity without introducing additional dissipation. The approximately \amount{21.5}{mm}-long cavity is coupled at either end via small capacitors to a waveguide with characteristic impedance $\ensuremath{Z_{0}}=\amount{50}{\Omega}$. {\bfseries c,} Micrograph of the CPT and its gate line. The CPT consists of a \amount{7}{nm} thick superconducting Al island with charging energy $\ensuremath{E_{c}}=e^2/2\ensuremath{C_{\Sigma}}= h\times \amount{8.7}{GHz}$, where \ensuremath{C_{\Sigma}}\ is the total island capacitance and coupled by small Josephson junctions to \amount{70}{nm} thick Al leads. The Josephson coupling energy $\ensuremath{E_{J}} = h\times\amount{17}{GHz}$, which is determined by the junction resistance, is comparable to \ensuremath{E_{c}}. {\bfseries d,} Current \ensuremath{I_{\text{CPT}}}\ through the CPT versus \ensuremath{V_{\text{dc}}}\ and \ensuremath{n_{g}}. {\bfseries e,} Sequential tunneling across the island-drain junction and co-tunneling across the CPT, both with simultaneous net photon emission. For sequential tunneling the energy released by tunneling across the island-drain junction must match the energy for addition of a photon to the cavity: $2eV_{\text{ID}}=\hbar\ensuremath{\omega_{0}}$. For cotunneling the source-drain voltage \ensuremath{V_{\text{SD}}}\ must satisfy a similar condition. Note that emission of multiple photons is also possible for other bias conditions. } \end{figure} The consequences of the ac Josephson effect for electrical transport in Josephson junction systems coupled to microwave fields or electrical resonators have been studied for decades, with the seminal discoveries of such classical phenomena as Shapiro steps\cite{Shapiro:1963} and Fiske modes\cite{Fiske:1964} occurring within only a few years of Josephson's predictions themselves.\cite{Josephson:1962} This physics was later extended to small junction systems for which electrical transport involves energy exchange between tunneling Cooper pairs and their electromagnetic environment.\cite{Hofheinz:2011} In the above work, the electromagnetic fields generated by the junctions could generally be considered classical and in the case of small junction systems did not act back on the electrical transport. One exception relates to recent use of a superconducting charge qubit coupled to a resonator to create a single artificial atom maser.\cite{Rodrigues:2007,Astafiev:2007} Even here, however, generation of quasiparticles during transport limited the quantum coherence of the overall system. In contrast, the cavtiy-embedded Cooper pair transistor (cCPT) involves only the interaction of Cooper pairs and photons, in a highly coherent and heretofore unstudied pumping process that does not depend on dissipative charge relaxation to establish transport. In our system the CPT is located at the voltage antinode at the center of a wavelength-long coplanar waveguide cavity with a resonant frequency $\ensuremath{\omega_{0}}= 2 \pi \times\amount{5.256}{GHz}$ and quality factor $Q=3.5\e{3}$ giving a photon decay rate $\kappa=2 \pi \times\amount{1.5}{MHz}$. The cavity, which is fabricated from \amount{100}{nm} thick Nb film, is modified by placement of dc bias lines at the voltage nodes located one quarter wavelength from either end of the cavity, as shown in Fig.~\ref{fig1}a. These bias lines allow application of a dc bias voltage \ensuremath{V_{\text{dc}}}\ to the central conductor of the cavity through a biasing impedance \ensuremath{Z_{b}}\ without affecting the microwave properties of the cavity at its resonant frequency.\cite{Chen:2011} The CPT itself is fabricated with its source coupled to the central cavity conductor and its drain coupled to the cavity ground plane. A separate gate voltage \ensuremath{V_{g}}\ is applied to the CPT island through a capacitance \ensuremath{C_{g}}\ to adjust its electrostatic potential via the gate charge $\ensuremath{n_{g}} = \ensuremath{C_{g}}\ensuremath{V_{g}}/e$. Escaping photons can be measured by microwave circuitry connected to the cavity's output port while the dc bias lines are used to probe electrical transport in the CPT (see Supplementary Information). For our purposes, the CPT is well described by considering only two charge states, \ensuremath{|0\rangle}\ and \ensuremath{|1\rangle}, corresponding to zero and one excess Cooper pairs on the island. These charge states are separated by a gate-dependent electrostatic energy difference $2\varepsilon=4\ensuremath{E_{c}}(1-\ensuremath{n_{g}})$, and coupled to each other via the Josephson energy \ensuremath{E_{J}}. By way of the ac Josephson effect, the dc bias gives rise to a characteristic drive frequency $\ensuremath{\omega_{d}}= e\ensuremath{V_{\text{SD}}}/\hbar$, which can be viewed approximately as the frequency of Josephson oscillations across each junction. Here \ensuremath{V_{\text{SD}}}\ is the source-drain voltage that exists at the CPT\@. Note that due to the nonlinearity of the system and the presence of the bias impedance $Z_{b}$, \ensuremath{V_{\text{SD}}}\ in general can differ from the applied voltage \ensuremath{V_{\text{dc}}}. Introducing cavity photon annihilation and creation operators $a$ and $a^{\dagger}$, the Hamiltonian of the cCPT can be expressed as \begin{equation}\label{heq} H = \hbar\ensuremath{\omega_{0}} a^{\dagger}a + \varepsilon \sigma_{z} - \ensuremath{E_{J}} \sigma_{x} \cos[\Delta (a + a^{\dagger}) + \ensuremath{\omega_{d}} t], \end{equation} where $\sigma_{x}$ and $\sigma_{z}$ are the Pauli matrices. The first two terms in (\ref{heq}) describe the cavity photons and the CPT charge. The third term describes the coupling between the CPT charge states and the cavity photons, and the effects of the voltage drive. In a standard CPT (with no cavity and at zero bias), this term would read $\ensuremath{E_{J}} \sigma_{x}\cos\varphi/2$ where $\varphi$, the total superconducting phase difference between the source and drain, can be treated as a classical variable.\cite{Joyez:1994} In our case, however, quantum fluctuations of the cavity photon field must be accounted for via the identification $\hat{\varphi} /2=\Delta (a+a^{\dagger})$, which is proportional to the electric field in the cavity at the location of the CPT. The dimensionless parameter $\Delta=\sqrt{\ensuremath{Z_{0}}/\ensuremath{R_{K}}}\approx 0.04$, where $\ensuremath{R_{K}}=h/e^{2}=\amount{25.8}{k\Omega}$ is the resistance quantum, describes the strength of the quantum phase fluctuations of the cavity field, which can be important for large numbers of photons in the cavity. (See Ref.~\citenum{Blencowe:2012} and the Supplementary Information. A different possible physical realization of a similar Hamiltonian was also recently studied.\cite{Marthaler:2011}) The current \ensuremath{I_{\text{CPT}}}\ through the CPT measured versus \ensuremath{V_{\text{dc}}}\ and \ensuremath{n_{g}}\ as in Fig.~\ref{fig1}d shows rich behavior, far more so than in similar measurements of Cooper pair or single electron transistors coupled to lower $Q$ resonators.\cite{Lu:2002,Pashkin:2011} The current is $2e$ periodic in gate charge for $\ensuremath{V_{\text{dc}}}\lesssim\amount{150}{\mu V}$, indicating that only Cooper pair transport is significant for sufficiently low bias. In general, two distinct varieties of transport process attributable to the interaction of the CPT with the cavity are observed. First, there are sequential tunneling processes involving photon emission, as indicated by the diagonal line in Fig.~\ref{fig1}d. As shown in the left panel of Fig.~\ref{fig1}e, such processes involve an allowed transition of a Cooper pair across either the source or drain junction, combined with net emission of a photon into the cavity. Second, there are also higher order virtual processes (cotunneling), indicated by the vertical line in Fig.~\ref{fig1}d. Here, as shown schematically in the right panel of Fig.~\ref{fig1}e, a Cooper pair is transferred from source to drain through an energetically forbidden state, again with net emission of a photon into the cavity. It is important to note that in the presence of large numbers of cavity photons, both sequential and co-tunneling can be strongly affected by the cavity field, \textit{i.~e.}\ stimulated emission of photons is important, as indicated schematically in Fig.~\ref{fig1}a. \begin{figure*}[t!] \includegraphics[width=17cm]{Fig2New.pdf} \caption{\label{fig2} {\bfseries Transport and photon emission in the single-pair Josephson laser a,} Current versus \ensuremath{V_{\text{dc}}}\ and \ensuremath{n_{g}}\ showing the portions of the parameter space chosen for detailed investigation. {\bfseries b,} $I$-$V$\ characteristic for two sweep directions, as indicated, along the horizontal magenta line in {\bfseries a}. The vertical dashed red and blue lines indicate the locations of the first and second cotunneling features for $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}/2$ ($\ensuremath{V_{\text{SD}}} =\amount{11}{\mu V}$) and $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}$ ($\ensuremath{V_{\text{SD}}} =\amount{22}{\mu V}$) respectively. {\bfseries c} and {\bfseries d,} Microwave spectral power density $S(\omega)$ of the cCPT over a \amount{5}{MHz} span versus detuning $\Delta f = f-\ensuremath{f_{0}}$ from the cavity resonant frequency $\ensuremath{f_{0}} = \ensuremath{\omega_{0}}/2\pi = \amount{5.256}{GHz}$ for the first ({\bfseries c}) and second ({\bfseries d}) cotunneling peaks along the red and blue vertical lines in {\bfseries a} } \end{figure*} To investigate the coupling to cavity photons more carefully, we restrict ourselves to very low applied voltages ($\ensuremath{V_{\text{dc}}}<\amount{30}{\mu V}$) and concentrate on the first two cotunneling features in the current as indicated in Fig.~\ref{fig2}a, which occur for $\ensuremath{V_{\text{SD}}}\approx 11$ and \amount{22}{\mu V}, and correspond to drive frequencies $\ensuremath{\omega_{d}} \approx \ensuremath{\omega_{0}}/2$ and \ensuremath{\omega_{0}}\ respectively. Referring to the diagram for cotunneling in Fig.~\ref{fig1}e, these features correspond to the net emission of one or two photons into the cavity during Cooper pair tunneling. Detailed behavior of the CPT current $I$ versus the measured source-drain voltage \ensuremath{V_{\text{SD}}}\ is shown in Fig.~\ref{fig2}b at a particular gate charge far from the charge degeneracy points at $\ensuremath{n_{g}}=\pm 1$. The current is hysteretic in \ensuremath{V_{\text{SD}}}, indicating the presence of bistability in the cCPT dynamics. Furthermore, there are sharp current steps at fixed voltages \ensuremath{V_{\text{SD}}}\ corresponding to the cotunneling features. Such current steps are well known from the phenomena of Shapiro steps\cite{Shapiro:1963} and Fiske modes\cite{Fiske:1964} to be an indication of frequency matching between Josephson oscillations and an electromagnetic field. In our context, the steps show that the very photons generated by the tunneling Cooper pairs in turn directly influence other tunneling events and subsequent photon emission, \textit{i.~e.}, they constitute direct evidence for stimulated photon emission. To conclusively demonstrate the connection between the current steps in Fig.~\ref{fig2}b and photon emission, we measure the microwave spectral power density $S(\omega)$ emitted by the cCPT for a series of different values of \ensuremath{n_{g}}\ for bias voltages surrounding the cotunneling features near $\ensuremath{V_{\text{SD}}} \approx\amount{11}{\mu V}$ and \amount{22}{\mu V}. In order to minimize noise in the dc bias voltage \ensuremath{V_{\text{dc}}}\ and accompanying jitter in the drive frequency \ensuremath{\omega_{d}}, for the emission data shown here we used low-noise bias circuitry that significantly improved emission stability (see Supplementary Information). For both values of \ensuremath{V_{\text{SD}}}\ there is strong emission close to the cavity resonance frequency \ensuremath{\omega_{0}}\ as can be seen in Fig.~\ref{fig2}c,d. For the first cotunneling feature, at which the drive frequency $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}/2$, emission at twice the drive frequency is a direct consequence of the strong nonlinearity of the system. The emission pattern is $2e$-periodic, as is the electrical transport, and shows interesting structure versus \ensuremath{n_{g}}. In particular the emission for $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}/2$ appears to develop internal structure for $\ensuremath{n_{g}}\gtrsim 0.6$ while for $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}$ there is a clear ``hot spot'' in the emission for $\ensuremath{n_{g}} \approx 0.7$. In both cases, the emission dies out as the gate charge approaches the charge degeneracy points at $\ensuremath{n_{g}}=\pm 1$. A detailed view of the emission at $\ensuremath{\omega_{d}}=\ensuremath{\omega_{0}}/2$ and $\ensuremath{\omega_{d}}=\ensuremath{\omega_{0}}$ reveals additional interesting features, as shown in Figs.~\ref{fig3} and \ref{fig4} which show the emission spectra $S(\omega)$ and cavity photon occupation $\ensuremath{n_{\text{ph}}}$ versus applied voltage \ensuremath{V_{\text{dc}}}\ at representative values of gate charge \ensuremath{n_{g}}\ for both the $\ensuremath{\omega_{d}} =\ensuremath{\omega_{0}}/2$ and $\ensuremath{\omega_{d}}=\ensuremath{\omega_{0}}$ cotunneling features. For the $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}/2$ resonance in Fig.~\ref{fig3} we see that as might be expected for a single-atom emitter there is no clear sign of a lasing threshold, with the cavity occupation \ensuremath{n_{\text{ph}}}\ climbing smoothly from zero as \ensuremath{V_{\text{dc}}}\ is increased. For low $\ensuremath{V_{\text{dc}}}\lesssim \amount{13}{\mu V}$ the emission linewidth of roughly \amount{1}{MHz} shows modest narrowing over the intrinsic cavity linewidth $\kappa = \amount{1.5}{M Hz}$. At $\ensuremath{V_{\text{dc}}}\approx\amount{13}{\mu V}$, there is a sharp change in the emission pattern: the linewidth suddenly drops by roughly an order of magnitude, to as low as \amount{70}{kHz}, slightly larger than the residual jitter of about \amount{25}{kHz} in the drive frequency $\ensuremath{\omega_{d}}/2\pi$. Over the same range in \ensuremath{V_{\text{dc}}}\ the cavity photon occupancy \ensuremath{n_{\text{ph}}}\ reaches a maximum value on the order of 100 before dropping sharply at $\ensuremath{V_{\text{dc}}}\approx\amount{13}{\mu V}$, stabilizing briefly, and then rapidly declining. Strikingly, as charge degeneracy is approached, the sharpened spectrum splits into two narrowly separated peaks at around $\ensuremath{n_{g}}=0.62$. The separation of these peaks increases as \ensuremath{n_{g}}\ approaches charge degeneracy, accompanied by a notable shift in the emission frequency toward negative detuning. There is an even more notable shift toward negative detuning as \ensuremath{n_{\text{ph}}}\ increases for fixed \ensuremath{n_{g}}. This latter tendency results in the characteristic ``v'' shape of the emission spectra versus \ensuremath{V_{\text{dc}}}\ as \ensuremath{n_{\text{ph}}}\ first rises and then falls with increasing bias. The $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}$ emission spectra shown in Fig.~\ref{fig4} share some features with those for $\ensuremath{\omega_{d}} =\ensuremath{\omega_{0}}/2$; there are, however, significant differences as well. There is again no clear sign of a threshold as \ensuremath{V_{\text{dc}}}\ is increased, and there is again a clear though less pronounced pulling toward negative detuning as charge degeneracy is approached or \ensuremath{n_{\text{ph}}}\ is increased. There is no sudden sharpening of the spectrum in this case; instead, the emission simply cuts off abruptly for $\ensuremath{V_{\text{dc}}}\gtrsim\amount{23}{\mu V}$, just after the cavity reaches its maximum occupancy of roughly $\ensuremath{n_{\text{ph}}}\approx 200$. The minimum linewidth of the emission spectra is roughly \amount{500}{kHz}, significantly below the bare cavity linewidth, but not by as much as for the $\ensuremath{\omega_{d}}=\ensuremath{\omega_{0}}/2$ resonance. All in all, the combination of linewidth reduction, high cavity photon number, and Shapiro-like $I$-$V$\ characteristics provide clear evidence of lasing driven by single-Cooper-pair tunneling. For many years, the standard theoretical approach for describing such tunneling has been so-called $P(E)$ theory,\cite{Devoret:1990,Girvin:1990} which considers the process of emission or absorption of photons by tunneling Cooper pairs to or from an environment described by a frequency-dependent impedance \ensuremath{Z_{t}(\omega)}. As recent experiments\cite{Hofheinz:2011} on an effective single junction system have shown, if the environmental impedance \ensuremath{Z_{t}(\omega)}\ is sharply peaked at a frequency $\ensuremath{\omega_{0}}$ (as for a resonance) there is substantial probability $P(E)$ of emission of a photon into the environment when the junction is biased at a voltage given by $V=\hbar\ensuremath{\omega_{0}}/2e$. The result is a peak in the current due to incoherent tunneling of Cooper pairs, combined with emission of microwave photons, similar in some respects to the results presented here.\cite{Hofheinz:2011,Gramich:2013} It is important to note, however, that the above picture fails to accurately describe the Josephson laser. The key point is the assumption essential to $P(E)$ theory that the environment is in thermal equilibrium. At low temperature $T = \amount{30}{mK}$ and high frequency $\ensuremath{\omega_{0}} = 2 \pi\times \amount{5.26}{GHz}$ the thermal occupation of the environmental modes is very small ($\ensuremath{n_{\text{th}}}=2\e{-4}$), and the probability $P(E)$ of absorption of a photon by the CPT is essentially zero. In our experiment, however, the occupation the cavity mode is in fact large ($\ensuremath{n_{\text{ph}}} \approx 100$) and absorption of photons is commonplace, as indicated by the Shapiro-like steps in the $I$-$V$\ characteristics of the CPT\@. Experimentally, this is a consequence of the very large $Q$ of the cavity in our experiment, which ensures that the photons generated by Cooper pair tunneling remain to interact with subsequent tunneling events for long after they are generated. Quantitatively, the product of the Cooper pair tunneling rate $I/2e$ and the photon dwell time of the cavity $2 \pi/\kappa$ is in our case $\pi I/\kappa e = 480$; for related experiments\cite{Hofheinz:2011} in a regime for which $P(E)$ theory is applicable we estimate $\pi I/\kappa e\approx 2$, over two orders of magnitude smaller. Recent experiments on lasing in a related single-electron transistor/cavity system have a photon dwell time comparable to that reported here.\cite{Astafiev:2007} However, in that case charge transport was based on a combination of Cooper pair and quasiparticle tunneling. The quasiparticle transitions served to establish population inversion of a three level system similar to that of single atom lasers.\cite{McKeever:2003} The typical photon emission rate for this artificial superconducting atom of roughly \amount{200}{MHz} was small compared to the quasiparticle tunneling rate, typically 3--\amount{5}{GHz}. As a result, any interactions between tunneling Cooper pairs and the cavity photons were incoherent. In contrast, for the single-Cooper-pair Josephson laser transport occurs via cotunneling of Cooper pairs, a very different transport process not clearly analogous to single-atom lasing. Furthermore, the decoherence-inducing influences of the electromagnetic environment and quasiparticle tunneling are vastly reduced relative to previous results. The result is a new regime of hybrid electronic/photonic transport that cannot be adequately treated within the framework of conventional $P(E)$ theory. Some aspects of this behavior can be explained within a semiclassical model of the cCPT system (see Supplementary Information for details). The emission of cavity photons at \ensuremath{\omega_{0}}\ for both $\ensuremath{\omega_{d}} = \ensuremath{\omega_{0}}/2$ and \ensuremath{\omega_{0}}\ resonances can be explained dynamically by considering the time dependence of the drive term $\sigma_{x} \cos[\Delta (a + a^{\dagger}) + \ensuremath{\omega_{d}} t]$ in the cCPT Hamiltonian. It can be shown that $\sigma_{x}$ oscillates at odd harmonics of the drive frequency \ensuremath{\omega_{d}}; the nonlinearity of the system in the form of the product of $\sigma_{x}$ with a sinusoidal function of \ensuremath{\omega_{d}}\ leads to an overall oscillation at the cavity resonant frequency \ensuremath{\omega_{0}}. The semiclassical analysis also correctly predicts the approximate number of photons \ensuremath{n_{\text{ph}}}\ in the cavity for both cases, as well as the vanishing of the emission at the charge degeneracy points. Nonetheless, many aspects of the emission remain unexplained, making the cCPT emission a rich topic for continued theoretical investigation. \begin{figure}[t!] \includegraphics[width=9cm]{Fig3New.pdf} \caption{\label{fig3} {\bfseries Detailed emission spectra for $\ensuremath{\omega_{d}} \approx \ensuremath{\omega_{0}}/2$ . a,} Emission versus gate charge \ensuremath{n_{g}}\ and detuning $\Delta f$ at close to maximum emission. {\bfseries b--d,} Top panels: emission versus applied bias voltage \ensuremath{V_{\text{dc}}}\ and detuning $\Delta f$ for $\ensuremath{n_{g}}=0.55$, 0.62 and 0.69, as indicated by the white vertical lines in {\bfseries a}. \ensuremath{V_{\text{dc}}}\ was swept from low to high bias in \amount{80}{nV} increments. Bottom panels: cavity photon occupation $\ensuremath{n_{\text{ph}}} = P/\kappa\hbar\ensuremath{\omega_{0}}$ versus \ensuremath{V_{\text{dc}}}\ for each gate voltage where $P$ is the integrated emission power. All emission spectra are plotted versus detuning $\Delta f$. } \end{figure} Finally, we turn to the question of the nature of the photon field in the cavity, and whether it is best described as a quantum state. A classical electrical current is well known (ignoring fluctuations) to produce a coherent state of the electromagnetic field. Given the highly quantum nature of the electronic/photonic transport in the cCPT it seems likely that the cavity photon field has non-classical correlations. By analogy with single-emitter lasers, which lack the noise associated with multiple emitters and spontaneous emission processes,\cite{Haroche:2006,McKeever:2003} we expect that the cavity photons will be number squeezed; \textit{i.~e.}, the cavity photon statistics will be sub-Poissonian, with a Fano factor $F = (\langle \ensuremath{n_{\text{ph}}}^{2}\rangle -\langle\ensuremath{n_{\text{ph}}}\rangle^{2})/\langle\ensuremath{n_{\text{ph}}}\rangle$ less than unity. This expectation is supported by several theoretical calculations.\cite{Armour:2013} Far from the charge degeneracy points we may treat the CPT as an effective single junction system. Making this approximation and using a rotating wave approximation, we find that the predicted Fano factor for system parameters comparable to the experiment is in the ranges $F\approx 0.6$--0.8, which would place the system firmly in the quantum regime for which $F<1$. In fact, suppression of photon number fluctuations leading to a reduced Fano factor seems to be a generic property of the system, whether biased on the cotunneling features considered here or on the sequential tunneling features\cite{Marthaler:2011} also visible in Fig.~\ref{fig1}d. (See Supplementary Information for details.) Future experimental studies will focus on use of recently developed state reconstruction techniques using linear detectors\cite{Eichler:2011,Eichler:2012} to accurately characterize the quantum state of photons generated by the Josephson laser. \begin{figure*}[t!] \includegraphics[width=12cm]{Fig4New.pdf} \caption{\label{fig4} \bfseries Detailed emission spectra for $\ensuremath{\omega_{d}} \approx \ensuremath{\omega_{0}}$ . a,} Emission versus gate charge \ensuremath{n_{g}}\ at close to maximum emission. {\bfseries b--d,} Top panels: emission versus applied bias voltage \ensuremath{V_{\text{dc}}}\ for $\ensuremath{n_{g}}=0.41$, 0.61 and 0.80, as indicated by the vertical white lines in {\bfseries a}. \ensuremath{V_{\text{dc}}}\ was swept from low to high bias in \amount{75}{nV} increments. Bottom panels: cavity photon occupation $\ensuremath{n_{\text{ph}}} = P/\kappa\hbar\ensuremath{\omega_{0}}$ versus \ensuremath{V_{\text{dc}}}\ for each gate voltage where $P$ is the integrated emission power. All emission spectra are plotted versus detuning $\Delta f$. All emission spectra are plotted versus detuning $\Delta f$. } \end{figure*} In conclusion, we have demonstrated lasing by means of a new quantum coherent transport process involving the interaction of Cooper pairs and photons. The single-Cooper-pair Josephson laser may ultimately serve as a convenient, easy-to-use source of amplitude squeezed light, and could form the basis of a new class of electrical or photonic amplifiers. It could also serve as an important platform for the study of the quantum dynamics of strongly nonlinear systems. \section*{Acknowledgements} This work was supported by the NSF under grants DMR-1104790 and DMR-1104821, by AFOSR/DARPA under agreement FA8750-12-2-0339, and by EPSRC (U.K.) under Grant No. EP/I017828/1. \section*{Author Contributions} F.C. and J.L. fabricated the samples. F.C. performed the measurements with assistance from J.L. and J.S\@. A.D.A., M.P.B. and E.B. developed the theoretical model. A.J.S. and R.W.S. fabricated the coplanar waveguide resonators. A.J.R., M.P.B. and A.D.A. conceived and designed the experiment, and co-wrote the manuscript with input from all authors. \section*{Competing Financial Interests} The authors acknowledge no competing financial interests.
\section{Introduction} In this paper, we discuss a rather technical issue concerning the sign of the partition function of 3d ${\cal N}=2$ supersymmetric theories. It is defined by the path integral \begin{equation} Z_{\cal M}=\int{\cal D}\Phi e^{-S}, \label{zdef} \end{equation} where ${\cal M}$ is the background manifold and $\Phi$ collectively represents all dynamical fields in the theory. The overall factor is often ignored because it does not affect some observables such as correlation functions. However, in recent progress of supersymmetric field theories, the partition function itself plays an important role. For example, we can determine the superconformal R-charge of a 3d ${\cal N}=2$ theory at infra-red fixed point by maximizing the real part of the free energy $F=-\log Z_{{\bm S}^3}$ \cite{Jafferis:2010un,Jafferis:2011zi}. In order to compute the partition function including the overall factor unambiguously through the path integral (\ref{zdef}) we need to fix the measure of the path integral carefully. A convenient way to do this is to exploit the fact that $Z_{\cal M}$ is obtained from the supersymmetric index of a 4d ${\cal N}=1$ theory as the small radius limit. Let us consider the case of ${\bm S}^3$ partition function. To obtain the 3d theory, we start from a 4d ${\cal N}=1$ theory in the background ${\bm S}^3\times{\bm S}^1$. We define ${\bm S}^3$ by \begin{equation} |z_1|^2+|z_2|^2=1,\quad z_1,z_2\in\mathbb{C}. \label{s3inc2} \end{equation} If we regard the ${\bm S}^1$ direction as a time, the path integral of the 4d theory is interpreted as the index \begin{equation} I(p_1,p_2,z_a)={\rm tr}\left[ (-1)^Fq^{D-\frac{R}{2}} (q^{-1}p_1)^{J_{1}+\frac{R}{2}} (q^{-1}p_2)^{J_{2}+\frac{R}{2}} z_a^{F_a} \right], \label{4dindex} \end{equation} where $F$, $R$, $D$, and $F_a$ are the fermion number, the $U(1)_R$ charge, the dilatation, and flavor charges, respectively. $J_1$ and $J_2$ are the angular momenta rotating $z_1$ and $z_2$, respectively. The exponent of $q$ in (\ref{4dindex}) \begin{equation} D-J_1-J_2-\frac{3}{2}R=\{Q,Q^\dagger\} \end{equation} is exact with respect to a supercharge $Q$, and (\ref{4dindex}) is independent of the variable $q$. Unlike the partition function $Z_{{\bm S}^3}$, there is a natural normalization of $I$; in the trace over the Hilbert space, every gauge invariant state contributes to the index by weight $1$, and there is no ambiguity of the normalization except for the signature. The ${\bm S}^3$ partition function is obtained as the $\beta\rightarrow 0$ limit of the index. In non-supersymmetric theories the small radius limit may in general diverge. However, in the reduction from a 4d ${\cal N}=1$ theory to a 3d ${\cal N}=2$ theory that we consider here, we can obtain a finite result due to the cancellation between bosonic and fermionic contributions.\footnote { For this cancellation we should carefully include the zero-point contribution, which is often neglected. } Therefore, once we obtain the 4d index, we can unambiguously obtain the partition function by the small radius limit \cite{Dolan:2011rp,Gadde:2011ia,Imamura:2011uw}. \begin{equation} Z_{{\bm S}^3}(b,\mu_a)= \lim_{\beta\rightarrow 0}I(p_i=e^{-\beta\omega_i},z_a=e^{-\beta\mu_a}),\quad b=\sqrt{\frac{\omega_1}{\omega_2}}. \label{smallradius} \end{equation} $\omega_i$ and $\mu_a$ are interpreted in the 3d theory as squashing parameters and real mass parameters. The partition function depends on $\omega_i$ through the single parameter $b$ \cite{Imamura:2011uw,Hama:2011ea,Imamura:2011wg,Alday:2013lba,Nian:2013qwa,Closset:2013vra,Tanaka:2013dca}. The ambiguity in the signature is due to the ambiguity in the statistics of the vacuum state. The statistics of states in the Hilbert space is fixed once that of the vacuum state is specified. However, there is no general rule to fix it, and we need an additional criterion to fix the overall sign. For some use of $Z_{\cal M}$, like $F$ maximization, we only need the absolute value of $Z_{\cal M}$, and one may think that the sign ambiguity is not important. However, if the theory has multiple sectors, we should sum up their contributions, and we need to fix the relative signs among them. This is the case when we consider a gauge theory on a manifold with non-trivial fundamental group. In such a case there are degenerate vacua labeled by holonomies associated with non-trivial cycles. In this paper we focus on the orbifold ${\bm S}^3/\mathbb{Z}_n$ defined from ${\bm S}^3$ in (\ref{s3inc2}) by the identification \begin{equation} (z_1,z_2)\sim (\omega z_1,\omega^{-1}z_2)\quad \omega=e^{\frac{2\pi i}{n}}. \label{zndef} \end{equation} The fundamental group of this manifold is $\mathbb{Z}_n$, and vacua are labeled by $\mathbb{Z}_n$-valued holonomies $h_a^{\rm (dyn)}$ associated with dynamical $U(1)_a$ gauge symmetries as well as continuous moduli parameters $\mu_a^{({\rm dyn})}$. It is also possible to introduce mass parameters $\mu _a^{\rm (ext)}$ and non-trivial holonomies $h_a^{\rm (ext)}$ for global $U(1)_a$ symmetries. (We label both gauge and global symmetries by $a$.) The partition function is obtained by summing up the contribution from sectors with different $h_a^{\rm (dyn)}$ \begin{equation} Z_{{\bm S}^3/\mathbb{Z}_n}(b,\mu_a^{(\rm ext)},h_a^{(\rm ext)}) =\sum_{h_a^{(\rm dyn)}}f(h)\int d\mu^{({\rm dyn})} e^{-S_{{\bm S}^3/\mathbb{Z}_n}^{\rm cl}}Z^{\rm 1-loop}_{{\bm S}^3/\mathbb{Z}_n}(b,\mu_a,h_a), \label{zorb} \end{equation} where $f(h)$ is the sign factor that we would like to determine, and the explicit form of the integrand will be given in the next section. A formula for the orbifold partition function has actually been already given in \cite{Benini:2011nc}. They derive the formula in two ways. One is the orbifold projection from ${\bm S}^3$ partition function, and the other is reduction from the lens space index, which is obtained from ${\bm S}^3\times{\bm S}^1$ index by the orbifold projection. In both derivations, they do not take account of the possible emergence of non-trivial sign factors. The formula has been used for some applications, and works well. However, in some cases, we need to introduce extra sign factors. For example, it is demonstrated in \cite{Imamura:2012rq} for a few examples of dual pairs that the matching of the orbifold partition function of dual theories requires non-trivial sign factors. A similar problem of relative weight also arises in the instanton sum in the ${\bm S}^4$ partition function. It would be instructive to understand how we can determine the relative weights in that case, before we explain our strategy for the ${\bm S}^3/\mathbb{Z}_n$ partition function. Let us consider an ${\cal N}=2$ supersymmetric gauge theory on ${\bm S}^4$. By equivariant localization, we can localize the dynamics of the theory at two poles of ${\bm S}^4$, and the partition function is written as \cite{Pestun:2007rz,Gomis:2011pf} \begin{equation} Z=\sum_{k_N,k_S=0}^{\infty}f(k_N,k_S)Z_N(k_N)Z_S(k_S), \end{equation} where $k_N$ and $k_S$ are, respectively, the instanton number at the north pole and the anti-instanton number at the south pole. (Precisely, we need to perform the integral over the Coulomb branch parameterized by constant scalar fields. Here we focus only on the instanton sum, and consider the contribution from a specific point in the Coulomb branch.) We introduced the unknown phase factor $f(k_N,k_S)$. This phase factor is strongly restricted by assuming the locality of the theory \cite{Callan:1976je}. If we assume the locality, the path integral for the localized modes at the two poles should be performed independently, and thus the partition function is factorized into contributions from the poles; \begin{equation} Z= \sum_{k_N=0}^{\infty} g(k_N)Z_N(k_N) \sum_{k_S=0}^{\infty} h(k_S)Z_S(k_S), \end{equation} where $g(k_N)$ and $h(k_S)$ are unknown phase factors. Now we use the fact that disconnected components of the configuration space of the theory is labeled only by the total instanton number $k_N-k_S$. This means that two configurations labeled by $(k^{(1)}_N,k^{(1)}_S)$ and $(k^{(2)}_N,k^{(2)}_S)$ are in the same component of the configuration space if $k_N^{(1)}-k_S^{(1)}=k_N^{(2)}-k_S^{(2)}$. We can interpolate them by continuous deformation and the relative phase between the contributions from them can be in principle determined unambiguously by the continuity of the action functional. Here let us assume for simplicity that there are no relative phases. Namely, $g(k_N+n)h(k_S+n)$ does not depend on $n$. Then, the relation \begin{equation} \frac{g(k_N+1)}{g(k_N)} =\frac{h(k_S)}{h(k_S+1)} \label{gh} \end{equation} holds. The left-(right-)hand side of this equation is independent of $k_S$ ($k_N)$, and (\ref{gh}) is a constant independent of both $k_N$ and $k_S$. Let $c$ be the constant. We obtain \begin{equation} g(k_N)=g(0)c^{k_N},\quad h(k_S)=h(0)c^{-k_S}, \end{equation} and the total partition function is \begin{equation} Z=f(0,0)\sum_{k_N,k_S=0}^{\infty}c^{k_N-k_S}Z_N(k_N)Z_S(k_S). \end{equation} Now we have determined the phase factor except the overall phase $f(0,0)$ and the constant $c$. The factor $c^{k_N-k_S}$ can be identified with the contribution of the topological $\theta$ term. In this way, the relative phases for instanton sectors of a 4d gauge theory can be fixed up to few parameters by the factorization of the partition function. We take the same strategy to determine the relative signs among holonomy sectors of the orbifold partition function. Actually, it is known that the ${\bm S}^3$ and ${\bm S}^2\times{\bm S}^1$ partition functions are factorized into factors so-called holomorphic blocks \cite{Krattenthaler:2011da,Pasquetti:2011fj,Dimofte:2011py,Beem:2012mb,Taki:2013opa}, and a similar factorization is expected for the orbifold. Each block is identified with the vortex partition function on a solid torus \cite{Pasquetti:2011fj,Beem:2012mb}. In the following, we determine the relative signs of holonomy sectors by requiring the factorization of the orbifold partition function. \section{Orbifold partition function} \subsection{Naive projection} Let us first summarize how the formula for the orbifold partition function $Z_{{\bm S}^3/\mathbb{Z}_n}$ is obtained from the ${\bm S}^3$ partition function by naive $\mathbb{Z}_n$ orbifold projection \cite{Benini:2011nc}. We consider a theory with general gauge group $G$ and matter representation $R$. At a generic point in the Coulomb branch the gauge group $G$ is broken to its Cartan subgroup $H$. Let $V_a$, $W_\alpha$, and $\Phi_i$ denote the vector multiplets for the Cartan part, W-bosons, and chiral multiplets, respectively. For later convenience, we include external vector multiplets for global $U(1)_a$ symmetries in $V_a$. For distinction, we denote dynamical and non-dynamical components by $V_a^{({\rm dyn})}$ and $V_a^{({\rm ext})}$, respectively. The scalar components $\mu_a^{({\rm dyn})}$ for the dynamical vector multiplets parameterize the Coulomb branch while those for external vector multiplets, $\mu_a^{({\rm ext})}$, are real mass parameters. By localization, we can reduce the path integral of an ${\cal N}=2$ supersymmetric theory on ${\bm S}^3$ into a finite dimensional matrix integral with the integrand consisting of the classical and one-loop factors: \begin{equation} Z_{{\bm S}^3}(\mu^{({\rm ext})})=\int d\mu^{({\rm dyn})} e^{-S^{\rm cl}_{{\bm S}^3}(\mu)}Z_{{\bm S}^3}^{\rm 1-loop}(\mu). \label{zs3} \end{equation} The integration measure is given by \begin{equation} \int d\mu^{({\rm dyn})}\equiv \frac{1}{|W|}\prod_a\int_{-\infty}^\infty d\mu^{({\rm dyn})}_a \end{equation} where $a$ runs over dynamical part of $V_a$, and $|W|$ is the order of the Weyl group of the gauge group. The one-loop factor is the product of the contributions of $W_\alpha$ and $\Phi_i$ \begin{equation} Z_{{\bm S}^3}^{\rm 1-loop}=\frac{1}{\prod_\alpha s_b(\mu_\alpha+\frac{iQ}{2})} \frac{1}{\prod_is_b(\mu_i-\frac{iQ}{2}(1-\Delta_i))}, \label{eq15} \end{equation} where $Q=b+b^{-1}$. $\Delta_i$ is the Weyl weight of the scalar component of a chiral multiplet $\Phi_i$. $\mu_\alpha$ and $\mu_i$ are the scalar components of the vector multiplets coupling to $W_\alpha$ and $\Phi_i$; \begin{equation} \mu_\alpha=q_{\alpha a}\mu_a,\quad \mu_i=q_{ia}\mu_a, \label{muamui} \end{equation} where $q_{\alpha a}$ and $q_{ia}$ are the $U(1)_a$ charge of W-boson $W_\alpha$ and the chiral multiplet $\Phi_i$, respectively. It is convenient to include R-charge in the charge matrix. We define $q_{\alpha 0}$ and $q_{i0}$ as the R-charges of the fermions in the multiplets $W_\alpha$ and $\Phi_i$, \begin{equation} q_{\alpha 0}=1,\quad q_{i0}=\Delta_i-1, \end{equation} and we set the corresponding scalar parameter by \begin{equation} \mu_0\equiv\mu_R=\frac{iQ}{2}. \end{equation} Including the contribution of the $U(1)_R$ symmetry, we define \begin{equation} \widehat\mu_\alpha=\mu_\alpha+\frac{iQ}{2},\quad \widehat\mu_i=\mu_i-\frac{iQ}{2}(1-\Delta_i). \end{equation} Then the one-loop factor (\ref{eq15}) is simply rewritten as \begin{equation} Z_{{\bm S}^3}^{\rm 1-loop}=\frac{1}{\prod_I s_b(\widehat\mu_I)}, \end{equation} where $I$ runs over both W-bosons and chiral multiplets. $s_b(z)$ is the double sine function, and can be expressed as an infinite product corresponding to the spherical harmonic expansion on ${\bm S}^3$. The contribution of a multiplet $I$ is \begin{equation} \frac{1}{s_b(\widehat\mu_I)} =\prod_{p,q=0}^\infty \frac{b(q+\frac{1}{2})+b^{-1}(p+\frac{1}{2})+i\widehat\mu_I} {b(p+\frac{1}{2})+b^{-1}(q+\frac{1}{2})-i\widehat\mu_I}. \label{s3oneloop} \end{equation} The denominator and the numerator come from the bosonic and the fermionic modes with angular momenta $(J_1,J_2)=(p,q)$, respectively. The classical factor exists when the theory has Chern-Simons terms. If the action contains the Chern-Simons term \begin{equation} \frac{ik_{ab}}{4\pi}\int A_adA_b, \label{eq22} \end{equation} the scalar field quadratic terms in the supersymmetric completion of (\ref{eq22}) give the classical factor \begin{equation} e^{-S_{{\bm S}^3}^{\rm cl}(\mu)}=e^{-\pi i k_{ab}\mu_a\mu_b}. \label{cscontribution} \end{equation} Let us move on to the orbifold partition function $Z_{{\bm S}^3/\mathbb{Z}_n}$. The orbifold ${\bm S}^3/\mathbb{Z}_n$ is defined from the ${\bm S}^3$ in (\ref{s3inc2}) by the identification (\ref{zndef}). The vacua are parameterized by the scalar field $\mu_a$ and holonomies \begin{equation} h_a=\frac{n}{2\pi}\oint _\gamma A_a, \label{holodef} \end{equation} where $\gamma$ is a non-trivial loop in ${\bm S}^3/\mathbb{Z}_n$ generating the fundamental group. $h_a$ is quantized to be integer, and $h_a$ and $h_a+n$ are identified because they are transformed to each other by a large gauge transformation. We define $h_I=\{h_\alpha,h_i\}$, holonomies coupling to $W_\alpha$ and $\Phi_i$, in a similar way to (\ref{muamui}): \begin{equation} h_I=\sum_a q_{Ia}h_a. \label{holonomyI} \end{equation} We can turn on non-trivial holonomies for global symmetries as well as the dynamical gauge symmetries. Note that we will not turn on the holonomy for the R-symmetry because it breaks supersymmetry. (For a more general lens space $L(p,q)$, we need to turn on non-trivial $U(1)_R$ holonomy to preserve supersymmetry.) After the orbifold projection, only modes compatible with the identification (\ref{zndef}) contribute to the one-loop factor. The condition for $\mathbb{Z}_n$ invariance for modes of multiplet $I$ is \begin{equation} p-q=h_I\mod n, \label{pqhn} \end{equation} and the one-loop factor is obtained by restricting the product over $p$ and $q$ in (\ref{s3oneloop}) by the condition (\ref{pqhn}). We define the function $s_{b,h_I}$ to express the contribution of each multiplet by \begin{equation} \frac{1}{s_{b,h_I}\left(\widehat\mu_I\right)} =\prod_{(p,q)\in\Lambda_{[h_I]}} \frac{b(q+\frac{1}{2})+b^{-1}(p+\frac{1}{2})+i\widehat\mu_I}{b(p+\frac{1}{2})+b^{-1}(q+\frac{1}{2})-i\widehat\mu_I}, \label{sbh1} \end{equation} where $\Lambda_{[h]}$ is the set consisting of $(p,q)$ satisfying the condition (\ref{pqhn}): \begin{equation} \Lambda_{[h]} =\{(p,q)|p,q\geq0,p-q=h\mod n\}. \end{equation} The one-loop factor (\ref{s3oneloop}) is replaced by \begin{equation} Z_{{\bm S}^3/\mathbb{Z}_n}^{\rm 1-loop} =\prod_I\frac{1}{s_{b,h_I}\left(\widehat\mu_I\right)}. \end{equation} For the classical part, the $\mathbb{Z}_n$ orbifolding gives rise to the extra $1/n$ factor in the action, and the Chern-Simons term gives the holonomy dependent phase \cite{Gang:2009wy,0209403,Griguolo:2006kp}: \footnote{We find a slightly different formula $e^{\frac{\pi i}{n}k_{ab}h_ah_b}$ for the holonomy dependent phase in the literature. This is, however, not gauge invariant even for integer Chern-Simons levels. A simple derivation of the holonomy dependent phase in (\ref{cscont}) is given in Appendix \ref{app:CSterm}. } \begin{equation} e^{-S_{{\bm S}^3/\mathbb{Z}_n}^{\rm cl}(\mu,h)} =e^{-\frac{\pi i}{n}k_{ab}\mu_a\mu_b} e^{-\frac{\pi i k_{ab}}{n}(n-1)h_ah_b}. \label{cscont} \end{equation} The integration measure is \begin{equation} \int d\mu^{({\rm dyn})} =\frac{1}{|W|}\prod_a\int_{-\infty}^\infty\frac{d\mu_a^{({\rm dyn})}}{n}. \end{equation} This is normalized by using the relation to the lens space index \cite{Benini:2011nc}. Combining these factors, we obtain the formula (\ref{zorb}). Before ending this section, we would like to comment on a subtlety lurking in the formula (\ref{zorb}). For the gauge invariance of the factor (\ref{cscont}), the Chern-Simons levels $k_{ab}$ must be integers. This is, however, not always the case. If the theory has parity anomaly, some components of the bare Chern-Simons level should be half odd integers to cancel the anomaly. In this case, the factor (\ref{cscont}) itself is not gauge invariant. Namely, it may change its sign under a large gauge transformation that shifts $h_a$ by $h_a\rightarrow h_a+nc_a$ ($c_a\in\mathbb{Z}$). Of course, this is not an essential problem. For the consistency, we only need the gauge invariance of the whole integrand in (\ref{zorb}) including the one-loop contribution. In the following, we propose a general formula for the sign factor $f(h)$, which will give the sign of the partition function for each holonomy sector in a gauge invariant way. \subsection{Projection operator} \label{subsec:pairing} As we mentioned in the introduction, we use the factorization to holomorphic blocks to determine the sign factor. For ${\cal M}={\bm S}^3$ and ${\bm S}^2\times {\bm S}^1$, it is known that the partition function is written in terms of holomorphic blocks by \cite{Krattenthaler:2011da,Pasquetti:2011fj,Dimofte:2011py,Beem:2012mb,Taki:2013opa} \begin{equation} Z_{\cal M}=\sum_A B^A(x_a,q)B^A(\widetilde x_a,\widetilde q). \label{factorization} \end{equation} In the case of ${\bm S}^3$, the variables $x_a$, $q$, $\widetilde x_a$, and $\widetilde q$ are given by \begin{equation} q=e^{2\pi i b^2},\quad x_a=e^{2\pi b\mu_a},\quad \widetilde q=e^{2\pi i b^{-2}},\quad \widetilde x_a=e^{2\pi b^{-1}\mu_a}. \label{s3arguments} \end{equation} This factorization is also expected for the orbifold partition function with different definition for the variables $x_a$, $q$, $\widetilde x_a$, and $\widetilde q$. This factorization is naturally interpreted in Higgs branch localization, in which the index $A$ labels Higgs vacua and the blocks are identified with the vortex partition functions\cite{Pasquetti:2011fj,Beem:2012mb}. In order to obtain a factorized form of the orbifold partition function, it is convenient to rewrite $s_{b,h_I}$ in (\ref{sbh1}) as \begin{equation} \frac{1}{s_{b,h_I}\left(\widehat\mu_I\right)} = {\cal P}_{(\widehat\mu_I,h_I)} \frac{1}{s_b(\widehat\mu_I)} \label{ps} \end{equation} where ${\cal P}_{(z,h)}$ with $z\in\mathbb{C}$ and $h\in\mathbb{Z}_n$ is the operator acting on a function of $z$ defined by \begin{equation} {\cal P}_{(z,h)} f(z) =\prod_{(k,l)\in L_{h}} f(z_{k,l}),\quad z_{k,l}\equiv \frac{z+ib(k-\frac{n-1}{2})+ib^{-1}(l-\frac{n-1}{2})}{n}, \label{Pdef} \end{equation} where \begin{equation} L_{h}=\{(k,l)| 0\leq k,l<n,\ k-l=h\mod n\}. \end{equation} An advantage of rewriting (\ref{sbh1}) with this operator is that the operator preserves the factorized form of the function. Namely, if a function $f(z)$ is the product of two functions $g(z)$ and $h(z)$, the relation ${\cal P}_{(z,h)}f(z)=({\cal P}_{(z,h)}g(z))({\cal P}_{(z,h)}h(z))$ holds. Therefore, if the ${\bm S}^3$ partition function $Z_{{\bm S}^3}(\widehat\mu)$ of a theory is factorized into holomorphic blocks as in (\ref{factorization}), and if the orbifold partition function is obtained from $Z_{{\bm S}^3}(\widehat\mu)$ by applying the operator ${\cal P}_{(\widehat\mu,h)}$, we can immediately obtain the factorized form of the orbifold partition function by applying ${\cal P}_{(\widehat\mu,h)}$ to the holomorphic blocks for ${\bm S}^3$. The operator ${\cal P}_{(z,h)}$ is defined to simplify the expression of the $\mathbb{Z}_n$ projection of the one-loop factor, and it is a priori not guaranteed that it correctly reproduces the classical factor $e^{-S^{\rm cl}_{{\bm S}^3/\mathbb{Z}_n}}$ in the orbifold partition function. Interestingly, up to the sign factor which we have not fixed yet, it reproduces the classical factor (\ref{cscont}) in the orbifold partition function from (\ref{cscontribution}) for ${\bm S}^3$. Let us consider a Chern-Simons term with factorized Chern-Simons level $k_{ab}=\kappa c_ac_b$. \begin{equation} \frac{i\kappa c_ac_b}{4\pi}\int A_adA_b. \label{cacb} \end{equation} For ${\bm S}^3$, this gives the classical factor \begin{equation} e^{-S^{\rm cl}_{{\bm S}^3}}=e^{-\kappa\pi i\widehat\mu^2},\quad \widehat\mu=c_a\mu_a. \end{equation} Applying the operator ${\cal P}_{(\widehat\mu,h)}$ on this function, we obtain \begin{align} {\cal P}_{(\widehat\mu,h)}e^{-\kappa\pi i\widehat\mu^2} &= e^{\frac{\pi i\kappa}{n}\frac{(b+b^{-1})^2}{12}(n^2-1)} \exp\left[ -\frac{\pi i\kappa}{n} \left(\widehat\mu^2 +[h](n-[h])\right) \right], \label{csxxxx} \end{align} where $h=c_ah_a$ and $[h]$ denotes the smallest non-negative integer in $h+n\mathbb{Z}$. (We have not yet assumed that $\kappa$ is an integer.) We rewrite this as \begin{equation} {\cal P}_{(\widehat\mu,h)}e^{-\kappa\pi i\widehat\mu^2} =(\mbox{$b$-dependent factor}) \rho(h)^{2\kappa} \exp\left[-\frac{\pi i\kappa}{n}\left(\widehat\mu^2 +(n-1)h^2\right) \right]. \label{eq40} \end{equation} In this paper, the factorization is used simply as a criterion for the correct choice of sign factors, and we are not interested in the prefactor depending only on $b$. The exponential factor is nothing but the classical factor in (\ref{cscont}), and $\rho(h)$ is \begin{equation} \rho(h)= e^{\frac{\pi i}{2n}[h](n-[h])} e^{-\frac{\pi i}{2n}(n-1)h^2}. \label{rhoh} \end{equation} This function always takes $+1$ or $-1$ depending on $h$. As we will show shortly, we can compose the sign factor $f(h)$ by using $\rho(h)$. \subsection{Factorization and sign factor} Let us consider a general non-gauge theory on ${\bm S}^3$. As is pointed out in \cite{Beem:2012mb}, it is convenient to decompose the Chern-Simons level into the part canceling the parity anomaly and the remaining part. \begin{equation} k_{ab}=\sum_\alpha\kappa_\alpha c_{\alpha a} c_{\alpha b} -\frac{1}{2}\sum_Iq_{Ia}q_{Ib}. \label{decomposition} \end{equation} $\kappa_\alpha$ and $c_{\alpha a}$ in the first term are integers, and the second term is the fractional contribution that cancels the parity anomaly. With this decomposition, we rewrite the partition function as \begin{equation} Z_{{\bm S}^3}(\mu)=e^{-\pi i k_{ab}\mu_a\mu_b}\frac{1}{\prod_Is_b(\widehat\mu_I)} =\prod_\alpha (e^{-\pi i \mu_\alpha^2})^{\kappa_\alpha} \prod_I Z_\Delta(\widehat\mu_I) \label{froduct} \end{equation} where $Z_\Delta(\widehat\mu)$ is the partition function of the ``tetrahedron theory,''\cite{Dimofte:2011ju} and given by \begin{equation} Z_\Delta(\widehat\mu)=\frac{e^{\frac{\pi i}{2}\widehat\mu^2}}{s_b(\widehat\mu)}. \label{tetra} \end{equation} The two kinds of factors in the product (\ref{froduct}) are known to be factorized into holomorphic blocks \cite{Beem:2012mb} \begin{align} Z_\Delta(\mu) & = e^{-\pi i\frac{b^2+b^{-2}}{24}} B_\Delta(x,q)B_\Delta(\widetilde x,\widetilde q),\label{zdelta}\\ e^{-\pi i \mu^2} &= e^{\pi i\frac{b^2+b^{-2}}{12}} B_{\rm CS}(x;q) B_{\rm CS}(\widetilde x;\widetilde q).\label{zcs} \end{align} The blocks are given by \begin{equation} B_\Delta(x;q)=(qx^{-1};q),\quad B_{\rm CS}(x;q)=\frac{1}{(-q^{\frac{1}{2}}x;q) (-q^{\frac{1}{2}}x^{-1};q)}, \label{holocs} \end{equation} where $(x;q)$ is the q-Pochhammer symbol \begin{equation} (x;q)=\prod _{k=0}^{\infty}(1-xq^k). \end{equation} An important feature of the factorization is that the information of the background manifold is encoded in the definition of the arguments of holomorphic blocks. For ${\bm S}^3$ they are given by (\ref{s3arguments}), and the blocks for the orbifold should be given by the same functions with different arguments. We can confirm this by computing the holomorphic blocks for the orbifold by applying the operator ${\cal P}_{(z,h)}$ to the holomorphic blocks for ${\bm S}^3$. Indeed, we can easily show \begin{eqnarray} && {\cal P}_{(\mu,h)}B_\Delta(x;q)=B_\Delta(x';q'),\quad {\cal P}_{(\mu,h)}B_\Delta(\widetilde x;\widetilde q)=B_\Delta(\widetilde x';\widetilde q'), \nonumber\\ && {\cal P}_{(\mu,h)} B_{\rm CS}(x;q)=B_{\rm CS}(x',q'),\quad {\cal P}_{(\mu,h)} B_{\rm CS}(\widetilde x;\widetilde q) =B_{\rm CS}(\widetilde x',\widetilde q'), \end{eqnarray} where the variables for the orbifold are \begin{equation} q'=\omega q^{\frac{1}{n}},\quad x'=\omega^h x^{\frac{1}{n}},\quad \widetilde q'=\omega\widetilde q^{\frac{1}{n}},\quad \widetilde x'=\omega^{-h}\widetilde x^{\frac{1}{n}}. \label{orbpairing} \end{equation} The arguments above guarantee that we obtain the orbifold partition function that is correctly factorized into the holomorphic blocks by simply applying the projection operator to the factors in ${\bm S}^3$ partition function (\ref{froduct}). \begin{align} Z_{{\bm S}^3/\mathbb{Z}_n}(\mu,h) &\propto \prod_\alpha ({\cal P}_{(\mu_\alpha,h_\alpha)}e^{-\pi i \mu_\alpha^2})^{\kappa_\alpha} \prod_I {\cal P}_{(\widehat\mu_I,h_I)}Z_\Delta(\widehat\mu_I) \nonumber\\ &\propto \prod_\alpha e^{\frac{\pi i}{n}\kappa_\alpha(\widehat\mu_\alpha^2+(n-1)h_\alpha^2)} \prod_I \frac{\rho(h_I)e^{\frac{\pi i}{2n}(\widehat\mu_I^2+(n-1)h_I^2)}}{s_{b,h_I}(\widehat\mu_I)} \nonumber\\ &= e^{-S^{\rm cl}_{{\bm S}^3/\mathbb{Z}_n}(\mu,h)} \prod_I \frac{\rho(h_I)}{s_{b,h_I}(\widehat\mu_I)} \end{align} ``$\propto$'' means the ignorance of prefactors depending only on $b$. By comparing this to (\ref{zorb}), we obtain \begin{equation} f(h)=\prod_I\rho(h_I). \end{equation} (We cannot fix the overall sign factor independent of $h$, which we are not interested in.) We can absorb the sign factor by the redefinition of the orbifold double sine function \begin{equation} s_{b,h}^{\rm imp}(z)=\rho(h)s_{b,h}(z), \label{improved} \end{equation} and then we can present the orbifold partition function in the same form with the original one. \begin{equation} Z_{{\bm S}^3/\mathbb{Z}_n}(\mu,h)=e^{-S_{{\bm S}^3/\mathbb{Z}_n}^{\rm cl}(\mu,h)}\frac{1}{\prod_Is_{b,h_I}^{\rm imp}(\widehat\mu_I)}. \label{improvedz} \end{equation} In \cite{Imamura:2012rq}, a similar improvement of the orbifold double sine function $\widehat s_{b,h}(z)=\sigma_h s_{b,h}(z)$ is proposed for odd $n=2m+1$. The extra sign factor $\sigma_h$ is related to $\rho(h)$ through \begin{equation} \sigma_h=(-)^{mh}\rho(h). \end{equation} In \cite{Imamura:2012rq}, only theories without parity anomaly are considered. This means that the charge assignment $q_{Ia}$ for every $U(1)$ gauge symmetry satisfies $\sum_I q_{Ia}\in 2\mathbb{Z}$. In such a case, the difference of $\sigma_h$ and $\rho(h)$ does not affect the partition function. In the case of even $n$, \cite{Imamura:2012rq} did not succeed in finding such an improvement. The reason is that in \cite{Imamura:2012rq} the sign factor $\sigma_h$ is assumed to be a periodic function of $h$ with period $n$. The function $\rho(h)$ does not satisfy this condition. It may change its sign under the shift $h\rightarrow h+n$. \begin{equation} \frac{\rho(h+n)}{\rho(h)}=(-1)^{(n-1)h+\frac{n(n-1)}{2}}. \end{equation} This means that the improved function $s_{b,h}^{\rm imp}$ may change its sign in the large gauge transformation $h\rightarrow h+n$. This, however, does not cause any problem. If the parity anomaly arising in the one-loop part is correctly canceled by the bare Chern-Simons term, the partition function (\ref{improvedz}) is invariant under the shift $h\rightarrow h+n$. \subsection{Gauge theories} In the previous subsection we gave a prescription to fix the relative signs among holonomy sectors. It is implemented by the redefinition of the orbifold double sine function (\ref{improved}). The orbifold partition function of an arbitrary non-gauge theory with this improvement is correctly factorized into holomorphic blocks. We expect that this improvement works for gauge theories, too. Namely, the orbifold partition function \begin{equation} Z_{{\bm S}^3/\mathbb{Z}_n}(\mu^{({\rm ext})},h^{({\rm ext})})=\sum_{h^{({\rm dyn})}}\int d\mu^{({\rm dyn})} e^{-S_{{\bm S}^3/\mathbb{Z}_n}^{\rm cl}(\mu,h)}\frac{1}{\prod_Is_{b,h}^{\rm imp}(\widehat\mu_I)} \label{improvedzgauge} \end{equation} for a gauge theory is correctly factorized into holomorphic blocks as in (\ref{factorization}). Unfortunately, we have not succeeded in proving this for an arbitrary gauge theory. We here consider two examples of gauge theories, SQED with $N_f=1$, and an $su(2)$ Chern-Simons theory with an adjoint chiral multiplet. \subsubsection{SQED} As the first example, let us consider SQED with one flavor $(q,\widetilde q)$. This theory has four $U(1)$ symmetries. One is a gauge symmetry $U(1)_G$, and the others are global symmetries. See Table \ref{table:1} for charge assignments. \begin{table}[htb] \caption{The charge assignments in the SQED and XYZ model. For $U(1)_R$ symmetry the charges of the fermion components are shown.} \label{table:1} \begin{center} \begin{tabular}{cccccccc} \hline \hline && $q$ & $\widetilde q$ & & $X$ & $Y$ & $Z$ \\ \hline $U(1)_R$ && $\Delta-1$ & $\Delta-1$ & & $-\Delta$ & $-\Delta$ & $2\Delta-1$ \\ $U(1)_G$ && $1$ & $-1$ & & - & - & - \\ $U(1)_V$ && $0$ & $0$ & & $1$ & $-1$ & $0$ \\ $U(1)_A$ && $1$ & $1$ && $-1$ & $-1$ & $2$ \\ \hline \end{tabular} \end{center} \end{table} $U(1)_R$ is an R-symmetry. $U(1)_A$ is a flavor symmetry acting on $q$ and $\widetilde q$ with charge $+1$. $U(1)_V$ is the topological symmetry, and the corresponding external gauge field $A_V$ couples to the $U(1)_G$ flux $dA_G$ through the Chern-Simons term \begin{equation} \frac{1}{2\pi}\int A_VdA_G. \end{equation} The partition function is \begin{equation} Z^{{\bm S}^3/\mathbb{Z}_n}_{\rm SQED} =\sum_{h_G=0}^{n-1}\int \frac{e^{\frac{2\pi i}{n}\mu_V\mu_G}e^{\frac{2\pi i}{n}h_Vh_G}} {s_{b,h_q}^{\rm imp}(\widehat\mu_q)s_{b,h_{\widetilde q}}^{\rm imp}(\widehat \mu_{\widetilde q})}\frac{d\mu_G}{n}, \label{zaqed} \end{equation} where $\widehat\mu_I$ and $h_I$ are defined by \begin{equation} \widehat\mu_q=(\Delta-1)\frac{iQ}{2}+\mu_A+\mu_G,\quad \widehat\mu_{\widetilde q}=(\Delta-1)\frac{iQ}{2}+\mu_A-\mu_G. \end{equation} \begin{equation} h_q=h_A+h_G,\quad h_{\widetilde q}=h_A-h_G. \end{equation} Let us confirm that this orbifold partition function is factorized into holomorphic blocks. Although it would not be difficult to directly prove the factorization by performing the integral by using the residue theorem, we take another way. In \cite{Imamura:2012rq}, it is confirmed that the orbifold partition function of the SQED with an appropriate choice of the sign factor coincides with that of the XYZ model, the system consisting of three chiral multiplets $X$, $Y$, and $Z$ interacting through the superpotential $W=XYZ$. Because XYZ model is a non-gauge theory and we have already proved the factorization of non-gauge theories, this duality relation guarantees the factorization of the partition function of the SQED. The charge assignments for XYZ model is also shown in Table \ref{table:1}, and the partition function is given by \begin{equation} Z^{{\bm S}^3/\mathbb{Z}_n}_{\rm XYZ}=\frac{1}{ s_{b,h_X}^{\rm imp}(\widehat\mu_X) s_{b,h_Y}^{\rm imp}(\widehat\mu_Y) s_{b,h_Z}^{\rm imp}(\widehat\mu_Z)} \label{zxtz} \end{equation} with the parameters \begin{equation} \widehat\mu_X=-\Delta\frac{iQ}{2}-\mu_A+\mu_V,\quad \widehat\mu_Y=-\Delta\frac{iQ}{2}-\mu_A-\mu_V,\quad \widehat\mu_Z=(2\Delta-1)\frac{iQ}{2}+2\mu_A. \end{equation} and \begin{equation} h_X=-h_A+h_V,\quad h_Y=-h_A-h_V,\quad h_Z=2h_A. \end{equation} It is easy to numerically check the coincidence of (\ref{zaqed}) and (\ref{zxtz}). If we use the result of \cite{Imamura:2012rq}, what we have to do to confirm the relation $Z_{\rm XYZ}=Z_{\rm SQED}$ is to show the sign factor determined in \cite{Imamura:2012rq} by requiring $Z_{\rm XYZ}=Z_{\rm SQED}$ is same as the extra sign factor introduced by replacing $s_{b,h}$ by $s_{b,h}^{\rm imp}$. Indeed, the product of the five sign factors corresponding to the five double sine functions in $Z_{\rm SQED}$ and $Z_{\rm XYZ}$ coincides with the factor given in \cite{Imamura:2012rq}. \begin{equation} \sigma(h_G,h_V,h_A)= \rho(h_q) \rho(h_{\widetilde q}) \rho(h_X) \rho(h_Y) \rho(h_Z). \end{equation} \subsubsection{$su(2)$ gauge theory} Next, as a simple example of non-Abelian gauge theory, we consider an $su(2)_1$ gauge theory coupled to an adjoint chiral multiplet $\Phi$. (We use $su(2)$ instead of $SU(2)$ to emphasize we do not specify the global structure of the gauge group.) Jafferis and Yin \cite{Jafferis:2011ns} proposed that this theory is dual to the theory consisting a single chiral multiplet $X\sim {\rm tr}\Phi ^2$. If we can show the matching of the partition functions for the dual pair the factorization of the $su(2)$ theory is guaranteed as the previous example. In general, we cannot completely specify a gauge theory only by local information. There may be different theories distinguished by global structure of the gauge group. In \cite{Aharony:2013hda} importance of such distinction in four-dimensional supersymmetric gauge theories is pointed out, and it is investigated how such theories are related to each other by Seiberg duality. The duality is checked in \cite{Razamat:2013opa} by matching the lens space index, which is sensitive to the global structure of the gauge group. Similar aspects in three-dimensional gauge theories are studied in \cite{Aharony:2013kma}. The $su(2)$ theory we discuss here is also an example of such a theory. It has only an adjoint chiral multiplet as a matter field, and no elementary fields are transformed by the center of $SU(2)$. Therefore, precisely speaking, there are two choices of the gauge group, $SU(2)$ and $SO(3)$. Although it is important problem to clarify how the different choices of the gauge group affect the duality and the factorization, we will not do any detailed analysis here. We only present the result of a preliminary analysis based on the numerical computation of the partition function. The symmetries and the charge assignments for the dual theories are shown in Table \ref{symJY}. \begin{table}[htb] \caption{The symmetries and the charge assignments in Jafferis-Yin duality. The $U(1)_R$ charges in the table are those for fermion component of the multiplets.} \label{symJY} \begin{center} \begin{tabular}{ccccc} \hline \hline && $\Phi$ & & $X$ \\ \hline $SU(2)_G$ && {\rm adj} && - \\ $U(1)_R$ && $\Delta-1$ & & $2\Delta-1$ \\ $U(1)_A$ && $1$ && $2$ \\ \hline \end{tabular} \end{center} \end{table} In the $su(2)$ theory, the action contains the Chern-Simons term \begin{equation} \frac{i}{4\pi}\int {\rm tr}_{\rm fund}\left(A_GdA_G-\frac{2i}{3}A_G^3\right) \label{su2csterm} \end{equation} for the dynamical gauge field, and \begin{align} \frac{(-3/2)i}{4\pi}\int ( (\Delta -1)A_R+A_A) d( (\Delta -1)A_R+A_A)+\frac{i}{4\pi}\int A_RdA_R \end{align} for the external $U(1)_R$ and $U(1)_A$ gauge fields. On the other hand, in the chiral free theory contains the Chern-Simons term \begin{align} \frac{(-1/2)i}{4\pi}\int ((2\Delta -1)A_R+2A_A)d((2\Delta -1)A_R+2A_A) \end{align} for the external gauge fields. We define the holonomy parameter $h$ for the dynamical $su(2)$ gauge group by \begin{equation} U=\exp\left(i\oint_\gamma A\right) = \begin{pmatrix} e^{\frac{\pi i h}{n}} \\ & e^{-\frac{\pi i h}{n}} \end{pmatrix}. \end{equation} If the gauge group is $SU(2)$, $U^n$ must be the unit matrix, and the holonomy is quantized by $h\in 2\mathbb{Z}$, while for $SO(3)$ gauge group $h$ can be an arbitrary integer. The periodicity of $h$ also depends on the global structure of the gauge group and the background manifold, and there are two possibilities, $h\sim h+2n$ or $h\sim h+n$. Here we will not argue which of these possibilities should be adopted. We simply compute the partition functions for all the holonomy sectors labeled by $h=0,1,\ldots,2n-1$, and infer from the duality which sectors should be summed up. The orbifold partition function of each holonomy sector of the $su(2)$ theory is \begin{align} Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A)&=\int \frac{d\mu _G}{2n}e^{-\frac{\pi i}{2n}\mu _G^2}e^{-\frac{\pi i}{2n}(n-1)h^2}e^{\frac{\pi i}{n}\left[ \frac{3}{2}\left\{ (\Delta -1)\frac{iQ}{2}+\mu _A \right\} ^2+\frac{Q^2}{4} \right]}e^{\frac{3\pi i}{2n}(n-1)h_A^2}\nonumber \\ &\quad \times \frac{1}{s^{\rm imp}_{b,h_{\phi _+}}(\widehat \mu _{\phi _+}) s^{\rm imp}_{b,h_{\phi _0}}(\widehat \mu _{\phi _0}) s^{\rm imp}_{b,h_{\phi _-}}(\widehat \mu _{\phi _-}) s^{\rm imp}_{b,h_{W^+}}(\widehat \mu _{W^+}) s^{\rm imp}_{b,h_{W^-}}(\widehat \mu _{W^-})}, \end{align} where the parameters are given by \begin{gather} \widehat \mu _{\phi _+} =(\Delta -1)\frac{iQ}{2}+\mu _A+\mu _G,\quad \widehat \mu _{\phi _0} =(\Delta -1)\frac{iQ}{2}+\mu _A,\quad \widehat \mu _{\phi _-} =(\Delta -1)\frac{iQ}{2}+\mu _A-\mu _G,\nonumber\\ \widehat \mu _{W^+}=\frac{iQ}{2}+\mu _G,\quad \widehat \mu _{W^-}=\frac{iQ}{2}-\mu _G,\\ h_{\phi _+}=h_A+h,\quad h_{\phi _0}=h_A,\quad h_{\phi _-}=h_A-h,\quad h_{W^+}=h,\quad h_{W^-}=-h. \end{gather} The orbifold partition function of the chiral free theory is \begin{align} Z^{{\bm S}^3/\mathbb{Z}_n}_X(h_A)=e^{\frac{\pi i}{2n}\left\{ (2\Delta -1)\frac{iQ}{2}+2\mu _A\right\} ^2}e^{\frac{2\pi i}{n}(n-1)h_A^2}\frac{1}{s^{\rm imp}_{b,h_X}(\widehat \mu _X)}, \end{align} where the parameters are given by \begin{align} \widehat \mu _X=(2\Delta -1)\frac{iQ}{2}+2\mu _A,\quad h_X=2h_A. \end{align} Numerical results are divided to four cases: \begin{itemize} \item In the case of $n\in 4{\mathbb Z}$, only the even sector coincides up to a constant factor. \begin{align} \sum _{h=0,2,\ldots ,2n-2}Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A)=2Z^{{\bm S}^3/\mathbb{Z}_n}_X(h_A). \end{align} \item In the case of $n\in 4{\mathbb Z}+1$, both even and odd sector coincides up to a constant factor. \begin{align} \sum_{h=0,2,\ldots ,2n-2} Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A) =\sum_{h=1,3,\ldots ,2n-1} Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A) =\sqrt2e^{\frac{\pi i}{4}}Z^{{\bm S}^3/\mathbb{Z}_n}_X(h_A). \end{align} \item In the case of $n\in 4{\mathbb Z}+2$, only the odd sector coincides up to a constant factor. \begin{align} \sum_{h=1,3,\ldots ,2n-1}Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A)=-2iZ^{{\bm S}^3/\mathbb{Z}_n}_X(h_A). \end{align} \item In the case of $n\in 4{\mathbb Z}+3$, both even and odd sector coincides with different constant factor. \begin{gather} \sum _{h=0,2,\ldots ,2n-2}Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A)=\sqrt2e^{\frac{\pi i}{4}}Z^{{\bm S}^3/\mathbb{Z}_n}_X(h_A),\\ \sum_{h=1,3,\ldots ,2n-1}Z^{{\bm S}^3/\mathbb{Z}_n}_{su(2)}(h,h_A)=-\sqrt2e^{\frac{\pi i}{4}}Z^{{\bm S}^3/\mathbb{Z}_n}_X(h_A). \end{gather} \end{itemize} In all cases, we do not need additional sign factor if we use the improved function $s_{b,h}^{\rm imp}$. These results strongly suggests that we should take the following holonomy sectors to sum up. \begin{itemize} \item $n\in 4\mathbb{Z}$: $h=0,2,\ldots,2n-2$. \item $n\in 4\mathbb{Z}+1$: $h=0,2,\ldots,2n-2$ or $h=1,3,\ldots,2n-1$. \item $n\in 4\mathbb{Z}+2$: $h=1,3,\ldots,2n-1$. \item $n\in 4\mathbb{Z}+3$: $h=0,2,\ldots,2n-2$ or $h=1,3,\ldots,2n-1$. \end{itemize} Unfortunately, we have no clear explanation for these non-trivial choices of the holonomy sectors. We hope we can return to this problem in future work. \section{The sign factor to the lens space index and the 3d index} In this section we discuss the effect of the sign factor $f(h)$ derived in Subsection \ref{subsec:pairing} to the 4d lens space index \cite{Benini:2011nc} and 3d superconformal index \cite{Kim:2009wb,Imamura:2011su,Kapustin:2011jm}. Since the orbifold partition function is related to the lens space index as well as the 3d superconformal index through dimensional reductions the same or similar sign factor should appear in those indices. Let us first summarize the lens space index and its reductions, and then, we discuss the sign factors. The lens space index can be obtained by the orbifold projection of the 4d index on ${\bm S}^3 \times {\bm S}^1$. The projection is performed along the Hopf fiber direction of the ${\bm S}^3$ of the ${\bm S}^3 \times {\bm S}^1$, and it is realized by leaving the modes compatible with the identification (\ref{zndef}). Since the rotation along the Hopf fiber direction is characterized by $(J_1 -J_2)$ the projection is realized by inserting the operator \begin{align} \frac{1}{n} \sum_{m=0}^{n-1} e^{2\pi i \frac{m}{n} (J_1-J_2)} \label{orbop} \end{align} to the index (\ref{4dindex}). Namely, the lens space index $I_n$ is \begin{align} I_n (p_1,p_2,z_a) = \frac{1}{n} \sum_{m=0}^{n-1} I(e^{\frac{2 \pi i}{n}m}p_1,e^{-\frac{2 \pi i}{n}m}p_2,e^{-\frac{2 \pi i}{n}h_a}z_a) , \end{align} where we introduced the holonomies for global symmetries $h_a$. Let us focus on a theory consisting of a single chiral multiplet $\Phi$ whose charge for a $U(1)$ global symmetry is 1. The 4d index on ${\bm S}^3\times {\bm S}^1$ defined by (\ref{4dindex}) for the theory can be rewritten as follows \cite{Romelsberger:2007ec}. \begin{align} I_\Phi (p_1,p_2,z) &= \mathop{\rm Pexp}\nolimits' \left[ I_\Phi^\mathrm{sp} \right] \\ I_\Phi^\mathrm{sp} (p_1,p_2,z) &= \sum_{i,j=0}^\infty \left(z p_1^i p_2^j -z^{-1} p_1^{j+1} p_2^{i+1} \right) \label{spphi} \\ &= \frac{z}{1-p_1 p_2} \left(\frac{1}{1-p_1} +\frac{p_2}{1-p_2}\right) -\frac{p_1 p_2 z^{-1}}{1-p_1 p_2} \left(\frac{p_1}{1-p_1} +\frac{1}{1-p_2}\right) , \end{align} where $z$ is a fugacity for the global symmetry, and the prime of the plethystic exponential denotes that it includes the zero point contributions; $\mathop{\rm Pexp}\nolimits' [x] = e^{x/2} \mathop{\rm Pexp}\nolimits[x]$. In this form the insertion of the operator (\ref{orbop}) is equivalent to leave the modes invariant under the condition \begin{align} i - j = h \mod n \label{condforind} \end{align} in (\ref{spphi}). Then, the lens space index for this theory can be written as follows. \begin{align} I_{n,\Phi} (z,h) &= \mathop{\rm Pexp}\nolimits'[I_{n,\Phi}^\mathrm{sp}(z,h)] \\ I_{n,\Phi}^\mathrm{sp}(z,h) &= \frac{z}{1-p_1 p_2} \left(\frac{p_1^{[h]}}{1-p_1^{n}} +\frac{p_2^{n-[h]}}{1-p_2^{n}}\right) -\frac{p_1 p_2 z^{-1}}{1-p_1 p_2} \left(\frac{p_1^{n-[h]}}{1-p_1^{n}} +\frac{p_2^{[h]}}{1-p_2^{n}}\right) . \end{align} Let us now consider the effect of the sign factor (\ref{rhoh}). The orbifold partition function is derived in the small radius limit (\ref{smallradius}) of the lens space index. Since the sign factor is independent of the radius $\beta$ it can be uplifted to the lens space index, and we define the improved index as \begin{equation} I_n^\mathrm{imp} (z,h)=\left( \prod_I\rho(h_I)\right) I_n (z,h) . \label{IandI0} \end{equation} For the theory considered above the improved lens space index is written as \begin{align} I_{n,\Phi}^\mathrm{imp} (z,h) &= e^{-\frac{i \pi}{2} h(1-h)} \mathop{\rm Pexp}\nolimits'[I_{n,\Phi}^\mathrm{imp, sp}(z,h)] \\ I_{n,\Phi}^\mathrm{imp, sp} (z,h) &= \frac{z}{1-p_1 p_2} \left(\frac{p_1^{h}}{1-p_1^{n}} +\frac{p_2^{n-h}}{1-p_2^{n}}\right) -\frac{p_1 p_2 z^{-1}}{1-p_1 p_2} \left(\frac{p_1^{n-h}}{1-p_1^{n}} +\frac{p_2^{h}}{1-p_2^{n}}\right) . \end{align} Note that $[h]$ in $I_{n,\Phi}$ is now replaced by $h$. Although both $I_{n,\Phi}$ and $\rho(h)$ include $[h]$ the combination of those (\ref{IandI0}) is analytic in terms of $h$. In order to illustrate the effect of the phase factor we obtained to the 3d index, we first describe the 3d index for the tetrahedron theory, which is written as follows \begin{align} I^\mathrm{3d}_{\Delta} &= (-1)^{-\frac{1}{2}|h|} \left( (-q^{\frac{1}{2}})^{\frac{|h|-h}{2} } z^{-\frac{|h|-h}{2} } \right) \mathop{\rm Pexp}\nolimits \left[ \frac{q^{\frac{|h|}{2} } z}{1-q} -\frac{q^{1+\frac{|h|}{2} } z^{-1}}{1-q} \right] . \label{dimofte} \end{align} Later, it is noticed in \cite{Dimofte:2011py} that the 3d index can be written in an analytic form by adding a simple phase factor: \begin{align} I^\mathrm{3d, imp}_{\Delta} &= i^{|h| } I^\mathrm{3d}_{\Delta} = \mathop{\rm Pexp}\nolimits \left[ \frac{q^{\frac{h}{2} } z}{1-q} -\frac{q^{1+\frac{h}{2} } z^{-1}}{1-q} \right] . \end{align} The phase factor is needed for the 3d index to be factorized. As is pointed out in \cite{Benini:2011nc}, the 3d index is obtained from the 4d lens space index by taking $n\rightarrow\infty$ limit. When we take this limit $h$ is kept finite and is identified with the magnetic flux in ${\bm S}^2\times {\bm S}^1$. In this limit the sign factor (\ref{rhoh}) can be rewritten as \begin{equation} \rho(h)= e^{\frac{\pi i}{2}(|h|-h^2)} . \label{rohlargeh} \end{equation} The contribution of the classical factor in (\ref{cscont}) in the large $n$ limit becomes \begin{equation} e^{-\pi ik_{ab}h_ah_b} . \end{equation} As the tetrahedron theory has Chern-Simons term with level $-1/2$ its contribution combined with the extra sign factor (\ref{rohlargeh}) is \begin{equation} e^{\frac{\pi i}{2}h^2}\times e^{\frac{\pi i}{2}(|h|-h^2)} = i^{|h|} . \label{3dphase} \end{equation} This is precisely the factor introduced in (\ref{dimofte}) to improve the 3d index. Note that the improved 3d index is analytic in terms of $h$, which is inherited from the analytic structure of the improved 4d index. \section{Conclusions} We focused on the sign of the orbifold partition function. We proposed a formula that systematically determine the relative signs among holonomy sectors. We use the factorization to the holomorphic blocks as a criterion to determine correct signs. For non-gauge theories we proved that the partition function with signs determined by the formula is correctly factorized into the holomorphic blocks. In the case of gauge theories, as a simple example, we considered two theories: SQED with $N_f=1$ and $su(2)$ gauge theory with an adjoint flavor. For the former we confirmed the factorization of the orbifold partition function. In the case of the $su(2)$ theory, we have not fully understood the summation over the holonomy sectors. We guessed which holonomy sector contributes to the partition function with the help of the duality to a non-gauge theory, which is known as Jafferis-Yin duality. We found that we have to choose an appropriate subset of the holonomy sectors to obtain the partition function consistent to the duality. The formula for the sign factor gives the correct relative phases for the contributing sectors. Therefore, up to the subtlety for the choice of the holonomy sector, which is probably related to the global aspects of the gauge bundle on the orbifold, the formula seems to give appropriate signs. We also discussed the sign factors in the lens space index of the 4d theories, and the index for 3d theories. The formulae for these indices have been known, and contain non-trivial sign factors, which are often implicit in the literature. We showed that they are closely related to the sign factor for the orbifold partition function. \section*{Acknowledgments} We would like to thank S.~Terashima for useful information and S.~Kim for valuable discussion. Y.~I. is partially supported by Grant-in-Aid for Scientific Research (C) (No.24540260), Ministry of Education, Science and Culture, Japan. H.~M. acknowledges the financial support from the Center of Excellence Program by MEXT, Japan through the ``Nanoscience and Quantum Physics'' Project of the Tokyo Institute of Technology. The work of D.Y. is supported in part by the National Research Foundation of Korea Grant NRF-2012R1A2A2A02046739, the Global Center of Excellence Program by MEXT Japan through the ``Nanoscience and Quantum Physics'' Project of the Tokyo Institute of Technology, and Perimeter Institute for Theoretical Physics. Research at Perimeter Institute is supported by the Government of Canada through Industry Canada and by the Province of Ontario through the Ministry of Economic Development \& Innovation.
\section{Introduction} Fast particles are present in magnetic fusion devices due to external plasma heating and eventually due to fusion born $\alpha$-particles. It is necessary that these super-thermal particles are well confined while they transfer their energy to the background plasma. Fast particle populations can interact with global electromagnetic waves, leading to the growth of MHD-like and kinetic instabilities -- e.g.\ Toroidicity-induced Eigenmodes (TAE) \cite{Cheng85, Cheng86}, Reversed Shear Alfv{\'e}n Eigenmodes (RSAE) \cite{Berk01,Breizman03} or Beta-induced Alfv{\'e}n Eigenmodes (BAE) \cite{Heidbrink93, Turnbull93}. This in turn redistributes the fast ions, enhances fast ion losses and affects the confinement. As a consequence, fusion power and fuel is reduced and the machine wall may suffer damage.\\ \textsc{Asdex} Upgrade (\textsc{Aug})\cite{Gruber84,GarciaRev09} and other machines of similar size (e.g.\ \textsc{DIII-D} \cite{zeeland11}), have the advantage to directly measure the fast ion losses due to core-localized MHD activity, since the particle orbits are relatively large compared to the machine size. The fast ion loss detector (\textsc{Fild}) at \textsc{Asdex} Upgrade provides energy and pitch angle resolved measurements of fast ion losses \cite{GarciaRev09,Garcia10}.\\ To find out about the loss mechanisms, loss features such as phase space pattern and ejection frequency are analyzed. Both can help to learn about whether or not the losses are caused by mode-particle (double) resonance (\emph{resonant losses}), phase space stochastization (\emph{diffusive losses}) \cite{Sigmar92} or none of both. Therefore, the losses are categorized: all particles that are unconfined solely due to their large orbit width -- caused by a high birth energy -- or a birth position radially far outside, are called \emph{prompt losses}. Their ejection frequency is not related to any of the mode frequencies, and the losses are therefore inherently \emph{incoherent}. Losses that are caused due to wave-particle interaction can be both, either \emph{coherent losses}, or incoherent as well. They are coherent, if they show a correlation with one of the mode frequencies or the beat frequency of different modes, but incoherent if they are ejected e.g.\ due to phase space stochastization. To learn about the role of the modes in the ejection mechanisms, the losses are further studied with respect to the number and type of modes present in the plasma and the $q$ profile\ (as this determines particle orbit widths and also mode growth). Eventually, this work aims to reproduce quantitatively the pitch and energy distribution and qualitatively the decomposition into \emph{coherent} and \emph{incoherent} losses of the of the measured fast ion losses \cite{Garcia10} at the different times in the \textsc{aug} discharge \#23824. From the code point of view, the study can be regarded as a code validation against the experiment, which is possible in small machines with large particle orbits (due to \textsc{Icrh}) such as \textsc{Asdex} Upgrade, where a large number of losses appears at the first wall and the \textsc{fild}.\\ This paper is organized as follows: \sref{sec:experiment} presents the experimental settings the numerical simulations are based on, as well as experimental loss observation. In \sref{sec:hagis}, the \textsc{Hagis} code is shortly introduced, especially the importance of the vacuum extension \cite{mwb_phd} is demonstrated. Next (\sref{sec:resplot}), the resonant and loss areas in phase space are given for the scenario that is studied. In \sref{sec:numloss}, the advance towards more realistic simulations is documented, followed by the presentation of the final results. \section{Experimental Observation}\label{sec:experiment} The \textsc{Icrh} minority-heated \textsc{aug} discharge \#23824 is especially suitable for the investigation of fast particle-wave interaction and losses, as it is characterized by an inverted $q$ profile\ after the current ramp-up phase, which relaxes to a monotonic profile as $q$ decreases (see \fref{q23824}). As particle confinement depends largely on the $q$ profile\ due to the proportionality between orbit width and $q$ value, the inverted $q$ profile\ with its higher absolute $q$ values in the core is expected to impact the amount of losses and also the wave-particle interaction.\\ The simulations are based on two different MHD equilibria -- at $t=1.16$\ s and $t=1.51$\ s. The total plasma current is constant between both time points ($I=800$\ kA), only the current density penetrates inwards. Consequentially the inverted $q$ profile\ at the earlier time point changes to a monotonic $q$ profile\ at the later time point. \section{The Hagis Code}\label{sec:hagis} The numerical investigations presented in this work are performed with the \textsc{Hagis} code \cite{Pinches98, sip_phd}, a nonlinear, driftkinetic, perturbative Particle-in-Cell code (current release 12.05). \textsc{Hagis} models the interaction between a distribution of energetic (fast) particles and a set of Alfv{\'e}n eigenmodes. It calculates the linear growth rates as well as the nonlinear behavior of the mode amplitudes and the fast ion distribution function that are determined by kinetic wave-particle nonlinearities. It is fully updated to work with MHD equilibria given by the recent \textsc{Helena} \cite{Huysmans91} version.\\ The plasma equilibrium and in particular the $q$ profile\ is determined by Alfv{\'e}n spectroscopy of the RSAEs: magnetic pick-up coil data and soft X-ray emission measurements \cite{Igochine03} are used to determine the $q$ profile\ minimum value and location. The plasma equilibrium for \textsc{Hagis} is based on the \textsc{Cliste} code \cite{Carthy12}, then transformed via \textsc{Helena} to straight field line coordinates and to Boozer coordinates via \textsc{Hagis}. \begin{minipage}{1\textwidth} \begin{minipage}{0.6\textwidth} \begin{figure}[H] \centering \includegraphics[width=0.8\textwidth, height=5.5cm]{figure1-eps-converted-to.pdf} \caption{\itshape $q$ profile\ of \textsc{aug} discharge \# 23824 at different times, obtained from \textsc{Cliste} calculation constraint by Alfv{\'e}n spectroscopy measurements as described in detail in ref.\ \cite{Lauber09}. The black solid line shows the $q$ profile\ at $t=1.16$\ s, whereas the blue dashed line refers to $t=1.51$\ s: the $q$ profile s differ in the shape (the earlier one is inverted, the later one monotonic), but also in the absolute values (the inverted $q$ profile\ has higher absolute $q$ values). Note: $q < 0$ due to \textsc{aug}'s helicity of the magnetic field. } \label{q23824} \end{figure} \end{minipage} \hfill \begin{minipage}{0.35\textwidth} \begin{figure}[H] \hspace{1cm}\includegraphics[width=0.9\textwidth, height=7cm]{figure2-eps-converted-to.pdf} \caption{\itshape Hagis equilibrium (flux surfaces in black) with vacuum region (red), representative banana orbit (blue) and \textsc{Fild} position (black square).} \label{hagis_equilibrium} \end{figure} \vspace{0.5cm} \end{minipage} \end{minipage} \newpage Recently, \textsc{Hagis} has been extended to the vacuum region (see \fref{hagis_equilibrium}), making it possible to track particles that re-enter the plasma \cite{mwb_phd}, and to investigate losses including finite Larmor radius effects. To allow for the second point within the frame of a drift kinetic ansatz, a tube around a particle's guiding center is assumed, whenever it approaches the wall. Where this tube intersects the wall, the particle is lost. For scenarios with many lower energetic losses, as they appear e.g.\ in all inverted $q$ profile\ simulations presented in this work, the use of the vacuum extension is crucial. \Fref{run0502+508_Eloss} visualizes the difference in the losses' energy spectra obtained with (black curve) and without (red curve) the vacuum extension -- if simulating without it, the lower energetic losses are massively overestimated. The overestimation results from to the fact, that particles are considered lost, as soon as they cross the separatrix. However, particles with very small gyroradii (i.e.\ low perpendicular energies), often re-enter the plasma and are not lost. The overestimation of lost markers if simulating without the vacuum extension is over 130\% in the given example. \textsc{Icrh}-generated fast particles are characterized by relatively low pitches $\lambda = v_\|/v$, thus, they have low perpendicular energies only if their total energy is in a lower range. These particles have small gyroradii, and therefore a higher probability to re-enter the plasma. Further, the distance between the separatrix and the first wall is important, which depends on the given equilibrium in the particular fusion device. The vacuum extension allows particles to travel through the vacuum and checks if the particle hits the machine specific first wall through assuming a `tube' around its guiding center with the radius of the gyroradius. Simulations including the vacuum region are therefore computationally more expensive (roughly between 120\% and 300\% compared to the simulations without the extended version). \begin{figure}[H] \centering \includegraphics[width=0.7\textwidth, height=3cm]{figure3-eps-converted-to.pdf} \caption{\itshape Energy spectra of losses simulated in the inverted $q$ profile\ with the eigenmodes given by \textsc{Ligka}\ using \textsc{Hagis} with its vacuum extension (black curve) and without (red). The amplitudes for these simulations were fixed at $\delta B/B = 5.1\cdot 10^{-3}$. The losses shown appeared in a time interval starting after approximately 10 RSAE wave periods ($t \in [0.2,1.5] \cdot 10^{-3}$\ s) to avoid the prompt losses.} \label{run0502+508_Eloss} \end{figure} \begin{figure}[H] \centering \subfigure[\itshape at $t = 1.16$\ s -- inverted $q$ profile]{\includegraphics[width=0.45\textwidth,height=5.5cm]{figure4a-eps-converted-to.pdf}}\hspace{1cm} \subfigure[\itshape at $t = 1.51$\ s -- monotonic $q$ profile]{\includegraphics[width=0.45\textwidth,height=5.5cm]{figure4b-eps-converted-to.pdf}} \caption{\itshape Resonance plots produced by the extended version of \textsc{Hagis} for the equilibria of \textsc{aug} discharge \#23824 at two different time points with different $q$ profile s. The vertical axis gives the radial position of the particle's bounce point (where pitch $\lambda =0$), the horizontal axis the energy. Pink indicates resonant areas, with respect to the main MHD modes as diagnosed experimentally ($n=5$ TAE of $120$\ kHz, $n=4$ RSAE of $55$\ kHz at $t = 1.16$\ s and $n=5$ TAE of $160$\ kHz, $n=4$ BAE of $70$\ kHz at $t = 1.51$\ s). Particles in the blue regions are lost. The loss region of the right picture is plotted superimposed in the left image to point out the difference.} \label{resplot_AUG-23824} \end{figure} \section{Resonances and Loss Regions}\label{sec:resplot} The \textsc{Icrh} minority-heating produces mainly trapped particles with their bounce points within the heating region that is at constant major radius above and below the magnetic axis (see \fref{loading-weighting}b). Trapped particles with a bounce frequency $\omega_{\mathrm{b}}$ and toroidal precession frequency $\omega_{\mathrm{tp}}$ interact with MHD modes of a certain frequency $\omega$, if the \emph{resonance condition}, $\omega - n\omega_{\mathrm{tp}}-p\omega_{\mathrm{b}} \approx 0$ is fulfilled, where $n$ is the toroidal mode harmonics and $p$ the particles' bounce harmonics \cite{porcelli94}. Plotting the left side of the resonance condition as a contour plot over the particle energy $E$ and the $z$ coordinate of the particle's bounce point over the magnetic axis gives the \emph{resonance plot} \cite{Pinches06} shown in \fref{resplot_AUG-23824}. Since at the bounce point, the pitch is $\lambda \equiv v_\|/v = 0$, the second dimension in energy space is kept fixed. The normalized $z$ coordinate can be translated to the radial coordinate $s$, the square root of the normalized poloidal flux: $s=(\psi_{\mathrm{pol}}/\psi_{\mathrm{pol, edge}})^{1/2}$. The darker blue areas give the \emph{loss regions}: in the respective phase space particles are unconfined.\\ The loss region is much broader in the earlier case, with the inverted $q$ profile. The reason for the decrease of the loss region is the decreasing $q$ in the plasma center region, leading to a decrease of orbit width (due to width $\propto q$) and therefore better particle confinement. \section{Fast Particle Losses in Numerical Simulations vs.\ Experimental Measurements}\label{sec:numloss} \subsection{Experimental Loss Measurements} For the \textsc{aug} discharge \#23824, the experiment gives a loss signal over time as depicted in figure 4 of ref.\ \cite{Garcia10}. The upper part shows the spectrogram of that data, which allows to identify the mode frequencies of the Alfv{\'e}n Eigenmodes. One can distinguish a large signal with a vast majority of incoherent losses at earlier time points, whereas after $t = 1.4$\ s the signal is very low, consisting of coherent losses only. A Fourier analysis revealed two types of losses: \emph{coherent} losses, i.e.\ losses ejected at a frequency correlated with the MHD mode frequency on and \emph{incoherent} losses, that do not show such a correlation. \subsection{Realistic Simulation Conditions}\label{sec:simcond} The fast particle population is simulated as protons (concerning mass and charge), according to how it is created by \textsc{Icrh} minority heating in a deuterium plasma \cite{Garcia10}. The volume averaged fast particle beta is set to $\beta_{\mathrm{fp}}$\ $\approx 0.05$\%. This can be estimated from the construction of the MHD equilibrium and different kinetic plasma profile measurements. However, the \textit{volume averaged} $\beta_{\mathrm{fp}}$\ does not alone determine the radial gradient of the distribution function, which governs the mode drive. For the most realistic modeling in the following sections, the $\beta_{\mathrm{fp}}$\ value given above is reduced to $\beta_{\mathrm{fp}}$\ $= 0.02$\% to obtain mode amplitude saturation levels comparable to the experimental values. However, the larger $\beta_{\mathrm{fp}}$\ value leads to a faster mode growing and slightly overestimated amplitude saturation levels. Since we are not interested in the total amount of losses but rather in their energy and pitch distribution as well as their nonlinear behavior (which is the same for both $\beta_{\mathrm{fp}}$), in the indicated simulations, the larger $\beta_{\mathrm{fp}}$\ is chosen to save computational costs. With either beta value, the nonlinear regime is steady-state without frequency chirping, since there is no damping involved.\\ To be able to simulate the interaction of these \textsc{Icrh}-generated fast ions with Alfv{\'e}nic waves, the distribution function in \textsc{hagis} was adapted (see also ref.\ \cite{mirjam_phd}): especially the anisotropy in the pitch $\lambda$ and orbit topology is taken into account: \textsc{Icrh}-generated fast ions are characterized by low pitches and are mainly trapped (banana orbits). The major challenges are the following: first, and different from \textsc{Nbi}-heated plasmas, the fast ion distribution function in \textsc{Icrh}-heated plasmas is almost unknown experimentally (especially for this discharge). Second, the strong anisotropies of the distribution function are in coordinates that are not constants of the motion (pitch $\lambda$ and poloidal angle $\vartheta$ -- coordinates of velocity and real space that change along the particle orbit), making it difficult to evolve such a distribution function in \textsc{Hagis}. Therefore, in the new implementation, the weighting is not performed in $\lambda$ and $\theta$ anymore, but in coordinates that are constants of the motion. The quantity $\Lambda$ fulfills this condition: it is connected to the pitch $\lambda$, but in contrast to that, is not subject to changes along the orbit: \begin{equation}\label{bigLambda} \Lambda ~:=~ \frac{\mu~B_\mathrm{mag}}{E_\mathrm{tot}} ~=~ \frac{B_\mathrm{mag}}{B} \left(1-\lambda^2\right), \end{equation} where $B_\mathrm{mag}$ is the magnetic field at the magnetic axis, introduced as normalization, $B$ is the magnetic field at the respective position $B(s,\theta)$ and $\lambda=\lambda(s,\theta)$ the pitch at this position. Due to the necessity of transforming $\lambda$ coordinate into the constant of the motion $\Lambda$, one obtains one velocity space coordinate dependent on the spacial coordinates: $\Lambda(s,\theta) = B_\mathrm{mag}/B(s,\theta)(1-\lambda^2)$. The distribution in $\Lambda$ is determined by marker loading in poloidal angle ($\theta$) and pitch ($\lambda$) space. Unless otherwise noted, markers start at pitches and poloidal angles of \begin{equation}\label{formula:lampollim} \lambda \in [0 \pm 0.2] \mathrm{~~~~~~~and~~~~~~~} \theta \in [\pm 90^{o} \pm 17.2^{o}]. \end{equation} A poloidal cut of the two initial cones, where the \textsc{Icrh}-generated fast particles are loaded is shown in \fref{loading-weighting}b. A reasonable $\Lambda$ weighting function is the modified Gaussian \begin{equation}\label{formula:Lambdis} f(\Lambda) = \exp{\{((\Lambda - \Lambda_0)/\Delta \Lambda^{0.2})^2\}}, \end{equation} where $\Lambda_0$ and $\Delta \Lambda$ result from $\lambda_0$ and $\Delta \lambda$ as well as from the (unperturbed) magnetic field $B(s,\theta)$ (according to \eref{bigLambda}) and are thus position-dependent. \begin{figure}[H] \centering \subfigure[\itshape radial direction $s$]{\includegraphics[width=0.4\textwidth,height=3.5cm]{figure5a-eps-converted-to.pdf} \hspace{2cm}\subfigure[\itshape loading along $s$ and $\theta$]{\includegraphics[width=0.3\textwidth,height=4.5cm]{figure5b-eps-converted-to.pdf}}\hspace{0.75cm} \caption{\itshape Regions where markers are loaded (green) in the different phase space directions and the distribution function for the unperturbed part $f_0$ (``weighting'', blue, only in (a)).} \label{loading-weighting} \end{figure} The chosen $\Lambda$ weighting results in a relatively broad $\Lambda$-distribution, which is realistic when considering also pitch angle scattering, but too broad to represent the instantaneous effect of the \textsc{Icrh} heating source. Therefore, the pattern of the losses appearing at the very beginning (prompt losses) is broader than what would be realistic. To investigate this pattern, additional simulations are carried out with marker loading in \begin{equation}\label{formula:lampollim2} \lambda \in [0 \pm 0.05] \mathrm{~~~~~~~and~~~~~~~} \theta \in [\pm 90^{o} \pm 5^{o}]. \end{equation} In radial position space, markers were loaded within $s_2 \leq 0.7$ (unless otherwise noted). The weighting function is implemented as constant for $s < s_0$ and drops down according to a Fermi-like potential law for $s \geq s_0$, with $s_0 = 0.3$ (unless otherwise noted), as shown in \fref{loading-weighting}a: \begin{equation}\label{formula:sdis} f(s) = \Bigg\{ \begin{array}{cc} \mathrm{const.} ~~& \mathrm{for}~~ s < s_0\\ (1-(s-s_0)^2)^a & \mathrm{for}~~ s \ge s_0, \end{array} \end{equation} Concerning the energy distribution $f(E)$, a slowing-down function is used, \begin{equation}\label{formula:Edis} \nonumber f(E) = \frac{1}{E^{3/2}+E^{3/2}_{\mathrm{c}}}\mathrm{erfc}\left(\frac{E-E_0}{\Delta E}\right), \end{equation} with $E_\mathrm{c}=19.34$\ keV, $\Delta E=149.9$\ keV and $E_0=1.0$\ MeV, although a slowing-down function is not considered the most adequate distribution function for \textsc{Icrh} generated fast particles. However, since it is almost unknown experimentally for this discharge, the \textsc{Toric-ssfpql} code \cite{Bilato12} was consulted. The data in the highest energy range calculates by \textsc{Toric-ssfpql} (red line in \fref{energy-disb}), which corresponds to the medium and most relevant energy range in the presented investigations, can justify the use of a slowing-down function in the \textsc{Hagis} simulations (green line in \fref{energy-disb}). \begin{figure}[H] \centering \includegraphics[width=0.8\textwidth,height=3.5cm]{figure6-eps-converted-to.pdf} \caption{\itshape Energy distribution function in \textsc{Hagis} (green) and as given by the \textsc{Toric-ssfpql} code (red, which focuses on the lower energy range and reaches only up to 430\ keV). The red dashed lines give the slowing-down fit to the \textsc{Toric-ssfpql} data.} \label{energy-disb} \end{figure} Still, $f$ is independent in each dimension, $f= f_s(s) f_E(E) f_\Lambda(\Lambda)$. A comparison with the distribution function as calculated by the \textsc{Toric-ssfpql} code confirms this as a valid approximation for the treatment in $\Lambda$. Further, it is comparable to previous approaches, as reported in ref.\ \cite{Zonca00}. However, concerning radial-energy space, further improvements towards more realistic distribution functions, which are non-separable in the two dimensions ($f(s,E)$), are currently under investigation. An analytical model is given in ref.\ \cite{Troia12}, and will be implemented into \textsc{Hagis}, once its parameters are optimized numerically (by \textsc{Toric-ssfpql}) and validated by the comparison with experimental findings. A sensitivity investigation (presented in the following) confirms robustness of non-prompt losses against changes in the presently used distribution function:\\ A first scan is performed with three different radial distributions of decreasing steepness while the radial extent broadens -- the parameter $s_0$ in the weighting function \eref{formula:sdis} takes the values 0.15, 0.25, 0.35 (for all three, it is $a=5$) and the marker loading reaches out until $s_2 =0.6,~ 0.7$ and $0.8$ respectively. The second scan is carried out for three poloidal loadings according to $\Delta \theta = \pm 17.2^o, \Delta \lambda = \pm 0.2$ and $\Delta \theta = \pm 11.5^o, \Delta \lambda =\pm 0.2$ as well as $\Delta \theta = \pm 2.5^o, \Delta \lambda =\pm 0.05$. In \fref{run0447+452+456_pEloss}, the boundaries of the loss pattern are shown, as obtained for the scan in the radial distribution function. One can see, that the prompt losses (a) are relatively sensitive to the distribution function. However, they cannot be determined quantitatively within the used model anyhow. The reason lies inherently in the code model, which does not calculate the slowing-down process of the (unperturbed) fast particle distribution function. Since the simulated time is short compared to the slowing-down time, a stationary situation is considered, where heating and collisional dissipation balances each other. The non-prompt losses (b) in contrast, are quite robust against minor changes in the radial and poloidal distribution function. A similar result is obtained for the poloidal scan (see \fref{run0447+471+472_pEloss}). \begin{figure}[H] \centering \subfigure[\itshape Prompt losses ($t < 10^{-4}$)]{\includegraphics[width=0.43\textwidth,height=5.cm]{figure6a-eps-converted-to.pdf} \subfigure[\itshape Resonant/diffusive losses]{\includegraphics[width=0.43\textwidth,height=5.cm]{figure6b-eps-converted-to.pdf} \caption{\itshape Losses at first wall in pitch angle-energy ($\lambda^o$-$E$) space for three double-mode simulations in the inverted $q$ profile\ (modes as explained below and shown in \fref{ligka-input}a, $\beta_{\mathrm{fp}}$=0.05\%). The simulations differ in the radial distribution function: the color code gives the prompt (a) and later (b) losses for the intermediate distribution function, the gray lines give the boundary of the pattern, when using the steeper radial distribution function; the white line gives the respective result for the flatter distribution function.} \label{run0447+452+456_pEloss} \end{figure} Having adapted the simulation conditions towards more realistic fast particle distribution functions, the input perturbation data is improved (see also ref.\ \cite{mirjam_phd}). Based on the resulting MHD equilibrium (originating from \textsc{Cliste} calculations \cite{Carthy12, Diarmuid+Schneider-priv}), the radial structure and frequency of the perturbation is calculated numerically with the linear gyrokinetic eigenvalue solver \textsc{Ligka} \cite{Lauber07}. \textsc{Hagis} is extended to read the real as well as the imaginary part of the electromagnetic perturbation. In general, the magnetic ($\tilde{\Psi} = (\nabla A)$) as well as the electric ($\tilde{\Phi}$) part can be taken into account separately. However, if the damping is small, as in this case, it can be assumed that $\tilde{\Psi} \propto \tilde{\Phi}$.\\ During \textsc{aug} discharge \# 23824, two dominating modes are seen in the experiment \cite{Garcia10,Lauber09} at $t = 1.16$\ s around the radial positions $s \approx 0.3$ and $s \approx 0.5$ with frequencies of $120$\ kHz ($n=4$) and $55$\ kHz ($n=4$) respectively. At the later time point, the frequencies have evolved upwards, resulting in $160$\ kHz ($n=5$) and $70$\ kHz ($n=4$). The higher frequency mode is identified as TAE, the lower as RSAE (=AC) at $t = 1.16$\ s but as BAE at $t = 1.51$\ s \cite{Lauber09}. \begin{figure}[H] \centering \subfigure[\itshape Prompt losses ($t < 10^{-4}$)]{\includegraphics[width=0.43\textwidth,height=5.cm]{figure7a-eps-converted-to.pdf}} \subfigure[\itshape Resonant/diffusive losses]{\includegraphics[width=0.43\textwidth,height=5.cm]{figure7b-eps-converted-to.pdf}} \caption{\itshape Losses at first wall in pitch angle-Energy ($\lambda^o$-$E$) space for two double-mode simulations in the inverted $q$ profile\ (modes as explained later and shown in \fref{ligka-input}a, $\beta_{\mathrm{fp}}$=0.05\%). The simulations differ in the $\Lambda$ distribution function: the color code gives the prompt (a) and later (b) losses for the $\Delta \theta = \pm 17.2^o, \Delta \lambda = \pm 0.2$ marker loading, the white lines give the boundary of the pattern, when using $\Delta \theta = \pm 11.5^o, \Delta \lambda =\pm 0.2$ loading, the gray lines for $\Delta \theta = \pm 2.5^o, \Delta \lambda =\pm 0.05$.} \label{run0447+471+472_pEloss} \end{figure} The following two sections are dedicated to the simulation of fast particle redistribution and losses under realistic simulation conditions with the vacuum extended version of \textsc{Hagis}. Therefore, the original eigenmodes given by \textsc{Ligka} are used with all their poloidal harmonics (unless otherwise noted), with a radial structure as depicted in \fref{ligka-input}a for the MHD equilibrium of \textsc{aug} discharge \#23824 at $t=1.16$\ s (inverted $q$ profile) and in \fref{ligka-input}b for the time point $t=1.51$\ s (monotonic $q$ profile). In the following, the two equilibria will be referred to as ``scenario1.16'' and ``scenario1.51'' respectively. The \textsc{Icrh}-like distribution function is used, and the fast particle beta value was chosen as $\beta_{\mathrm{fp}}$=0.02\%, a quite realistic value, that leads to mode amplitudes comparable to those measured experimentally \cite{Garcia10}, or slightly higher ($\delta B/B \in [5\cdot 10^{-3},2\cdot 10^{-2}]$). \subsection{Internal Transport Study in Two Different $q$ profile\ Equilibria} The mode amplitude evolution in both equilibria are shown in \fref{run0461+462_ampl}: in scenario1.16, one can see clearly the double-resonance effect as described in ref.\ \cite{Schneller12} leading a superimposed oscillation (insert), and also to the TAE (blue curve) exceeding the initially much faster growing RSAE (pink curve) in the late nonlinear phase. The examination below (\fref{run0461+462_Esdf-res}a for scenario1.16) confirms that both modes reach the stochasticity threshold. In scenario1.51, the TAE (green curve) grows much faster than the low frequency mode (in this case a BAE, red curve), but both grow slower and saturate at a lower amplitude compared to scenario1.16. The TAE's stochasticity threshold is reduced by the stochastization due to the BAE. Since the simulated time scale in this case is only one order of magnitude below the slowing-down time, the effect of energy dissipation becomes slightly visible with the TAE amplitude decreasing at the end of the simulation. Beyond this time point, the present model is not valid any more. One would have to extend it for a fast particle source term and take damping mechanisms into account. \begin{figure}[H] \centering \subfigure[\itshape at $t = 1.16$~s: 55~kHz RSAE (gray) and 120~kHz TAE]{\includegraphics[width=0.49\textwidth, height=4cm]{figure8a-eps-converted-to.pdf}} \subfigure[\itshape at $t = 1.51$~s: 70~kHz BAE (gray) and 160~kHz TAE]{\includegraphics[width=0.49\textwidth, height=4cm]{figure8b-eps-converted-to.pdf}} \caption{\itshape The radial structure including poloidal harmonics with $m=4...10$ (a) and $m=4..11$ (b) (from black to blue/violet) of the $n=4$ TAE and the $m=4,~n=4$ RSAE of (a)/BAE (b) (gray) as calculated by \textsc{Ligka} for the \textsc{aug} discharge \#23824 equilibrium at two different time points. The given perturbation amplitude refers to the real (solid lines) and the imaginary (dashed lines) part of the electric perturbation potential $\tilde{\Phi}$. The values are normalized, such that unity corresponds to the initial value of $\delta B/B = 10^{-7}$.} \label{ligka-input} \end{figure} To understand the mode-particle interaction in the different stages of the mode evolution, it is helpful to look at the processes in phase space. In \fref{run0461+462_Esdf-res}b, redistribution in $E$-$s$ space is shown during the mode's resonant phase for scenario1.51. In this phase, the redistribution in phase space takes place along the resonance lines, from higher energies and lower radial positions to lower energies further outside. As the modes in scenario1.51 stay longer in the resonant phase, the redistribution pattern is better visible in this case, but can be seen in scenario1.16 as well. \begin{figure}[H] \centering \includegraphics[width=0.5\textwidth,height=5cm]{figure9-eps-converted-to.pdf} \caption{\itshape Amplitude evolution in two double-mode simulations with different MHD equilibria of the \textsc{aug} discharge \#23824 ($\beta_{\mathrm{fp}}$=0.02\%): pink and blue: $n=4$ RSAE ($55$\ kHz) and $n=4$ TAE ($120$\ kHz) (radial structure see \fref{ligka-input}a), simulated in the equilibrium at $t= 1.16$\ s, red and green: $n=4$ BAE ($70$\ kHz) and $n=5$ TAE ($160$\ kHz) (see \fref{ligka-input}b) in the equilibrium at $t= 1.51$\ s. The marker are loaded according to the \textsc{Icrh}-like distribution function described in \sref{sec:simcond}.} \label{run0461+462_ampl} \end{figure} When the modes, or at least one of both, reach the stochastization level, a massive radial gradient depletion takes place over the whole energy space. No resonance pattern is visible any more, as visualized in \fref{run0461+462_Esdf-res}a (for scenario1.16). This happens in both scenarios, however much later in scenario1.51 around $t \approx 1.5 \cdot 10^{-2}$\ s whereas at $t \approx 1.5\cdot 10^{-3}$\ s in scenario1.16. In both cases, at lower energies ($E \in [50,200]$\ keV), the resonance regions are still slightly visible in the redistribution pattern. However, only in scenario1.16 does the redistribution (along the $p = 0,~1$ resonance lines of both modes) cross the loss boundary. The redistribution takes place independently on the energy, but once the radial gradient is depleted in the energy region of $E \in [400,600]$\ keV, the redistribution is radially broadened at lower energies. In \sref{sec-poloidalharmonics}, it will be shown that this broad radial redistribution is caused by the outer poloidal harmonics of the TAE. \subsection{Fast Particle Losses in Two Different $q$ profile\ Equilibria} The redistribution plots already indicate fewer losses in scenario1.51, due to the smaller loss area, and the weaker redistribution especially in the lower energy range. In fact, scenario1.16 gives much more losses -- prompt (i.e.\ appearing before any influence of the mode) as well as losses in the later phase of the simulation (compare \fref{run0461+462_tEloss}a and b). In scenario1.51, there are no losses with an energy below $300$\ keV at all. This is in accordance with the redistribution pattern of this case, giving no redistribution across the loss boundary for $E < 300$\ keV, due to the smaller loss region. The prompt losses appear in both cases in the higher energy range. They reach down to $600$\ keV (scenario1.16) and $750$\ keV (scenario1.51).\\ As the resonant phase is not distinguishable very clearly in scenario1.16, it is difficult to find resonant losses at very distinct energies. However, the losses in this phase have energies that match with the resonant phase space redistribution across the loss boundary, although not as clearly as in scenario1.51 (see \fref{run0461+462_Esdf-res}b). In scenario1.16, the resonances are at energies of $E \in [450,700]$\ keV and $\in [750,950]$\ keV. In scenario1.51, the distinct energies result in losses around $400$\ keV, $550$\ keV, $700$\ keV and $850$\ keV (see \fref{run0461+462_tEloss}b). \begin{figure}[H] \centering \subfigure[\itshape scenario1.16, $t_2=1.5\cdot10^{-2}$\ s.]{\includegraphics[width=0.4\textwidth, height=5.5cm]{figure10a-eps-converted-to.pdf}} \hspace{0.7cm}\subfigure[\itshape scenario1.51, $t_1=2.3\cdot10^{-3}$\ s.]{\includegraphics[width=0.4\textwidth, height=5.5cm]{figure10b-eps-converted-to.pdf}} \caption{\itshape Redistribution (in terms of fast particle pressure $E \delta f$) in energy and radial space at two different stages in the simulation. Red indicates particle accumulation, blue means particles move away. In b, the pink lines give the resonance lines for trapped particles, the blue areas the loss region (both according to \fref{resplot_AUG-23824}b). The horizontal lines denote the modes' radial positions. One can see clearly the redistribution caused by the wave-particle resonance along the resonance lines in the mode regions.} \label{run0461+462_Esdf-res} \end{figure} \begin{figure}[H] \centering \subfigure[\itshape scenario1.16]{\includegraphics[width=0.48\textwidth,height=5.5cm]{figure11a-eps-converted-to.pdf}} \hspace{0.3cm}\subfigure[\itshape scenario1.51]{\includegraphics[width=0.48\textwidth,height=5.5cm]{figure11b-eps-converted-to.pdf}} \caption{\itshape Temporal evolution of losses appearance at the first wall in energy space for the two different scenarios. The white areas indicate the prompt losses as carried out in an extra simulation with a thinner $\Lambda$ distribution function \eref{formula:lampollim2}, as explained before. The gray lines give the mode amplitude evolution as shown in detail in \fref{run0461+462_ampl}.} \label{run0461+462_tEloss} \end{figure} When stochastization sets in, a higher amount of energetic losses appears, due to the broad redistribution across the loss boundary over a large energy range. In scenario1.16, the losses reach down to low energies. The cause is a phase space channeling effect within the redistribution (according to the domino effect proposed theoretically in \cite{Berk95-III}), as discussed above (\fref{run0461+462_Esdf-res}a): first, energies $E \in [300,600]$\ keV are redistributed until the radial gradient in this energy area has flattened and the redistribution continues in $E \in[200,400]$\ keV. In the loss spectrum over time (\fref{run0461+462_tEloss}), the successive transport across the loss boundary can be recognized as losses. Due to the many poloidal harmonics resulting in a very broad radial structure of the TAE (as will be shown in \sref{sec-poloidalharmonics}), even a very low energy resonance is able to transport particles across the loss boundary, that in turn appear as a thin loss peak at $E \in [100,150]$\ keV, once the TAE has reached the stochastization level. \subsubsection{Influence of the $q$ Profile} In scenario1.51, where the loss region is smaller, the TAE is not broad enough, to redistribute particles with $E < 350$\ keV across the loss boundary. Although one reason for this is the lower amplitude levels reached in scenario1.51 (compared to scenario1.16), the main reason for the missing of the lowest energetic losses (around $\approx 180$\ keV) and the small amount in the whole range of $E \in [300,600]$\ keV is found in the different equilibrium, i.e.\ the smaller loss region and the larger distance between the resonance lines. This is proved by simulating the same TAE and RSAE eigenmode structures of scenario1.16 in both equilibria with a fixed mode amplitude of $\delta B/B = 5.1\cdot 10^{-3}$ (an experimentally realistic value): as shown in \fref{run0502+506+507_Eloss}, even with the same mode amplitude, structure, mode number $n$ and frequency, the monotonic $q$ profile\ case leads to far fewer losses. No losses at all are observed in $E \in [300,600]$\ keV. The complete inhibition of the very low energy peak around 180\ keV however, is caused by a combination of the monotonic $q$ profile\ equilibrium and the corresponding eigenmode.\\ The drop in the height of the very low energy peak not only results from the smaller loss area in the monotonic $q$ profile\footnote{$~$Note that the resonances for the scenario resulting in the red curve differ from what is shown in \fref{resplot_AUG-23824}b, because as mode frequencies, the values from the modes in the inverted $q$ profile\ equilibrium were used.}, but also from a different redistribution mechanism: due to the less dense resonances in phase space around $E \in [100;600]$\ keV, the redistribution in the monotonic $q$ profile\ is less efficient but spreads wider towards higher energies. In combination with the different loss boundaries, this leads to the missing of losses in the energy range of $E \in [300,600]$\ keV. The higher the energy is, the smaller is the effect of the equilibrium on the loss spectrum. This is quite logical, since high energy particles are on large orbits, and therefore require only a small perturbation to be expelled. \begin{figure}[H] \centering \includegraphics[width=0.7\textwidth,height=4cm]{figure12-eps-converted-to.pdf} \caption{\itshape Energy spectra of losses as numerically calculated within scenario1.16 (black curve), within scenario1.51 (green curve) and as simulated within the monotonic $q$ profile\ equilibrium, using the eigenmodes given by \textsc{ligka} for the inverted $q$ profile\ equilibrium (\fref{ligka-input}a) (red curve). The amplitudes for these simulations were fixed at $\delta B/B = 5.1\cdot 10^{-3}$. The losses shown appeared in a time interval starting after c.a.\ 10 RSAE wave periods ($t \in [0.2,1.5] \cdot 10^{-3}$\ s).} \label{run0502+506+507_Eloss} \end{figure} \subsubsection{Influence of the Poloidal Harmonics}\label{sec-poloidalharmonics} Next, the effect resulting from the many poloidal harmonics of the TAE is investigated. By calculating the eigenmodes with the \textsc{Ligka} solver, the poloidal harmonics are known and can be used within the \textsc{Hagis} simulation. Comparing two simulations -- one with the poloidal harmonics $m=4$ and $m=5$ only, the other one with all harmonics (from $m=4$ to $m=10$) as shown in \fref{ligka-input}a (lower) -- shows a similar qualitative mode amplitude evolution with similar mode saturation levels. However, the TAE growth rate is lower if simulating with all poloidal harmonics, as the energy has to be distributed over more poloidal harmonics. In the saturation phase, this effect is compensated by the wider radial range, i.e.\ the mode with all poloidal harmonics is able to tap energy from the radial gradient also radially further outside. \begin{figure}[H] \centering \subfigure[\itshape Redistribution in phase space ($t =3.6 \cdot 10^{-3}$\ s)]{\includegraphics[width=0.48\textwidth,height=5.5cm]{figure13a-eps-converted-to.pdf}} \hspace{0.3cm}\subfigure[\itshape Losses over time]{\includegraphics[width=0.48\textwidth,height=5.5cm]{figure13b-eps-converted-to.pdf}} \caption{\itshape Redistribution (a, blue means particles move away, red means they accumulate) in energy and radial position space and losses at the first wall (b) of a simulation in the inverted $q$ profile\ ($\beta_{\mathrm{fp}}$=0.05\%), with modes given by \textsc{Ligka} (see \fref{ligka-input}a), but the TAE is simulated with only its first two poloidal harmonics. The blue line in a gives the loss boundary (as explained in \fref{resplot_AUG-23824}a), the gray curves in b the mode amplitude evolution.} \label{run0451_Esdf+tEloss} \end{figure} This effect is well visible in the redistribution: only if all poloidal harmonics are present, the TAE redistributes particles at very low energies (around $E \approx 100$\ keV) and far outside radial positions ($s > 0.8$), as becomes clear when comparing \fref{run0461+462_Esdf-res}a with \fref{run0451_Esdf+tEloss}a. This redistribution results in the observed low energy losses of $E < 200$\ keV, that are not observed in the simulation with only two poloidal harmonics (compare \fref{run0461+462_tEloss}a and \fref{run0451_Esdf+tEloss}b).\\ In summary, it is clear now, that the losses with $E < 200$\ keV result from the combination of the broad TAE with many harmonics and the large loss region appearing in the inverted $q$ profile\ equilibrium. However, it was shown above (see \fref{run0502+506+507_Eloss}), that significantly less of these losses appear when simulating the same modes at the same amplitudes in the monotonic $q$ profile, due to the smaller loss region. Concerning the losses in the lower energy range between around 300\ keV up to c.a.\ 600\ keV, it was found that these are due to the dense resonance lines meeting the loss boundary in the inverted $q$ profile\ scenario, quite independent of the mode structure. In the monotonic $q$ profile, these losses do therefore not occur. \subsubsection{Double Mode Effect} As the mode amplitudes in the nonlinear saturation are known now from simulations with consistent mode evolution, \textsc{Hagis} is run with these mode amplitudes kept fixed. This offers the possibility to investigate the losses caused by each of the modes alone vs.\ the losses caused by double-mode interaction, but without the additional effect that different amplitudes are reached in single-mode simulations compared to double-mode simulations. \Fref{run0502+503+504_Eloss} shows the energy spectra of the losses appearing in such double-mode simulation (black curve), as well as those obtained, when simulating both modes individually (blue for the TAE and pink for the RSAE). Note that the loss spectra curves obtained in the single-mode simulations are shown multiplied by a factor of 10. For each of the three spectra, the same time interval is chosen, starting after approximately 10 wave RSAE periods, to avoid the effect of prompt losses. All mode amplitudes are fixed at $\delta B/B = 5.1\cdot 10^{-3}$, which is a reasonable value, when comparing to experimentally measured amplitudes. Further, it is of the order of the minimum saturation level reached in the previous simulations with consistent mode evolution. It is not surprising that the more core-localized RSAE causes significantly fewer losses than the TAE, which is broad and located at a higher radial position. However, for both single-mode simulations, the losses' energy spectra exhibits clearly the different resonances: losses appear at energies, where resonance lines meet the loss boundary (with \fref{resplot_AUG-23824}a the bounce harmonic numbers $p$ can be identified). But the most interesting fact is found, when looking at the double-mode simulation's losses: they do by far exceed the sum of the losses of the single-mode simulations, although the amplitude levels are the same. The redistribution in phase space is consistent with the appearance of losses in the double-mode simulation versus the single-mode simulations. As well as the losses, also the redistribution in the double-mode simulation exceeds by far those of the single-mode cases. Thus, the large number of losses in multi-mode situations is not only caused by the higher mode amplitudes that are possibly reached then (as was shown in ref.\ \cite{Schneller12}). The same reason that leads to higher mode amplitude levels -- the lowering of the stochasticity threshold -- also causes more losses, because the redistribution is enhanced by the stochastization. If the loss boundary is large, as in the inverted $q$ profile\ equilibrium, this redistribution leads to a large number of losses.\\ \begin{minipage}{0.48\textwidth} \vspace{2cm} \begin{figure}[H] \centering \includegraphics[width=1\textwidth,height=3.5cm]{figure14-eps-converted-to.pdf} \caption{\itshape Energy spectra of losses in three different} \label{run0502+503+504_Eloss} \end{figure} \vspace{-0.5cm}{\itshape \small simulations with fixed mode amplitudes at $\delta B/B = 5.1\cdot 10^{-3}$ for a time interval starting after ap\-prox\-imately 10 RSAE wave periods ($t \in [2.0,1.2] \cdot 10^{-3}$\ s): the pink and the blue curves give the loss spectrum of the RSAE and TAE single-mode simulation, stretched by a factor of 10. The black curve gives the losses appearing if simulating both modes (\fref{ligka-input}a) together.} \vspace{0.4cm} \end{minipage} \hfill\begin{minipage}{0.50\textwidth} \begin{figure}[H] \centering \includegraphics[width=1\textwidth,height=5.5cm]{figure15-eps-converted-to.pdf} \caption{\itshape Redistribution in energy and radial position} \label{run0503+504_Esdf3300} \end{figure} \vspace{-0.5cm}{\itshape \small space caused by the TAE (color code) and the RSAE (boundary lines) in the fixed amplitude simulation (at $\delta B/B = 5.1\cdot 10^{-3}$). Red indicates particle ac\-cu\-mu\-la\-tion (higher $E\delta f$), blue that particles move away (lower $E\delta f$). The dark line gives the loss boundary of the inverted $q$ profile\ equilibrium.} \vfill \end{minipage}\\ In the double-mode scenario, stochastization is enhanced due to the dense coverage of phase space with both resonance lines and perturbation structure. In \fref{run0503+504_Esdf3300}, this effect can be observed clearly: the RSAE transports energetic particles into phase space areas from where they are further redistributed by the TAE. Since this effect also occurs vice versa, both modes mutually refill the phase space areas with particles, where the other mode redistributes them away from. \subsubsection{Comparison of the Loss Ejection Signature} The first striking statement of ref.\ \cite{Garcia10} is the difference in the \textbf{total amount of losses} between the earlier time points, $t < 1.4$\ s, when a large number of losses appears in the whole energy range, and later times, $t > 1.4$\ s, when there are only few losses in the high energy channel \cite{Garcia10} and almost none in the low energy channel \cite{GarciaIAEA09}. Focusing on the two time points $t=1.16$\ s and $t=1.51$\ s within the experimentally measured loss signal reveals (figure 4 of ref.\ \cite{Garcia10}): at $t=1.16$\ s, an equally high amount of losses is found in the low and the high energy channel, whereas at $t=1.51$\ s, no low energy losses appear -- except for a general noise level, and fewer ($\approx 18\%$) in the high energy range. This result is well reproduced by the numerical simulations, where scenario1.16 gives a large number of losses in the lower and in the higher energy range, with quite comparable levels, relative to each other. In scenario1.51, no losses appear below 300\ keV and a few for higher energies ($E \gtrsim 300$\ keV), mostly within $E \in [550;750]$\ keV. The drop of losses in the higher energy range between scenario1.16 and scenario1.51 is in the range of around 10\% in the simulation. However, to compare the losses quantitatively with the ones measured at the \textsc{Fild} is only possible in the lower energy channel, since in the higher energy channel, prompt losses are superimposed. These prompt losses cannot be quantified within the model, as explained in \sref{sec:simcond}.\\ Next, the \textbf{nature of losses -- incoherent or coherent} is discussed. In the experiment, incoherent diffusive losses are observed \cite{Garcia10}, identified by a typical quadratic scaling with the mode amplitude ($\propto (\delta B/B)^2$) \cite{Sigmar92}. It is an important result of the presented realistic simulations, that this quadratic scaling is now found also numerically (\fref{run0461_tloss}b). Further, the quadratic scaling is another evidence for the diffusive character of the losses in the lower energy range of scenario1.16. The simulations for $t=1.51$\ s (see \fref{run0461+462_tEloss}b) give losses only in the high energy range. There are both prompt, i.e.\ incoherent losses and non-prompt losses, which show clearly an energy spectrum correlated to the resonance energies at the loss boundary. However, there is also a small amount of non-prompt incoherent losses (diffusive). Thus, the few incoherent losses seen also in the \textsc{Fild} signal can be both prompt or diffusive. But taking into account, that the mode amplitudes were slightly overestimated in the simulation, it can be concluded that there are only very few losses with diffusive origin in the experiment. \begin{figure}[H] \centering \subfigure[\itshape {Losses' signature for the energy range of $E\in [90,350]$\ keV. Light blue: underlying incoherent losses.}]{\includegraphics[width=0.8\textwidth,height=3cm]{figure16a-eps-converted-to.pdf}} \subfigure[\itshape {Time integrated losses with $E\in [90,350]$\ keV over TAE amplitude (green) and quadratic fit (red).}]{\includegraphics[width=0.8\textwidth,height=3cm]{figure16b-eps-converted-to.pdf}} \caption{\itshape Losses at the first wall in scenario1.16.} \label{run0461_tloss} \end{figure} In the experiment, the ratio of coherent to incoherent losses at $t=1.16$\ s is about 1:5 in the low and roughly 1:2.5 in the high energy channel. A possible overestimation of incoherent losses due to the smearing effects (caused by a small deuterium fast ion population, magnetic field ripple, etc.) has to be taken into account. In the simulation, coherent (resonant) losses are found as well, especially in the medium to higher energy range. They can be clearly identified via a Fourier analysis of the ejection signature, which revealed the main peak at the mode's beat frequency. The maximum ratio of incoherent (caused by phase space stochastization, combined with the large loss region of the inverted $q$ profile) to coherent (resonant) losses appears in the low energy range and is about 1:1 (as illustrated by \fref{run0461_tloss}a). Thus, it is lower than the experimental finding, but might be increased when taking into account more than two modes. In the higher energy range, a quantitative analysis is not possible due to the prompt losses (as mentioned above). Though, a rough estimate is, that a large fraction of the losses in this energy range is prompt: since no prompt losses occur in the low energy range, one can calculate a scaling factor between the experimental loss amplitude and the simulated one. With this factor, the simulated losses in the high energy range can be compared to the measured ones. The difference gives the prompt losses, that are missing in the simulation. In the case of the most realistic simulation, this allows one to estimate, that between around 5\% and 50\% of the measured incoherent losses in the high energy channel would be prompt. However, this can be only a rough estimate, due to the uncertainties that enter the comparison. Besides the limitations of the model (discussion see ref.\ \cite{mirjam_phd}), these are the width of the energy channels, the \textsc{Fild} noise level, and the fact, that the scaling between experiment and simulation can depend on the energy, since the losses at different energies are caused by different loss mechanisms. If one considers smearing effects at the \textsc{Fild}, that overestimate the incoherent losses, the amount of prompt losses within the incoherent losses are expected to be higher. \subsubsection{Comparison of the Phase Space Pattern} In the next step, the \textbf{energy-pitch angle characteristics }of the numerical losses is compared with the experimental measurement. Using the synthetic \textsc{Fild} diagnostics as described in ref.\ \cite{mwb_phd}, it is found that experimental (colored red in \fref{Numeric_vs_Exp-116}, from ref.\ \cite{Garcia10}) and numerical loss pattern (green and blue line for the boundary of the prompt and the non-prompt loss pattern) match very well in phase space. The numerical results indicate, that the higher energy (or gyroradius-) incoherent losses are prompt losses, whereas the lower energy (or gyroradius-) incoherent losses are caused by phase space stochastization due to the presence of multiple modes. The highest energies (gyro radii) detected at the \textsc{Fild} do not appear in the simulation, as the maximum energy simulated was $E_\mathrm{max} = 1.2$\ MeV (indicated by the gray line in \fref{Numeric_vs_Exp-116}).\\ To allow for a direct comparison to the \textsc{Fild} measurement, the numerical values in \fref{Numeric_vs_Exp-116} have been drift-corrected. This means that both energy (gyro radius) and pitch of each lost particle has to be shifted to account for a deviation caused by the experimental method of the \textsc{Fild}: since the particle continues drifting within the detector, the measured gyro radius and the pitch angle are overestimated. Ref.\ \cite{mwb_phd}, p.\ 115-117 gives a more detailed explanation and calculated the deviation for comparable cases to $\approx 6$\% in the gyro radius and $\approx 9^o$ in pitch angle. \section{Conclusions and Outlook} In this work, multi-mode Alfv{\'e}nic fast particle transport was modeled with the vacuum-extended \textsc{Hagis} code \cite{mwb_phd}. One aim of the investigation was to compare numerical fast particle losses with experimental measurements of the \textsc{Icrh}-heated \textsc{Asdex} Upgrade discharge \#23824 \cite{Garcia10} and to gain a deeper understanding of the transport processes. This was achieved through the investigation of the internal redistribution in combination with existing resonances, as well as phase space and frequency analysis of the losses. The simulations have been carried out within a realistic model in various respects: on the energetic particle side, a more general, consistent \textsc{Icrh}-like distribution function was implemented. It accounts for the strong anisotropy of the \textsc{Icrh}-generated fast particle population. However, it is still assumed to be separable in its coordinates. This is a constraint concerning a realistic representation that is hoped to be overcome soon, especially for the radial and energy coordinates: the analytical model from ref.\ \cite{Troia12}, adjusted to realistic conditions with the help of the \textsc{Toric-ssfpql} code is planned to be implemented into \textsc{Hagis}. For the results presented in this work, crucial changes were ruled out via a sensitivity scan on the distribution function. \begin{figure}[H] \centering \includegraphics[width=0.7\textwidth,height=5.5cm]{figure17-eps-converted-to.pdf} \caption{\itshape Loss pattern (colored red) in phase space (gyroradius $\rho$ over pitch angle $\lambda^o$) as measured at the \textsc{Fild} in \textsc{aug} discharge \#23824 at time $t \in [1.14,1.16]$\ s from ref.\ \cite{GarciaIAEA09}. Further, the loss pattern as resulting from the simulation of an \textsc{Icrh}-generated fast particle distribution function in the MHD equilibrium of \textsc{aug} \#23824 at $t= 1.16$\ s with the corresponding perturbation given by \textsc{Ligka} (see \fref{ligka-input}a) is shown. For the distribution function see \sref{sec:simcond}, esp.\ \eref{formula:Lambdis},\eref{formula:lampollim}. The blue line gives the boundary of the non-prompt losses at the first wall -- where the majority appears in the respective lower $\rho$ part within the boundary. The green line depicts the boundary of the prompt losses' appearance at the first wall (simulated with the distribution function of \eref{formula:Lambdis},\eref{formula:lampollim2}). Since the simulation range in the energy was restricted to $E \leq 1.2$\ MeV, there is an area in $\rho$-$\lambda^o$ space that has not been simulated: it is indicated by the gray line. The numerical values have been drift-corrected as explained in ref.\ \cite{mwb_phd}.} \label{Numeric_vs_Exp-116} \end{figure} On the perturbation side, \textsc{Hagis} has been extended to read kinetic perturbation data from the eigenvalue solver \textsc{Ligka} \cite{Lauber07}. The simulated losses' phase space pattern was found to coincide very well with the experimental one. Especially in multi-mode scenarios with different mode frequencies, stochastic redistribution sets in over a broad energy range, leading to lower energetic diffusive (incoherent) losses. The quadratical scaling of the amount of losses with the mode amplitude, which is predicted by theory \cite{Sigmar92} and seen in the experiment could be reproduced in realistic numerical simulations. Resonant losses appear from the late linear phase on, mainly in the intermediate to higher energy range, showing good coherence with the mode frequencies and especially their beat frequencies. The higher energetic part of experimentally measured incoherent losses has been identified as mainly prompt losses. The simulations using eigenmode structure given by \textsc{Ligka} revealed that the lowest energetic losses result from a combination of two interconnected facts: first, the many poloidal harmonics of the Toroidicity-induced Alfv{\'e}n eigenmode, which are caused by the gap alignment in the continuum of an inverted $q$ profile\ equilibrium. Second, due to particle drift orbits that are radially extended on the size of the machine's small radius. In scenarios like this, especially in the presence of multiple modes with different frequencies, i.e.\ with resonances, that are complementary in phase space, a domino \cite{Berk95-III} effect can occur: particles in the high energy range leave the plasma as prompt losses, followed by resonant and diffusive losses in the lower energy region. At the same time, the redistribution caused by the core-localized mode refills the particles, that have been transported radially outwards by the outer mode. When the outer mode reaches the stochasticity threshold, low energy losses appear (down to 1/3 of the birth energy), and if many poloidal harmonics are present, they transport even very low energetic particles (down to 1/10 of the birth energy) across the loss boundary. This phase space channeling effect, caused by the presence of multiple modes, is clearly a nonlinear phenomenon, although the regime is still weakly nonlinear. Losses are enhanced by orders of magnitude. This result stresses the importance of the mode structure and thus infers a possible control of energetic particle transport via gap-dealignment by $q$ profile\ and density shaping.\\ Another important result is the crucial role of the linearly subdominant mode for the nonlinear energetic particle transport. To follow up this finding, further studies are planned in the near future, investigating the nonlinear behaviour of a ``sea'' of linearly stable or weakly unstable modes. Such scenario is one of those considered most realistic for \textsc{Iter}. \section*{Acknowledgment} This work was facilitated by the Max-Planck/Princeton Center for Plasma Physics. The simulation results were obtained with the help of the high performance computing resources provided by the RZG in Garching, Germany, on the \textsc{Hpc-FF} system at Forschungszentrum J\"ulich, Germany, and on CSC's \textsc{Iferc-Helios} in Japan. Thanks. \bibliographystyle{apsrev}
\section{Introduction} Approximate Bayesian Computation (ABC) is a popular method for likelihood-free Bayesian inference. ABC methods were originally introduced in population genetics, but are now widely used in applications as diverse as epidemiology~\citep{TanFraLucSis06, BluTra10, WalAllLeeSma10}, materials science~\citep{BorColSis07}, parasitology~\citep{DroPet11}, genetic evolution~\citep{ThoAnd06, Fag-etal07, Rat-etal09, WegExc10, Beau10, Wil-etal11}, population migration~\citep{Gui-etal10} and conservation studies~\citep{LopBoe10}. One of the earliest articles about the ABC approach, covering applications in population genetics, is~\citet{TavBalGriDon97}. Newer developments and extensions of the method include placing ABC in an MCMC context~\citep{Maj-etal03}, sequential ABC~\citep{Sis-etal07}, enhancing ABC with nonlinear regression models~\citep{BluFra10}, agent-based modelling~\citep{SotTav10}, and a Gibbs sampling ABC approach~\citep{Wil-etal11}. A survey of recent developments is given by~\citet{MarEtAl12}. ABC methods allow for inference in a Bayesian setting where the parameter $\theta\in\mathbb{R}^p$ of a statistical model is assumed to be random with a given prior distribution~$f_\theta$, we have observed data $X \in\mathbb{R}^d$ from a given distribution~$f_{X|\theta}$ depending on $\theta$ and we want to use these data to draw inference about~$\theta$. In the areas where ABC methods are used, the likelihood $f_{X|\theta}$ is typically not available in an explicit form. The term ``likelihood-free'' is used to indicate that ABC methods do not make use of the likelihood $f_{X|\theta}$, but only work with samples from the joint distribution of $\theta$ and $X$. In the context of ABC methods, the data are usually summarised using a statistic $S\colon \mathbb{R}^d\to \mathbb{R}^q$, and analysis is then based on $S(X)$ instead of~$X$. The choice of $S$ affects the accuracy of the ABC estimates: the estimates can only be expected to converge to the true posterior if $S$ is a sufficient statistic, otherwise additional error is introduced. The basic idea of ABC methods is to replace samples from the exact posterior distribution $\theta\,|\,X\!=\!x^{\mathstrut*}$ or $\theta\,|\,S(X)\!=\!s^{\mathstrut*}$ with samples from an approximating distribution like $\theta \,|\, S(X)\!\approx\! s^{\mathstrut*}$. There are many variants of the basic ABC method available, including different implementations of the condition $S(X) \approx s^{\mathstrut*}$. All variants use a tolerance parameter~$\delta$ which controls the trade-off between fast generation of samples (large values of $\delta$) and accuracy of samples (small values of~$\delta$). The easiest approach to implement the condition $S(X) \approx s^{\mathstrut*}$, considered in this paper, is to use $\| S(X) - s^{\mathstrut*} \| \leq \delta$ where $\|{\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}}\|$ is some norm on~$\mathbb{R}^q$. Different approaches to choosing the statistic~$S$ are used; a semi-automatic approach is described in \citet{FeaPra12}. In many cases considerable improvements can be achieved by choosing the norm for comparison of $S(X)$ to $s^{\mathstrut*}$ in a problem-specific way. Despite the popularity of ABC methods, theoretical analysis is still in its infancy. The aim of this article is to provide a foundation for such analysis by providing rigorous results about the convergence properties of the ABC method. Here, we restrict discussion to the most basic variant, to set a baseline to which different ABC variants can be compared. We consider Monte Carlo estimates of posterior expectations, using the ABC samples for the estimate. Proposition~\ref{p:convergence} shows that such ABC estimates converge to the true value under very weak conditions; once this is established we investigate the rate of convergence in theorem~\ref{t:rate}. Similar results, but in the context of estimating posterior densities rather than posterior expectations can be found in \citet{Blum10} and \citet{Biau13}. The choice of norm and of the tolerance parameter~$\delta$ are major challenges in the practical application of ABC methods, but not many results are available in the literature. A numerical study of the trade-off between accuracy and computational cost, in the context of sequential ABC methods, can be found in~\citet{SiFiStu13}. \citet{Wil13} establishes that an ABC method which accepts or rejects proposals with a probability based on the difference between the observed and proposed data converges to the correct solution under assumptions on model or measurement error. The error of an ABC estimate is affected both by the bias of the ABC samples, controlled by the tolerance parameter~$\delta$, and by Monte Carlo error, controlled by the number~$n$ of accepted ABC samples. One of the key ingredients in our analysis is the result, shown in lemma~\ref{l:bias}, that for ABC estimates the bias satisfies \begin{equation*} \mathrm{bias} \sim \delta^2. \end{equation*} Similarly, in lemma~\ref{l:cost} we show that for the ABC estimate we have \begin{equation*} \mathrm{cost} \sim n \delta^{-q}, \end{equation*} where $q$ is the dimension of the observation~$s^{\mathstrut*}$. It is well-known that for Monte Carlo methods the error decays, as a function of computational cost, proportional to $\cost^{-1/2}$, where the exponent $-1/2$ is independent of dimension \citep[see \textit{e.g.}][section~3.2.2]{Vo13}. In contrast, the main result of this article, theorem~\ref{t:rate}, shows that, under optimal choice of $\delta$, the basic ABC method satisfies \begin{equation*} \mathrm{error} \sim \mathrm{cost}^{-2 / (q+4)}. \end{equation*} Corollary~\ref{c:rate-hat} shows that this result holds whether we fix the number of accepted samples (controlling the precision of our estimates but allowing the computational cost to be random) or fix the number of proposals (allowing the number of accepted samples to be random). The former is aesthetically more satisfying, but in practice most users have a fixed computational budget and hence must fix the number of proposals. In either case, the rate of decay for the error gets worse as the dimension~$q$ increases and even in the one-dimensional case the exponent $-2/(1+4) = -2/5$ is worse than the exponent $-1/2$ for Monte Carlo methods. \citet{FeaPra12} obtain the same exponent $-2/(q+4)$ for the specific summary statistic $S(X) = \mathbb{E}\cond{\theta}{X}$. For the problem of estimating the posterior density, \citet{Blum10} reports the slightly worse exponent $-2/(q+5)$. The difference between Blum's results and ours is due to the fact that for kernel density estimation an additional bandwidth parameter must be considered. We continue by giving a very short introduction to the basic ABC method in section~\ref{S:ABC}. The main results, proposition~\ref{p:convergence} and theorem~\ref{t:rate} are presented in section~\ref{S:results}, together with their proofs. Section~\ref{S:fixed-N} shows that our results are independent of whether the number of proposals or of accepted samples are fixed in the algorithm. Section~\ref{S:experiments} illustrates the results of this paper with the help of numerical experiments. Finally, in section~\ref{S:discussion}, we consider the practical implications of our results. \section{Approximate Bayesian Computation} \label{S:ABC} This section gives a short introduction to the basic ABC algorithm. A more complete description is, for example, given in~\citet[section~5.1]{Vo13}. We describe the algorithm in the context of the following Bayesian inference problem: \begin{itemize} \item A parameter vector $\theta\in\mathbb{R}^p$ is assumed to be random. Before observing any data, our belief about its value is summarised by the prior distribution~$f_\theta$. The value of~$\theta$ is unknown to us and our aim is to make inference about~$\theta$. \item The available data $X \in\mathbb{R}^d$ are assumed to be a sample from a distribution $f_{X|\theta}$, depending on the parameter~$\theta$. Inference about $\theta$ is based on a single observed sample $x^{\mathstrut*}$ from this distribution; repeated observations can be assembled into a single vector if needed. \item In the context of ABC, the data are often summarised using a statistic $S\colon \mathbb{R}^d \to \mathbb{R}^q$. Since $X$ is random, $S = S(X)$ is random with a distribution $f_{S|\theta}$ depending on~$\theta$. If a summary statistic is used, inference is based on the value $s^{\mathstrut*} = S(x^{\mathstrut*})$ instead of on the full sample~$x^{\mathstrut*}$. \end{itemize} Our aim is to explore the posterior distribution of $\theta$, \textit{i.e.}\ the conditional distribution of $\theta$ given $X = x^{\mathstrut*}$, using Monte Carlo methods. More specifically, we aim to estimate the posterior expectations \begin{equation}\label{eq:ABC-answer} y = E\condl{h(\theta)}{X=x^{\mathstrut*}} \end{equation} for given test functions~$h\colon \mathbb{R}^p\to \mathbb{R}$. Expectations of this form allow us to study many relevant properties of the posterior distribution, including the posterior mean when $h(\theta) = \theta_i$ with $i=1, \ldots, p$ and posterior second moments when $h(\theta) = \theta_i \theta_j$ with $i,j=1, \ldots, p$. If $h(\theta) = 1_A(\theta)$ for a given set $A\subseteq\mathbb{R}^p$, the expectation $E\condl{h(\theta)}{X=x^{\mathstrut*}}$ equals the posterior probability of hitting~$A$; for example, the CDF can be approximated by choosing $A=(-\infty, a]$ for $a\in\mathbb{R}$. The basic ABC method for generating approximate samples from the posterior distribution is given in the following algorithm. \begin{algorithm}\label{alg:ABC} For a given observation $s^{\mathstrut*}\in\mathbb{R}^q$ and given tolerance $\delta>0$, {\em ABC samples} approximating the distribution $\theta\,|\,S=s^{\mathstrut*}$ are random samples $\theta^{(\delta)}_j \in \mathbb{R}^p$ computed by the following algorithm: \begin{algorithmic}[1] \STATE let $j \gets 0$ \WHILE{$j < n$} \STATE sample $\theta$ with density $f_\theta({\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}})$ \STATE sample $X$ with density $f_{X|\theta}\cond{{\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}}}{\theta}$ \IF{$\|S(X) - s^{\mathstrut*}\|_A \leq \delta$} \STATE let $j \gets j + 1$ \STATE let $\theta^{(\delta)}_j \gets \theta$ \ENDIF \ENDWHILE \end{algorithmic} \end{algorithm} \noindent The norm $\| {\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}} \|_A$ used in the acceptance criterion is defined by $\| s \|_A^2 = s^\top A^{-1} s$ for all $s\in\mathbb{R}^q$, where $A$ is a positive definite symmetric matrix. This includes the case of the Euclidean distance $\| {\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}} \| = \| {\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}} \|_I$ for $A=I$. In practical applications, the matrix~$A$ is often chosen in a problem-specific way. Using the output of the algorithm, an estimate for the posterior expectation~\eqref{eq:ABC-answer} can be obtained as \begin{equation}\label{eq:ABC-MC-basic} Y^{(\delta)}_n = \frac{1}{n} \sum_{j=1}^n h(\theta^{(\delta)}_j) \end{equation} where $\theta^{(\delta)}_1, \ldots, \theta^{(\delta)}_n$ are computed by algorithm~\ref{alg:ABC}. Since the output of the ABC algorithm approximates the posterior distribution, the Monte Carlo estimate~$Y^{(\delta)}_n$ can be used as an approximation to the posterior expectation. The ABC samples are only approximately distributed according to the posterior distribution, and thus the estimate $Y^{(\delta)}_n$ will not exactly converge to the true value~$y$ as $n\to\infty$. The quality of the approximation can be improved by decreasing the tolerance parameter~$\delta$, but this leads at the same time to a lower acceptance probability in algorithm~\ref{alg:ABC} and thus, ultimately, to higher computational cost for obtaining the estimate $Y^{(\delta)}_n$. Hence, a trade-off must be made between accuracy of the results and speed of computation. Since the algorithm is not using $x^{\mathstrut*}$ directly, but uses $s^{\mathstrut*} = S(x^{\mathstrut*})$ instead, we require $S$ to be a sufficient statistic, so that we have \begin{equation*} y = E\condl{h(\theta)}{X=x^{\mathstrut*}} = E\condl{h(\theta)}{S=s^{\mathstrut*}}. \end{equation*} If $S$ is not sufficient, an additional error will be introduced. For {\em application} of the ABC method, knowledge of the distributions of $\theta$, $X$ and $S=S(X)$ is not required; instead we assume that we can simulate large numbers of samples of these random variables. In contrast, in our {\em analysis} we will assume that the joint distribution of $\theta$ and $S$ has a density $f_{S,\theta}$ and we will need to consider properties of this density in some detail. To conclude this section, we remark that there are two different approaches to choosing the sample size used to compute the ABC estimate~$Y^{(\delta)}_n$. If we denote the number of proposals required to generate $n$ output samples by $N \geq n$, then one approach is to choose the number $N$ of proposals as fixed; in this case the number $n \leq N$ of accepted proposals is random. Alternatively, for given $n$, the loop in the ABC algorithm could be executed until $n$ samples are accepted, resulting in random~$N$ and fixed~$n$. In order to avoid complications with the definition of $Y^{(\delta)}_n$ for $n=0$, we follow the second approach here and defer discussion of the case of fixed $N$ until section~\ref{S:fixed-N}. \section{Results} \label{S:results} This section presents the main results of the paper, in proposition~\ref{p:convergence} and theorem~\ref{t:rate}, followed by proofs of these results. Throughout, we assume that the joint distribution of $\theta$ and $S$ has a density $f_{S,\theta}$, and we consider the marginal densities of $S$ and $\theta$ given by \begin{equation*} f_S(s) = \int_{\mathbb{R}^p} f_{S,\theta}(s,t) \,dt \end{equation*} for all $s\in\mathbb{R}^q$ and \begin{equation*} f_\theta(t) = \int_{\mathbb{R}^q} f_{S,\theta}(s,t) \,ds \end{equation*} for all $t\in \mathbb{R}^p$, respectively. We also consider the conditional density of $\theta$ given $S=s$, defined by \begin{equation*} f_{\theta|S}\cond{t}{s} = \begin{cases} \frac{f_{S,\theta}(s,t)}{f_S(s)}, & \mbox{if $f_S(s) > 0$, and} \\ 0 & \mbox{otherwise.} \end{cases} \end{equation*} Our aim is to study the convergence of the estimate~$Y^{(\delta)}_n$ to $y$ as $n \to \infty$. The fact that ABC estimates converge to the correct value as $\delta \downto 0$ is widely accepted and easily shown empirically. This result is made precise in the following proposition, showing that the convergence holds under very weak conditions. \begin{prop}\label{p:convergence} Let $h\colon \mathbb{R}^p\to \mathbb{R}$ be such that $\mathbb{E}\bigl(|h(\theta)|\bigr) < \infty$. Then, for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$, the ABC estimate $Y^{(\delta)}_n$ given by~\eqref{eq:ABC-MC-basic} satisfies \begin{enumerate} \item $\displaystyle \lim_{n\to\infty} Y^{(\delta)}_n = \mathbb{E}\bigl( Y^{(\delta)}_n \bigr)$ almost surely for all $\delta>0$; and \item $\displaystyle \lim_{\delta\downto 0} \mathbb{E}\bigl( Y^{(\delta)}_n \bigr) = E\condl{h(\theta)}{S=s^{\mathstrut*}}$ for all $n\in\mathbb{N}$. \end{enumerate} \end{prop} We note that the assumptions required in our version of this result are very modest. Since our aim is to estimate the posterior expectation of $h(\theta)$, the assumption that the prior expectation $\mathbb{E}\bigl(|h(\theta)|\bigr)$ exists is reasonable. Similarly, the phrase ``for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$'' in the proposition indicates that the result holds for all $s^{\mathstrut*}$ in a set~$A\subseteq\mathbb{R}^q$ with $P\bigl(S(X)\in A\bigr) = 1$. Since in practice the value $s^{\mathstrut*}$ will be an observed sample of $S(X)$, this condition forms no restriction to the applicability of the result. The proposition could be further improved by removing the assumption that the distributions of $\theta$ and $S(X)$ have densities. While a proof of this stronger result could be given following similar lines to the proof given below, here we prefer to keep the setup consistent with what is required for the result in theorem~\ref{t:rate}. \bigskip Our second result, theorem~\ref{t:rate}, quantifies the speed of convergence of $Y^{(\delta)}_n$ to~$y$. We consider the mean squared error \begin{equation*} \MSE(Y^{(\delta)}_n) = \mathbb{E}\bigl( ( Y^{(\delta)}_n - y )^2 \bigr) = \Var(Y^{(\delta)}_n) + \bias(Y^{(\delta)}_n)^2, \end{equation*} and relate this to the computational cost of computing the estimate~$Y^{(\delta)}_n$. For the computational cost, rather than using a sophisticated model of computation, we restrict ourselves to the na{\"\i}ve approach of assuming that the time required to obtain the estimate $Y^{(\delta)}_n$ in algorithm~\ref{alg:ABC}, denoted $\cost(Y^{(\delta)}_n)$, is proportional to $a + bN$, where $a$ and~$b$ are constants and $N$ is the number of iterations the while-loop in the algorithm has to perform until the condition $j = n$ is met (\textit{i.e.}\ the number of proposals required to generate $n$ samples). To describe the asymptotic behaviour of MSE and cost, we use the following notation. \begin{notation} For sequences $(a_n)_{n\in\mathbb{N}}$ and $(b_n)_{n\in\mathbb{N}}$ of positive real numbers we write $a_n \sim b_n$ to indicate that the limit $c = \lim_{n\to\infty} a_n / b_n$ exists and satisfies $0 < |c| < \infty$. \end{notation} Our result about the speed of convergence requires the density of $S(X)$ to have continuous third partial derivatives. More specifically, we will use the following technical assumptions on~$S$. \begin{assumption} The density $f_S$ and the function $s \mapsto \int_{\mathbb{R}^p} h(t) \,f_{S,\theta}(s,t) \,dt$ are three times continuously differentiable in a neighbourhood of~$s^{\mathstrut*}\in\mathbb{R}^q$. \end{assumption} \begin{theorem}\label{t:rate} Let $h\colon \mathbb{R}^p\to \mathbb{R}$ be such that $\mathbb{E}\bigl(h(\theta)^2\bigr) < \infty$ and let $S$ be sufficient and satisfy assumption~A. Assume $\Var\condl{h(\theta)}{S=s^{\mathstrut*}} > 0$ and \begin{equation}\label{eq:C-s-ast} C(s^{\mathstrut*}) = \frac{\Delta \phi_h(s^{\mathstrut*}) - y \cdot \Delta\phi_1(s^{\mathstrut*})}% {2(q+2) \phi_1(s^{\mathstrut*})} \neq 0. \end{equation} Then, for $f_S$-almost all $s^{\mathstrut*}$, the following statements hold: \begin{enumerate} \item Let $(\delta_n)_{n\in\mathbb{N}}$ be a sequence with $\delta_n \sim n^{-1/4}$. Then the mean squared error satisfies \begin{equation*} \MSE(Y^{(\delta_n)}_n) \sim \mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)^{-4/(q+4)} \end{equation*} as $n\to\infty$. \item The exponent $-4/(q+4)$ given in the preceding statement is optimal: for any sequence $(\delta_n)_{n\in\mathbb{N}}$ with $\delta_n \downarrow 0$ as $n\to \infty$ we have \begin{equation*} \liminf_{n\to\infty} \frac{\!\MSE(Y^{(\delta_n)}_n)\;}{\;\mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)^{-4/(q+4)}\!} > 0. \end{equation*} \end{enumerate} \end{theorem} The rate of convergence in the first part of theorem~\ref{t:rate} should be compared to the corresponding rate for the usual Monte Carlo estimate. Since Monte Carlo estimates are unbiased, the root-mean squared error for a Monte Carlo estimate is proportional to $1/\sqrt{n}$, while the cost of generating $n$ samples is proportional to $n$. Thus, for Monte Carlo estimates we have $\RMSE \sim \cost^{-1/2}$. The corresponding exponent from theorem~\ref{t:rate}, obtained by taking square roots, is $-2/(q+4)$, showing slower convergence for ABC estimates. This reduced efficiency is a consequence of the additional error introduced by the bias in the ABC estimates. The statement of the theorem implies that, as computational effort is increased, $\delta$ should be decreased proportional to $n^{-1/4}$. For this case, the error decreases proportionally to the cost with exponent $r = -4/(q+4)$. The second part of the theorem shows that no choice of $\delta$ can lead to a better (\textit{i.e.}\ more negative) exponent. If the tolerances $\delta_n$ are decreased in a way which leads to an exponent $\tilde r > r$, in the limit for large values of $n$ such a schedule will always be inferior to the choice $\delta_n = c n^{-1/4}$, for any constant~$c>0$. It is important to note that this result only applies to the limit $n\to\infty$. The result of theorem~\ref{t:rate} describes the situation where, for a fixed problem, the tolerance~$\delta$ is decreased. Caution is advisable when comparing ABC estimates for different~$q$. Since the constant implied in the $\sim$-notation depends on the choice of problem, and thus on~$q$, the dependence of the error on $q$ for fixed~$\delta$ is not in the scope of theorem~\ref{t:rate}. Indeed, \citet{Blum10} suggests that in practice ABC methods can be successfully applied for larger values of~$q$, despite results like theorem~\ref{t:rate}. \bigskip Before we turn our attention to proving the results stated above, we first remark that, without loss of generality, we can assume that the acceptance criterion in the algorithm uses Euclidean distance $\| {\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}} \|$ instead of $\| {\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}} \|_A$. This can be achieved by considering the modified statistic $\tilde S(x) = A^{-1/2} S(x)$ and $\tilde s^{\mathstrut*} = \tilde S(x^{\mathstrut*}) = A^{-1/2} s^{\mathstrut*}$, where $A^{-1/2}$ is the positive definite symmetric square root of~$A^{-1}$. Since \begin{align*} \| S(X) - s^{\mathstrut*} \|_A^2 &= \bigl(S(X) - s^{\mathstrut*}\bigr)^\top A^{-1} \bigl(S(X) - s^{\mathstrut*}\bigr) \\ &= \Bigl( A^{-1/2} \bigl(S(X) - s^{\mathstrut*}\bigr) \Bigr)^\top \Bigl( A^{-1/2} \bigl(S(X) - s^{\mathstrut*}\bigr) \Bigr) \\ &= \| \tilde S(X) - \tilde s^{\mathstrut*} \|^2, \end{align*} the ABC algorithm using $S$ and $\|{\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}}\|_A$ has identical output to the ABC algorithm using $\tilde S$ and the Euclidean norm. Finally, a simple change of variables shows that the assumptions of proposition~\ref{p:convergence} and theorem~\ref{t:rate} are satisfied for $S$ and $\|{\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}}\|_A$ if and only they are satisfied for $\tilde S$ and $\|{\setbox0\hbox{$x$}\hbox to\wd0{\hss$\cdot$\hss}}\|$. Thus, for the proofs we will assume that Euclidean distance is used in the algorithm. The rest of this section contains the proofs of proposition~\ref{p:convergence} and theorem~\ref{t:rate}. In the proofs it will be convenient to use the following technical notation. \begin{definition}\label{def:phi} For $h\colon \mathbb{R}^p\to \mathbb{R}$ with $\mathbb{E}\bigl(|h(\theta)|\bigr) < \infty$ we define $\phi_h\colon \mathbb{R}^q\to \mathbb{R}$ to be \begin{equation*} \phi_h(s) = \int_{\mathbb{R}^p} h(t) f_{S,\theta}(s, t) \,dt \end{equation*} for all $s\in\mathbb{R}^q$ and $\phi^{(\delta)}_h\colon \mathbb{R}^q\to \mathbb{R}$ to be \begin{equation}\label{eq:phi-delta} \phi^{(\delta)}_h(s^{\mathstrut*}) = \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} \phi_h(s) \,ds, \end{equation} for all $s^{\mathstrut*} \in\mathbb{R}^q$, where $\bigl|B(s^{\mathstrut*}, \delta) \bigr|$ denotes the volume of the ball $B(s^{\mathstrut*}, \delta)$. \end{definition} Using the definition for $h\equiv 1$ we get $\phi_1 \equiv f_S$ and for general $h$ we get $\phi_h(s) = f_S(s) \, \mathbb{E}\condl{h(\theta)}{S=s}$. Both the exact value~$y$ from~\eqref{eq:ABC-answer} and the mean of the estimator $Y^{(\delta)}_n$ from~\eqref{eq:ABC-MC-basic} can be expressed in the notation of definition~\ref{def:phi}. This is shown in the following lemma. \begin{lemma}\label{l:phi} Let $h\colon \mathbb{R}^p\to \mathbb{R}$ be such that $\mathbb{E}\bigl(|h(\theta)|\bigr) < \infty$. Then \begin{equation*} \mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}} = \frac{\phi_h(s^{\mathstrut*})}{\phi_1(s^{\mathstrut*})} \end{equation*} and \begin{equation*} \mathbb{E}\bigl( Y^{(\delta)}_n \bigr) = \mathbb{E}\bigl( h(\theta^{(\delta)}_1) \bigr) = \frac{\phi^{(\delta)}_h(s^{\mathstrut*})}{\phi^{(\delta)}_1(s^{\mathstrut*})} \end{equation*} for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$. \end{lemma} \begin{proof} From the assumption $\mathbb{E}\bigl(|h(\theta)|\bigr) < \infty$ we can conclude \begin{equation*} \begin{split} \int_{\mathbb{R}^q} \int_{\mathbb{R}^p} \bigl| h(t) \bigr| f_{S,\theta}(s,t) \,dt \,ds &= \int_{\mathbb{R}^p} \bigl| h(t) \bigr| \int_{\mathbb{R}^q} f_{S,\theta}(s,t) \,ds \,dt \\ &= \int_{\mathbb{R}^p} \bigl| h(t) \bigr| f_\theta(t) \,dt = \mathbb{E}\bigl(|h(\theta)|\bigr) < \infty, \end{split} \end{equation*} and thus we know that $\int_{\mathbb{R}^p} \bigl| h(t) \bigr| f_{S,\theta}(s,t) \,dt < \infty$ for almost all $s\in\mathbb{R}^q$. Consequently, the conditional distribution $\mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}}$ exists for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$. Using Bayes' rule we get \begin{equation*} \begin{split} \mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}} &= \int_{\mathbb{R}^p} h(t) f_{\theta|S}\cond{t}{s^{\mathstrut*}} \,dt \\ &= \int_{\mathbb{R}^p} h(t) \frac{f_{S,\theta}(s^{\mathstrut*}, t)}{f_S(s^{\mathstrut*})} \,dt = \frac{\phi_h(s^{\mathstrut*})}{\phi_1(s^{\mathstrut*})}. \end{split} \end{equation*} On the other hand, the samples $\theta^{(\delta)}_j$ are distributed according to the conditional distribution of $\theta$ given $S \in B(s^{\mathstrut*}, \delta)$. Thus, the density of the samples $\theta^{(\delta)}_j$ can be written as \begin{equation*} f_{\theta^{(\delta)}_j}(t) = \frac{1}{Z} \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} f_{S,\theta}(s,t) \,ds, \end{equation*} where the normalising constant $Z$ satisfies \begin{equation*} Z = \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} f_S(s) \,ds = \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} \phi_1(s) \,ds = \phi^{(\delta)}_1(s^{\mathstrut*}), \end{equation*} and we get \begin{equation*} \begin{split} \mathbb{E}\bigl( Y^{(\delta)}_n \bigr) &= \mathbb{E}\bigl( h(\theta^{(\delta)}_1) \bigr) \\ &= \frac{1}{\phi^{(\delta)}_1(s)} \int_{\mathbb{R}^p} h(t) \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} f_{S,\theta}(s,t) \,ds \,dt \\ &= \frac{1}{\phi^{(\delta)}_1(s)} \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} \phi_h(s) \,ds \\ &= \frac{\phi^{(\delta)}_h(s^{\mathstrut*})}{\phi^{(\delta)}_1(s^{\mathstrut*})}. \end{split} \end{equation*} This completes the proof. \end{proof} Using these preparations, we can now present a proof of proposition~\ref{p:convergence}. \begin{proof}[Proof of proposition~\ref{p:convergence}] Since $\mathbb{E}\bigl( |h(\theta)| \bigr) < \infty$ we have \begin{equation*} \mathbb{E}\bigl(|Y^{(\delta)}_n|\bigr) \leq \mathbb{E}\bigl(|h(\theta_1)|\bigr) = \frac{\phi_{|h|}(s^{\mathstrut*})}{\phi_1(s^{\mathstrut*})} <\infty \end{equation*} whenever $\phi_1(s^{\mathstrut*}) = f_S(s^{\mathstrut*}) > 0$, and by the law of large numbers $Y^{(\delta)}_n$ converges to $\mathbb{E}\bigl( Y^{(\delta)}_n \bigr)$ almost surely. For the second statement, since $\phi_1 \equiv f_S \in L^1(\mathbb{R}^q)$, we can use the Lebesgue differentiation theorem \citep[theorem~7.7]{Rudin66} to conclude that $\phi^{(\delta)}_1(s^{\mathstrut*}) \to \phi_1(s^{\mathstrut*})$ as $\delta \downto 0$ for almost all $s^{\mathstrut*}\in\mathbb{R}^q$. Similarly, since \begin{equation*} \int_{\mathbb{R}^q} \bigl| \phi_h(s) \bigr| \,ds \leq \int_{\mathbb{R}^p} \bigl| h(t) \bigr| \int_{\mathbb{R}^q} f_{S,\theta}(s,t) \,ds \,dt = \int_{\mathbb{R}^p} \bigl| h(t) \bigr| f_\theta(t) \,dt < \infty \end{equation*} and thus $\phi_h \in L^1(\mathbb{R}^q)$, we have $\phi^{(\delta)}_h(s^{\mathstrut*}) \to \phi_h(s^{\mathstrut*})$ as $\delta \downto 0$ for almost all $s^{\mathstrut*}\in\mathbb{R}^q$ and using lemma~\ref{l:phi} we get \begin{equation*} \lim_{\delta\downto 0} \mathbb{E}\bigl( Y^{(\delta)}_n \bigr) = \lim_{\delta\downto 0} \frac{\phi^{(\delta)}_h(s^{\mathstrut*})}{\phi^{(\delta)}_1(s^{\mathstrut*})} = \frac{\phi_h(s^{\mathstrut*})}{\phi_1(s^{\mathstrut*})} = \mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}} \end{equation*} for almost all $s^{\mathstrut*}\in\mathbb{R}^q$. This completes the proof. \end{proof} For later use we also state the following simple consequence of proposition~\ref{p:convergence}. \begin{corollary}\label{C:variance} Assume that $\mathbb{E}\bigl(h(\theta)^2\bigr) < \infty$. Then, for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$ we have \begin{equation*} \lim_{\delta\downarrow0} n \Var(Y^{(\delta)}_n) = \Var\condl{h(\theta)}{S=s^{\mathstrut*}}, \end{equation*} uniformly in~$n$. \end{corollary} \begin{proof} From the definition of the variance we know \begin{equation*} \Var\bigl( Y^{(\delta)}_n \bigr) = \frac{1}{n} \Var\bigl( h(\theta^{(\delta)}_j) \bigr) = \frac{1}{n} \Bigl( \mathbb{E}\bigl( h(\theta^{(\delta)}_j)^2 \bigr) - \mathbb{E}\bigl( h(\theta^{(\delta)}_j) \bigr)^2 \Bigr). \end{equation*} Applying proposition~\ref{p:convergence} to the function $h^2$ first, we get $\lim_{\delta\downarrow0} \mathbb{E}\bigl(h(\theta^{(\delta)}_j)^2\bigr) = \mathbb{E}\condl{h(\theta)^2}{S=s^{\mathstrut*}}$. Since $\mathbb{E}\bigl(h(\theta)^2\bigr) < \infty$ implies $\mathbb{E}\bigl(|h(\theta)|\bigr) < \infty$, we also get $\lim_{\delta\downarrow0} \mathbb{E}\bigl(h(\theta^{(\delta)}_j)\bigr) = \mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}}$ and thus \begin{align*} \lim_{\delta\downarrow 0} n \Var\bigl( Y^{(\delta)}_n \bigr) &= \mathbb{E}\condl{h(\theta)^2}{S=s^{\mathstrut*}} - \mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}}^2 \\ &= \Var\condl{h(\theta)}{S=s^{\mathstrut*}}. \end{align*} This completes the proof. \end{proof} The rest of this section is devoted to a proof of theorem~\ref{t:rate}. We first consider the bias of the estimator~$Y^{(\delta)}_n$. As is the case for Monte Carlo estimates, the bias of the ABC estimate~$Y^{(\delta)}_n$ does not depend on the sample size~$n$. The dependence on the tolerance parameter~$\delta$ is given in the following lemma. This lemma is the key ingredient in the proof of theorem~\ref{t:rate}. \begin{lemma}\label{l:bias} Assume that $\mathbb{E}\bigl(|h(\theta)|\bigr) < \infty$ and that $S$ satisfies assumption~A. Then, for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$, we have \begin{equation*} \bias(Y^{(\delta)}_n) = C(s^{\mathstrut*}) \, \delta^2 + O(\delta^3) \end{equation*} as $\delta \downto 0$ where the constant $C(s^{\mathstrut*})$ is given by equation~\eqref{eq:C-s-ast} and $\Delta$ denotes the Laplace operator. \end{lemma} \begin{proof} Using lemma~\ref{l:phi} we can write the bias as \begin{equation}\label{eq:bias} \bias(Y^{(\delta)}_n) = \frac{\phi^{(\delta)}_h(s^{\mathstrut*})}{\phi^{(\delta)}_1(s^{\mathstrut*})} - \frac{\phi_h(s^{\mathstrut*})}{\phi_1(s^{\mathstrut*})}. \end{equation} To prove the lemma, we have to study the rate of convergence of the averages~$\phi^{(\delta)}_h(s^{\mathstrut*})$ to the centre value $\phi_h(s^{\mathstrut*})$ as $\delta\downto 0$. Using Taylor's formula we find \begin{equation*} \begin{split} \phi_h(s) &= \phi_h(s^{\mathstrut*}) + \nabla \phi_h(s^{\mathstrut*}) (s - s^{\mathstrut*}) + \frac12 (s - s^{\mathstrut*})^\top H_{\phi_h}(s^{\mathstrut*}) (s-s^{\mathstrut*}) \\ &\hskip1cm + r_3(s - s^{\mathstrut*}) \end{split} \end{equation*} where $H_{\phi_h}$ denotes the Hessian of $\phi_h$ and the error term $r_3$ satisfies \begin{equation}\label{eq:error-bound} \bigl| r_3(v) \bigr| \leq \max_{|\alpha|=3} \sup_{s\in B(s^{\mathstrut*},\delta)} \bigl| \d^\alpha_s \phi_h(s) \bigr| \cdot \sum_{|\beta| = 3} \frac{1}{\beta!} \bigl| v^\beta \bigr| \end{equation} for all $s \in B(s^{\mathstrut*}, \delta)$, and $\d^\alpha_s$ denotes the partial derivative corresponding to the multi-index~$\alpha$. Substituting the Taylor approximation into equation~\eqref{eq:phi-delta}, we find \begin{equation}\label{eq:f-delta-1} \begin{split} \phi^{(\delta)}_h(s^{\mathstrut*}) &= \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} \Bigl( \phi_h(s^{\mathstrut*}) + \nabla \phi_h(s^{\mathstrut*}) (s - s^{\mathstrut*}) \Bigr. \\ &\hskip2cm \Bigl. + \frac12 (s - s^{\mathstrut*})^\top H_{\phi_h}(s^{\mathstrut*}) (s-s^{\mathstrut*}) + r_3(s - s^{\mathstrut*}) \Bigr) \,ds \\ &= \phi_h(s^{\mathstrut*}) + 0 \\ &\hskip1cm + \frac{1}{2\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} (s - s^{\mathstrut*})^\top H_{\phi_h}(s^{\mathstrut*}) (s-s^{\mathstrut*}) \,ds \\ &\hskip1cm + \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} r_3(s - s^{\mathstrut*}) \,ds. \end{split} \end{equation} Since the Hessian $H_{\phi_h}(s^{\mathstrut*})$ is symmetric and since the domain of integration is invariant under rotations we can choose a basis in which $H_{\phi_h}(s^{\mathstrut*})$ is diagonal, such that the diagonal elements coincide with the eigenvalues $\lambda_1, \ldots, \lambda_q$. Using this basis we can write the quadratic term in~\eqref{eq:f-delta-1} as \begin{align*} \hskip5mm&\hskip-5mm \frac{1}{2\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} (s - s^{\mathstrut*})^\top H_{\phi_h}(s^{\mathstrut*}) (s-s^{\mathstrut*}) \,ds \\ &= \frac{1}{2\bigl|B(0, \delta)\bigr|} \int_{B(0, \delta)} \sum_{i=1}^q \lambda_i u_i^2 \,du \\ &= \frac{1}{2\bigl|B(0, \delta)\bigr|} \sum_{i=1}^q \lambda_i \frac{1}{q} \sum_{j=1}^q \int_{B(0, \delta)} u_j^2 \,du \\ &= \frac{1}{2\bigl|B(0, \delta)\bigr|} \sum_{i=1}^q \lambda_i \frac{1}{q} \int_{B(0, \delta)} |u|^2 \,du \\ &= \tr H_{\phi_h}(s^{\mathstrut*}) \cdot \frac{1}{2q\bigl|B(0, \delta)\bigr|} \int_{B(0, \delta)} |u|^2 \,du \end{align*} where $\tr H_{\phi_h} = \Delta \phi_h$ is the trace of the Hessian. Here we used the fact that the value $\int_{B(0, \delta)} u_i^2 \,du$ does not depend on $i$ and thus equals the average $\frac{1}{q} \sum_{j=1}^q \int_{B(0, \delta)} u_j^2 \,du$. Rescaling space by a factor $1/\delta$ and using the relation $\int_{B(0,1)} |x|^2 \,dx = \bigl|B(0,1)\bigr| \cdot q/(q+2)$ we find \begin{align*} \hskip5mm&\hskip-5mm \frac{1}{2\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} (s - s^{\mathstrut*})^\top H_{\phi_h}(s^{\mathstrut*}) (s-s^{\mathstrut*}) \,ds \\ &= \Delta\phi_h(s^{\mathstrut*}) \cdot \frac{1}{2q \delta^q \bigl|B(0, 1)\bigr|} \int_{B(0, 1)} |y|^2 \delta^{q+2} \,dy \\ &= \frac{\Delta\phi_h(s^{\mathstrut*})}{2(q+2)} \delta^2. \end{align*} For the error term we can use a similar scaling argument in the bound~\eqref{eq:error-bound} to get \begin{equation*} \frac{1}{\bigl|B(s^{\mathstrut*}, \delta)\bigr|} \int_{B(s^{\mathstrut*}, \delta)} \bigl| r_3(s - s^{\mathstrut*}) \bigr| \,ds \leq C \cdot \delta^3 \end{equation*} for some constant~$C$. Substituting these results back into equation~\eqref{eq:f-delta-1} we find \begin{equation}\label{eq:f-delta-2} \phi^{(\delta)}_h(s^{\mathstrut*}) = \phi_h(s^{\mathstrut*}) + a_h(s^{\mathstrut*}) \cdot \delta^2 + \mathcal{O}(\delta^3) \end{equation} where \begin{equation*} a_h(s^{\mathstrut*}) = \frac{\Delta\phi_h(s^{\mathstrut*})}{2(q+2)}. \end{equation*} Using formula~\eqref{eq:f-delta-2} for $h\equiv 1$ we also get \begin{equation*} \phi^{(\delta)}_1(s^{\mathstrut*}) = \phi_1(s^{\mathstrut*}) + a_1(s^{\mathstrut*}) \cdot \delta^2 + \mathcal{O}(\delta^3) \end{equation*} as $\delta\downto 0$. Using representation~\eqref{eq:bias} of the bias, and omitting the argument $s^{\mathstrut*}$ for brevity, we can now express the bias in powers of~$\delta$: \begin{align*} \bias(Y^{(\delta)}_n) &= \frac{\phi_1\phi^{(\delta)}_h - \phi^{(\delta)}_1\phi_h}{\phi^{(\delta)}_1\phi_1} \\ &= \frac{\phi_1 \Bigl( \phi_h + a_h \delta^2 + \mathcal{O}(\delta^3) \Bigr) - \Bigl( \phi_1 + a_1 \delta^2 + \mathcal{O}(\delta^3) \Bigr)\phi_h}{\phi^{(\delta)}_1\phi_1} \\ &= \frac{\phi_1 a_h - \phi_h a_1}{\phi^{(\delta)}_1\phi_1} \cdot \delta^2 + \mathcal{O}(\delta^3). \end{align*} Since $1/\phi^{(\delta)}_1 = 1/\phi_1 \cdot \bigl(1 + \mathcal{O}(\delta^2)\bigr)$ and $y = \phi_h / \phi_1$, the right-hand side can be simplified to \begin{equation*} \bias(Y^{(\delta)}_n) = \frac{a_h(s^{\mathstrut*}) - y a_1(s^{\mathstrut*})}{\phi_1(s^{\mathstrut*})} \cdot \delta^2 + \mathcal{O}(\delta^3). \end{equation*} This completes the proof. \end{proof} To prove the statement of theorem~\ref{t:rate}, the bias of $Y^{(\delta)}_n$ has to be balanced with the computational cost for computing $Y^{(\delta)}_n$. When $\delta$ decreases, fewer samples will satisfy the acceptance condition $\| S(X) - s^{\mathstrut*} \| \leq \delta$ and the running time of the algorithm will increase. The following lemma makes this statement precise. \begin{lemma}\label{l:cost} Let $f_S$ be continuous at~$s^{\mathstrut*}$. Then the expected computational cost for computing the estimate $Y^{(\delta)}_n$ satisfies \begin{equation*} \mathbb{E}\bigl(\cost(Y^{(\delta)}_n)\bigr) = c_1 + c_2 n \delta^{-q} \bigl( 1 + c_3(\delta) \bigr) \end{equation*} for all $n\in\mathbb{N}$, $\delta > 0$ where $c_1$ and $c_2$ are constants, and $c_3$ does not depend on~$n$ and satisfies $c_3(\delta) \to 0$ as $\delta \downto 0$. \end{lemma} \begin{proof} The computational cost for algorithm~\ref{alg:ABC} is of the form $a + bN$ where $a$ and $b$ are constants, and $N$ is the random number of iterations of the loop required until $n$ samples are accepted. The number of iterations required to generate one ABC sample is geometrically distributed with parameter \begin{equation*} p = P\bigl(S \in B(s^{\mathstrut*}, \delta)\bigr) \end{equation*} and thus $N$, being the sum of $n$ independent geometrically distributed values, has mean $\mathbb{E}(N) = n / p$. The probability~$p$ can be written as \begin{equation*} p = \int_{B(s^{\mathstrut*}, \delta)} f_S(s) \,ds = \bigl| B(s^{\mathstrut*}, \delta) \bigr| \cdot \phi_1^{(\delta)}(s^{\mathstrut*}) = \delta^q \bigl| B(s^{\mathstrut*}, 1) \bigr| \cdot \phi_1^{(\delta)}(s^{\mathstrut*}). \end{equation*} Since $\phi_1 = f_S$ is continuous at~$s^{\mathstrut*}$, we have $\phi^{(\delta)}_1(s^{\mathstrut*}) \to \phi_1(s^{\mathstrut*})$ as $\delta \downto 0$ and thus $p = c \delta^q \bigl( 1 + o(1) \bigr)$ for some constant~$c$. Thus, we find that the computational cost satisfies \begin{align*} \mathbb{E}\bigl(\cost(Y^{(\delta)}_n)\bigr) &= a + b \cdot \mathbb{E}(N) \\ &= a + b \cdot \frac{n}{c \delta^q \bigl( 1 + o(1) \bigr)} \\ &= a + \frac{b}{c} \, n \delta^{-q} \, \frac{1}{1 + o(1)}, \end{align*} and the proof is complete. \end{proof} Finally, the following two lemmata give the relation between the approximation error caused by the bias on the one hand and the expected computational cost on the other hand. \begin{lemma}\label{l:optimal} Let $\delta_n \sim n^{-1/4}$ for $n\in\mathbb{N}$ . Assume that $\mathbb{E}\bigl( h(\theta)^2 \bigr) < \infty$, that $S$ satisfies assumption~A and $ \Var\condl{h(\theta)}{S = s^{\mathstrut*}} > 0$. Then, for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$, the error satisfies $\MSE(Y^{(\delta_n)}_n) \sim n^{-1}$, the expected computational cost increases as $\mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr) \sim n^{1+q/4}$ and we have \begin{equation*} \MSE(Y^{(\delta_n)}_n) \sim \mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)^{-4/(q+4)} \end{equation*} as $n\to\infty$. \end{lemma} \begin{proof} By assumption, the limit $D = \lim \delta_n n^{1/4}$ exists. Using lemma~\ref{l:bias} and corollary~\ref{C:variance}, we find \begin{align*} \lim_{n\to\infty} \frac{\MSE(Y^{(\delta_n)}_n)}{n^{-1}} &= \lim_{n\to\infty} n\Bigl( \Var(Y^{(\delta_n)}_n) + \bigl( \bias(Y^{(\delta_n)}_n) \bigr)^2 \Bigr) \\ &= \lim_{n\to\infty} n \Var(Y^{(\delta_n)}_n) + \lim_{n\to\infty} n \bigl( C(s^{\mathstrut*}) \, \delta_n^2 + O(\delta_n^3) \bigr)^2 \\ &= \Var\condl{h(\theta)}{S = s^{\mathstrut*}} + \lim_{n\to\infty} \Bigl( C(s^{\mathstrut*}) + \frac{\mathcal{O}(\delta_n^3)}{\delta_n^2} \Bigr)^2 (\delta_n n^{1/4})^4 \\ &= \Var\condl{h(\theta)}{S = s^{\mathstrut*}} + C(s^{\mathstrut*})^2 D^4. \end{align*} On the other hand, using lemma~\ref{l:cost}, we find \begin{align*} \lim_{n\to\infty} \frac{\mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)}{n^{1+q/4}} &= \lim_{n\to\infty} \frac{c_1 + c_2 n \delta_n^{-q} \bigl( 1 + c_3(\delta_n) \bigr)}{n^{1+q/4}} \\ &= 0 + c_2 \lim_{n\to\infty} \frac{1}{(\delta _n n^{1/4})^q} \bigl(1 + \lim_{n\to\infty} c_3(\delta_n) \bigr) \\ &= c_2 / D^q. \end{align*} Finally, combining the rates for cost and error we get the result \begin{equation}\label{eq:optimal} \begin{split} \hskip5mm&\hskip-5mm \lim_{n\to\infty} \frac{\!\MSE(Y^{(\delta_n)}_n)\;}{\; \mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)^{-4/(q+4)} \!} \\ &= \lim_{n\to\infty} \frac{\MSE(Y^{(\delta_n)}_n)}{n^{-1}} / \lim_{n\to\infty} \Bigl( \frac{\mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)}{n^{1+q/4}} \Bigr)^{-4/(q+4)} \\ &= \Bigl( \Var\condl{h(\theta)}{S = s^{\mathstrut*}} + C(s^{\mathstrut*})^2 D^4 \Bigr) \cdot \Bigl( \frac{D^q}{c_2} \Bigr)^{-4/(q+4)}. \end{split} \end{equation} Since the right-hand side of this equation is non-zero, this completes the proof. \end{proof} Lemma~\ref{l:optimal} only specifies the optimal {\em rate} for the decay of $\delta_n$ as a function of~$n$. By inspecting the proof, we can derive the corresponding optimal constant: $\delta_n$ should be chosen as $\delta_n = D n^{-1/4}$ where $D$ minimises the expression in equation~\eqref{eq:optimal}. The optimal value of $D$ can analytically be found to be \begin{equation*} D_{\mathrm{opt}} = \Bigl( \frac{q \Var\condl{h(\theta)} {S = s^{\mathstrut*}}}{4 C(s^{\mathstrut*})^2} \Bigr)^{1/4}, \end{equation*} where $C(s^{\mathstrut*})$ is the constant from lemma~\ref{l:bias}. The variance $\Var\condl{h(\theta)}{S = s^{\mathstrut*}}$ is easily estimated, but it seems difficult to estimate $C(s^{\mathstrut*})$ in a likelihood-free way. The following result completes the proof of theorem~\ref{t:rate} by showing that no other choice of $\delta_n$ can lead to a better rate for the error while retaining the same cost. \begin{lemma}\label{l:bound} For $n\in\mathbb{N}$ let $\delta_n \downto 0$. Assume that $\mathbb{E}\bigl( h(\theta)^2 \bigr) < \infty$ with $\Var\condl{h(\theta)}{S = s^{\mathstrut*}} > 0$, that $S$ satisfies assumption~A and $C(s^{\mathstrut*}) \neq 0$. Then, for $f_S$-almost all $s^{\mathstrut*}\in\mathbb{R}^q$, we have \begin{equation*} \liminf_{n\to\infty}\frac{\!\MSE(Y^{(\delta_n)}_n)\;}{\;\mathbb{E}\bigl(\cost(Y^{(\delta_n)}_n)\bigr)^{-4/(q+4)}\!} > 0. \end{equation*} \end{lemma} \begin{proof} From lemma~\ref{l:bias} we know that \begin{equation}\label{eq:bound1} \begin{split} \MSE(Y^{(\delta_n)}_n) &= \Var\bigl(Y^{(\delta_n)}_n\bigr) + \bigl( \bias(Y^{(\delta_n)}_n) \bigr)^2 \\ &= \frac{\Var\bigl(h(\theta^{(\delta_n)}_j)\bigr)}{n} + \bigl( C(s^{\mathstrut*}) \, \delta_n^2 + \mathcal{O}(\delta_n^3) \bigr)^2 \\ &= \frac{\Var\bigl(h(\theta^{(\delta_n)}_j)\bigr)}{n} + C(s^{\mathstrut*})^2 \, \delta_n^4 + \mathcal{O}(\delta_n^5) \\ &\geq \frac{\Var\bigl(h(\theta^{(\delta_n)}_j)\bigr)}{n} + \frac{C(s^{\mathstrut*})^2}{2} \, \delta_n^4 \end{split} \end{equation} for all sufficiently large~$n$, since $C(s^{\mathstrut*})^2 > 0$. By lemma~\ref{l:cost}, as $n\to\infty$, we have \begin{equation}\label{eq:bound2} \begin{split} \hskip5mm&\hskip-5mm \mathbb{E}\bigl( \cost(Y^{(\delta_n)}_n) \bigr)^{-4/(q+4)} \\ &\sim (n \delta_n^{-q})^{-4/(q+4)} \\ &= \bigl(n^{-1}\bigr)^{4/(q+4)} \bigl(\delta_n^4\bigr)^{q/(q+4)} \\ &\sim \Bigl(\frac{4}{q+4} \frac{\Var\bigl(h(\theta^{(\delta_n)}_j)\bigr)}{n} \Bigr)^{4/(q+4)} \Bigl(\frac{q}{q+4} \, \frac{C(s^{\mathstrut*})^2}{2} \delta_n^4 \Bigr)^{q/(q+4)} \end{split} \end{equation} where we were able to insert the constant factors into the last term because the $\sim$-relation does not see constants. Using Young's inequality we find \begin{equation}\label{eq:bound3} \begin{split} \hskip5mm&\hskip-5mm \frac{\Var\bigl(h(\theta^{(\delta_n)}_j)\bigr)}{n} + \frac{C(s^{\mathstrut*})^2}{2} \delta_n^4 \\ &\geq \Bigl(\frac{4}{q+4} \frac{\Var\bigl(h(\theta^{(\delta_n)}_j)\bigr)}{n} \Bigr)^{4/(q+4)} \Bigl(\frac{q}{q+4} \, \frac{C(s^{\mathstrut*})^2}{2} \delta_n^4 \Bigr)^{q/(q+4)} \end{split} \end{equation} and combining equations \eqref{eq:bound1}, \eqref{eq:bound2} and~\eqref{eq:bound3} we get \begin{equation*} \MSE(Y^{(\delta_n)}_n) \geq c \cdot \mathbb{E}\bigl( \cost(Y^{(\delta_n)}_n) \bigr)^{-4/(q+4)} \end{equation*} for some constant $c > 0$ and all sufficiently large~$n$. This is the required result. \end{proof} The statement of theorem~\ref{t:rate} coincides with the statements of lemmata \ref{l:optimal} and~\ref{l:bound} and thus we have completed the proof of theorem~\ref{t:rate}. \section{Fixed Number of Proposals} \label{S:fixed-N} So far, we have fixed the number $n$ of accepted samples; the number~$N$ of proposals required to generate $n$ samples is then a random variable. In this section we consider the opposite approach, where the number $\hat N$ of proposals is fixed and the number $\hat n$ of accepted samples is random. Compared to the case with fixed number of samples, the case considered here includes two additional sources of error. First, with small probability no samples at all will be accepted and we shall see that this introduces a small additional bias. Secondly, the randomness in the value of $\hat n$ introduces additional variance in the estimate. We show that the additional error introduced is small enough not to affect the rate of convergence. The main result here is corollary~\ref{c:rate-hat}, below, which shows that the results from theorem~\ref{t:rate} still hold for random~$\hat n$. In analogy to the definition of $Y^{(\delta)}_n$ from equation~\eqref{eq:ABC-MC-basic}, we define \begin{equation*} \hat Y^{(\delta)}_{\hat N} = \begin{cases} \frac{1}{\hat n} \sum_{j=1}^{\hat n} h\bigl(\theta_j^{(\delta)}\bigr), & \mbox{if $\hat n > 0$, and} \\ \hat c & \mbox{if $\hat n = 0$,} \end{cases} \end{equation*} where $\hat c$ is an arbitrary constant (\textit{e.g.}\ zero or the prior mean) to be used if none of the proposals are accepted. For a given tolerance $\delta > 0$, each of the $\hat N$ proposals is accepted with probability \begin{equation*} p = P\bigl(S(X) \in B(s^{\mathstrut*}, \delta)\bigr), \end{equation*} and the number of accepted samples is binomially distributed with mean $\mathbb{E}(\hat n) = \hat N p$. In this section we consider arbitrary sequences of $\hat N$ and $\delta$ such that $\hat N p \to \infty$ and we show that asymptotically the estimators $Y^{(\delta)}_{\lfloor\hat N p\rfloor}$ (using $n = \lfloor \hat N p \rfloor$ accepted samples) and $\hat Y^{(\delta)}_{\hat N}$ (using $\hat N$ proposals) have the same error. \begin{prop}\label{p:MSE-fixed-N} Let $\hat Y^{(\delta)}_{\hat N}$ and $Y^{(\delta)}_{\lfloor\hat N p\rfloor}$ be as above. Then \begin{equation*} \frac{\MSE\bigl(\hat Y^{(\delta)}_{\hat N}\bigr)}{\MSE\bigl(Y^{(\delta)}_{\lfloor\hat N p\rfloor}\bigr)} \longrightarrow 1 \end{equation*} as $\hat N p \to \infty$. \end{prop} \begin{proof} We start the proof by comparing the variances of the two estimates $\hat Y^{(\delta)}_{\hat N}$ and $Y^{(\delta)}_{\lfloor\hat N p\rfloor}$. For the estimate with a fixed number of proposals we find \begin{align*} \Var\bigl(\hat Y^{(\delta)}_{\hat N}\bigr) &= \mathbb{E}\Bigl(\Var\condl{\hat Y^{(\delta)}_{\hat N}}{\hat n}\Bigr) + \Var\Bigl(\mathbb{E}\condl{\hat Y^{(\delta)}_{\hat N}}{\hat n}\Bigr) \\ &= \mathbb{E}\Bigl(\frac{\sigma_\delta^2}{\hat n} 1_{\{\hat n > 0\}}\Bigr) + (\hat c - y_\delta)^2 P(\hat n > 0) P(\hat n = 0), \end{align*} where $\sigma_\delta^2 = \Var\condl{h(\theta)}{S\in B(s^{\mathstrut*},\delta)}$ and $y_\delta = \mathbb{E}\condl{h(\theta)}{S\in B(s^{\mathstrut*},\delta)}$. Thus, \begin{equation}\label{eq:Nhat-Var} \begin{split} \frac{\Var\bigl(\hat Y^{(\delta)}_{\hat N}\bigr)}{\Var\bigl(Y^{(\delta)}_{\lfloor\hat N p\rfloor}\bigr)} &= \frac{\mathbb{E}\Bigl(\frac{\sigma_\delta^2}{\hat n} 1_{\{\hat n > 0\}}\Bigr) + (\hat c - y_\delta)^2 P(\hat n > 0) P(\hat n = 0)}% {\frac{\sigma_\delta^2}{\lfloor\hat N p\rfloor}} \\ &\leq \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{\hat n > 0\}}\Bigr) + \frac{(\hat c - y_\delta)^2}{\sigma_\delta^2} \lfloor\hat N p\rfloor (1-p)^{\hat N}. \end{split} \end{equation} Our aim is to show that the right-hand side of this equation converges to~$1$. To see this, we split the right-hand side into four different terms. First, since $\log\bigl( (1-p)^{\hat N}\bigr) = \hat N \log(1-p) \leq -\hat N p$, we have $(1-p)^{\hat N} \leq \exp(-\hat N p)$ and thus \begin{equation}\label{eq:Nhat-term0} \lfloor\hat N p\rfloor (1-p)^{\hat N} \to 0, \end{equation} \textit{i.e.}\ the right-most term in \eqref{eq:Nhat-Var} disappears as $\hat N p \to\infty$. Now let $\eps > 0$. Then we have \begin{equation}\label{eq:Nhat-term1} \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{\hat n > (1+\eps)\hat N p\}}\Bigr) \leq \frac{1}{1+\eps} \cdot P\Bigl(\frac{\hat n}{\hat N p} > 1+\eps\Bigr) \\ \leq \frac{1}{\eps^2} \Var\Bigl(\frac{\hat n}{\hat N p}\Bigr) \to 0 \end{equation} as $\hat N p\to\infty$ by Chebyshev's inequality. For the lower tails of $\hat n / \hat N p$ we find \begin{align*} \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{0 < \hat n \leq (1-\eps)\hat N p\}}\Bigr) &\leq \hat N p P\Bigl( 0 < \hat n \leq (1-\eps)\hat N p \Bigr) \\ &\leq (1-\eps) (\hat N p)^2 P\Bigl( \hat n = \bigl\lfloor (1-\eps)\hat N p \bigr\rfloor \Bigr). \end{align*} Using the relative entropy \begin{equation*} H\cond{q}{p} = q \log\Bigl(\frac{q}{p}\Bigr) + (1-q)\log\Bigl(\frac{1-q}{1-p}\Bigr), \end{equation*} lemma~2.1.9 in~\citet{DeZei98} states $P(\hat n = k) \leq \exp\bigl(-\hat N H\cond{k/\hat N}{p}\bigr)$ and since $q \mapsto H\cond{q}{p}$ is decreasing for $q < p$ we get \begin{equation*} \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{0 < \hat n \leq (1-\eps)\hat N p\}}\Bigr) \leq (1-\eps) (\hat N p)^2 \exp\Bigl(-\hat N \cdot H\condl{(1-\eps)p}{p}\Bigr). \end{equation*} It is easy to show that $H\condl{(1-\eps)p}{p} / p > \eps + (1-\eps)\log(1-\eps) > 0$ and thus \begin{equation} \begin{split} \hskip5mm&\hskip-5mm \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{0 < \hat n \leq (1-\eps)\hat N p\}}\Bigr) \\ &\leq (1-\eps) (\hat N p)^2 \exp\Bigl(-\hat N p \bigl( \eps + (1-\eps)\log(1-\eps) \bigr) \Bigr) \longrightarrow 0 \end{split} \end{equation} as $\hat Np \to \infty$. Finally, we find \begin{equation}\label{eq:Nhat-term3a} \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{(1-\eps)\hat N p < \hat n \leq (1+\eps)\hat N p\}}\Bigr) \leq \frac{1}{1-\eps} P\Bigl( \bigl| \frac{\hat n}{\hat Np} - 1 \bigr| \leq \eps \Bigr) \longrightarrow \frac{1}{1-\eps} \end{equation} and \begin{equation}\label{eq:Nhat-term3b} \mathbb{E}\Bigl(\frac{\hat N p}{\hat n} 1_{\{(1-\eps)\hat N p \leq \hat n \leq (1+\eps)\hat N p\}}\Bigr) \geq \frac{1}{1+\eps} P\Bigl( \bigl| \frac{\hat n}{\hat Np} - 1 \bigr| \leq \eps \Bigr) \longrightarrow \frac{1}{1+\eps} \end{equation} as $\hat N p\to \infty$. Substituting equations \eqref{eq:Nhat-term0} to~\eqref{eq:Nhat-term3b} into \eqref{eq:Nhat-Var} and letting $\eps \downto0$ we finally get \begin{equation*} \frac{\Var\bigl(\hat Y^{(\delta)}_{\hat N}\bigr)}{\Var\bigl(Y^{(\delta)}_{\lfloor\hat N p\rfloor}\bigr)} \longrightarrow 1 \end{equation*} as $\hat N p\to \infty$. From the previous sections we know $\bias(Y^{(\delta)}_n) = y_\delta - y$. Since \begin{equation*} \bias(\hat Y^{(\delta)}_{\hat N}) = \mathbb{E}\bigl( \mathbb{E}\cond{\hat Y^{(\delta)}_{\hat N}}{\hat n} - y \bigr) = (y_\delta - y) P(\hat n > 0) + (\hat c - y) P(\hat n = 0), \end{equation*} and since we already have seen $P(\hat n = 0) = (1-p)^{\hat N} = o(1/\hat Np)$, we find \begin{equation}\label{eq:Nhat-MSE1} \begin{split} \frac{\MSE(\hat Y^{(\delta)}_{\hat N})}{\MSE(Y^{(\delta)}_{\lfloor\hat N p\rfloor})} &= \frac{\Var(\hat Y^{(\delta)}_{\hat N}) + \bias(\hat Y^{(\delta)}_{\hat N})^2}{\Var(Y^{(\delta)}_{\lfloor\hat N p\rfloor}) + \bias(Y^{(\delta)}_{\lfloor\hat N p\rfloor})^2} \\ &= \frac{\frac{\sigma_\delta^2}{\lfloor \hat N p\rfloor}\bigl(1+o(1)\bigr) + (y_\delta - y)^2 \bigl(1 + o(1)\bigr) + o( 1 / \hat N p)} {\frac{\sigma_\delta^2}{\lfloor \hat N p\rfloor} + (y_\delta - y)^2} \\ &= \frac{\sigma_\delta^2\bigl(1+o(1)\bigr) + \lfloor \hat N p\rfloor(y_\delta - y)^2 \bigl(1 + o(1)\bigr) + o(1)} {\sigma_\delta^2 + \lfloor \hat N p\rfloor(y_\delta - y)^2}. \end{split} \end{equation} Let $\eps > 0$ and let $\hat N p$ be large enough that all $o(1)$ terms in~\eqref{eq:Nhat-MSE1} satisfy $-\eps < o(1) < \eps$. Then we have \begin{equation*} \frac{\MSE(\hat Y^{(\delta)}_{\hat N})}{\MSE(Y^{(\delta)}_{\lfloor\hat N p\rfloor})} \leq \frac{\sigma_\delta^2\bigl(1+\eps\bigr) + \lfloor \hat N p\rfloor(y_\delta - y)^2 \bigl(1 + \eps\bigr) + \eps} {\sigma_\delta^2 + \lfloor \hat N p\rfloor(y_\delta - y)^2} \leq \frac{\sigma_\delta^2(1+\eps) + \eps}{\sigma_\delta^2}, \end{equation*} where the second inequality is found by maximising the function $x \mapsto \bigl(\sigma_\delta^2(1+\eps) + x(1+\eps) + \eps\bigr)/(x+\sigma_\delta^2)$. Similarly, we find \begin{equation*} \frac{\MSE(\hat Y^{(\delta)}_{\hat N})}{\MSE(Y^{(\delta)}_{\lfloor\hat N p\rfloor})} \geq \frac{\sigma_\delta^2\bigl(1-\eps\bigr) + \lfloor \hat N p\rfloor(y_\delta - y)^2 \bigl(1 - \eps\bigr) + \eps} {\sigma_\delta^2 + \lfloor \hat N p\rfloor(y_\delta - y)^2} \geq \frac{\sigma_\delta^2(1-\eps) - \eps}{\sigma_\delta^2} \end{equation*} for all sufficiently large $\hat N p$. Using these bounds, taking the limit $\hat N p \to \infty$ and finally taking the limit $\eps \downto 0$ we find \begin{equation*} \frac{\MSE(\hat Y^{(\delta)}_{\hat N})}{\MSE(Y^{(\delta)}_{\lfloor\hat N p\rfloor})} \longrightarrow 1 \end{equation*} as $\hat N p \to \infty$. This completes the proof. \end{proof} We note that the result of proposition~\ref{p:MSE-fixed-N} does not require $p$ to converge to~$0$. In this case neither $\MSE\bigl(\hat Y^{(\delta)}_{\hat N}\bigr)$ nor $\MSE\bigl(Y^{(\delta)}_{\lfloor\hat N p\rfloor}\bigr)$ converges to~$0$, but the proposition still shows that the difference between fixed $N$ and fixed $n$ is asyptotically negligible. \begin{corollary}\label{c:rate-hat} Let the assumptions of theorem~\ref{t:rate} hold and let $(\delta_{\hat N})_{\hat N\in\mathbb{N}}$ be a sequence with $\delta_{\hat N} \sim \hat N^{-1/(q+4)}$, where $\hat N$ denotes the (fixed) number of proposals. Then the mean squared error satisfies \begin{equation*} \MSE\bigl(\hat Y_{\hat N}^{\delta_{\hat N}}\bigr) \sim \cost\bigl(\hat Y_{\hat N}^{\delta_{\hat N}}\bigr)^{-4/(q+4)} \end{equation*} as $\hat N\to\infty$, and the given choice of $\delta_{\hat N}$ is optimal in the sense of theorem~\ref{t:rate}. \end{corollary} \begin{proof} Let $p_{\hat N} = P\bigl(S(X) \in B(s^{\mathstrut*}, \delta_{\hat N})\bigr)$. Then \begin{equation*} \hat N p_{\hat N} \sim \hat N \delta_{\hat N}^q \sim \hat N ^ {1 - q/(q+4)} = \hat N ^ {4/(q+4)} \to \infty \end{equation*} and thus we can apply proposition~\ref{p:MSE-fixed-N}. Similarly, we have \begin{equation*} \delta_{\hat N} \sim \hat N ^ {- 1/(q+4)} \sim (\hat N p_{\hat N})^{-1/4} \sim \lfloor\hat N p_{\hat N}\rfloor^{-1/4} \end{equation*} and thus we can apply theorem~\ref{t:rate}. Using these two results we get \begin{equation*} \MSE\bigl( \hat Y_{\hat N}^{\delta_{\hat N}} \bigr) \sim \MSE\bigl( Y^{(\delta_{\hat N})}_{\lfloor\hat N p_{\hat N}\rfloor} \bigr) \sim \mathbb{E}\Bigl(\cost\bigl( Y^{(\delta_{\hat N})}_{\lfloor\hat N p_{\hat N}\rfloor} \bigr)\Bigr)^{-4/(q+4)}. \end{equation*} Since $\hat Y_{\hat N}^{\delta_{\hat N}}$ is always computed using $\hat N$ proposals we have \begin{equation*} \cost\Bigl(\hat Y_{\hat N}^{\delta_{\hat N}}\Bigr) \sim \hat N \end{equation*} and thus, using lemma~\ref{l:cost}, we find \begin{equation*} \mathbb{E}\Bigl(\cost\bigl( Y^{(\delta_{\hat N})}_{\lfloor\hat N p_{\hat N}\rfloor} \bigr)\Bigr) \sim \bigl\lfloor\hat N p_{\hat N}\bigr\rfloor \delta_{\hat N}^{-q} \sim \hat N \sim \cost\Bigl(\hat Y_{\hat N}^{\delta_{\hat N}}\Bigr). \end{equation*} This completes the proof of the first statement of the corollary. For the second statement there are two cases. If $\delta_{\hat N}$ is chosen such that $\hat N p_{\hat N} \to \infty$ for a sub-sequence, the statement is a direct consequence of proposition~\ref{p:MSE-fixed-N}. Finally, if $\hat N p_{\hat N}$ is bounded, $\MSE\bigl( \hat Y_{\hat N}^{\delta_{\hat N}}\bigr)$ stays bounded away from~$0$ and thus \begin{equation*} \frac{\MSE\bigl(\hat Y_{\hat N}^{\delta_{\hat N}}\bigr)} {\cost\bigl(\hat Y_{\hat N}^{\delta_{\hat N}}\bigr)^{-4/(q+4)}} \longrightarrow \infty > 0 \end{equation*} as $\hat N\to \infty$. This completes the proof of the second statement. \end{proof} \section{Numerical Experiments} \label{S:experiments} To illustrate the results from section~\ref{S:results}, we present a series of numerical experiments for the following toy problem. \begin{itemize} \item We choose $p=1$, and assume that our prior belief in the value of the single parameter $\theta$ is standard normally distributed. \item We choose $d=2$, and assume the data $X$ to be composed of i.i.d.\ samples $X_1, X_2$ each with distribution $X_i \,|\, \theta \sim \mathcal{N}(\theta , 1)$. \item We choose $q=2$, and the (non-minimal) sufficient statistic to be $S(x) = x$ for all $x\in\mathbb{R}^2$. \item We consider the test function $h(\theta)=1_{[-1/2,1/2]}(\theta)$, \textit{i.e.}\ the indicator function for the region $[-1/2,1/2]$. The ABC estimate is thus an estimate for the posterior probability $P\bigl( \theta\in [-1/2,1/2] \bigm| S=s^{\mathstrut*} \bigr)$. \item We fix the observed data to be $s^*=(1,1)$. \end{itemize} This problem is simple enough that all quantities of interest can be determined explicitly. In particular, $\theta|S$ is $\mathcal{N}\bigl((s_1+s_2)/3, 1/3\bigr)$ distributed, $\theta|S\!=\!s^*$ is $\mathcal{N}(2/3, 1/3)$ distributed, and $S$ is bivariate normally distributed with mean~$0$ and covariance matrix \begin{equation*} \Sigma = \begin{pmatrix}2&1\\1&2\end{pmatrix}. \end{equation*} Therefore, the prior and posterior expectation for $h(\theta)$ are $\mathbb{E}\bigl(h(\theta)\bigr) = 0.3829$ and $\mathbb{E}\condl{h(\theta)}{S=s^{\mathstrut*}} = 0.3648$, respectively. Similarly, the constant from lemma~\ref{l:bias} can be shown to be $C(s^*) = 0.0323$. Assumption~A can be shown to hold. The function \begin{equation*} \phi_1(s) = f_S(s) = \frac{1}{2\pi\sqrt{3}} e^{-\frac{1}{3} (s_1^2-s_1s_2+s_2^2)} \end{equation*} is multivariate normal, so its third derivatives exist, and are bounded and continuous. Similarly, the function \begin{equation*} \phi_h(s) = \int_{-1/2}^{1/2} f_{\theta,S}(t,s) dt \leq \phi_1(s) \end{equation*} also has bounded and continuous third derivatives. Thus, the assumptions hold. The programs used to perform the simulations described in this section, written using the R environment \citep{R} are available as supplementary material. \subsection*{Experiment 1} Our first experiment validates the statement about the bias given in lemma~\ref{l:bias}. Since we know the exact posterior expectation, the bias can be determined experimentally. For fixed $\delta$, we generate $k$ independent ABC estimates, each based on $n$ proposals. For each of the $k$ estimates, we calculate its distance from the true posterior expectation. We then calculate the mean and standard error of these differences to obtain a Monte Carlo estimate of the bias. Repeating this procedure for several values of $\delta$, we can produce a plot of the estimated bias against $\delta$, with $95\%$ error bars. Figure~\ref{fig:parabola} shows the result of a simulation, using $n=500$ samples for each ABC estimate and $k=5000$ ABC estimates for each value of~$\delta$. For comparison, the plot includes the theoretically predicted asymptotic bias $C(s^{\mathstrut*}) \delta^2$, using the value~$C(s^*) = 0.0323$. The plot shows that for small values of $\delta$ the theoretical curve is indeed a good fit to the numerical estimates of the bias. The result of the lemma is only valid as $\delta \downto 0$ and indeed the plot shows a discrepancy for larger values. This discrepancy is caused by only a small fraction of the sample being rejected; as $\delta$ increases, the distribution of the ABC samples approaches the prior distribution. \begin{figure} \begin{center} \includegraphics{fig1.pdf} \end{center} \caption{\label{fig:parabola}Simulation results illustrating the relationship between bias and $\delta$. Circles give the mean empirical bias from $5000$ simulations for each value of~$\delta$. The error bars indicate mean $\pm 1.96$ standard errors. The solid line shows the theoretically predicted asymptotic bias from lemma~\ref{l:bias}.} \end{figure} \subsection*{Experiment 2} Our second experiment validates the statement of theorem~\ref{t:rate}, by numerically estimating the optimal choice of delta and the corresponding MSE. For fixed values of expected computational cost and $\delta$, we estimate the mean squared error by generating $k$ different ABC estimates and taking the mean of their squared distance from the true posterior expectation. This reflects how the bias is estimated in experiment~1. Repeating this procedure for several values of $\delta$, the estimates of the MSE are plotted against~$\delta$. Our aim is to determine the optimal value of $\delta$ for fixed computational cost. From lemma~\ref{l:cost} we know that the expected cost is of order $n \delta^{-q}$ and thus we choose $n \sim \delta^2$ in this example. From lemma~\ref{l:bias} we know that $\bias \sim \delta^2$. Thus, we expect the MSE for constant expected cost to be of the form \begin{equation}\label{eq:fitted} \MSE(\delta) = \frac{\Var}{n} + \bias^2 = a\delta^{-2} +b\delta^4 \end{equation} for some constants $a$ and~$b$. Thus, we fit a curve of this form to the numerically estimated values of the MSE. The result of one such simulation, using $k=500$ samples for each~$\delta$, is shown in figure~\ref{fig:opt-delta}. The curve fits the data well. \begin{figure} \begin{center} \includegraphics{fig2.pdf} \end{center} \caption{\label{fig:opt-delta}The estimated MSE as a function of the tolerance~$\delta$ for fixed expected cost. The fitted curve has the form given in equation~\eqref{eq:fitted}. The location of the optimal~$\delta$ is marked by the vertical line.} \end{figure} We estimate the optimal values of $\delta$ and MSE, given an expected computational cost, to be those at the minimum of the fitted curve. Given the good fit between the predicted form~\eqref{eq:fitted} of the curve and the empirical MSE values, this procedure promises to be a robust way to estimate the optimal value of $\delta$. The direct approach would likely require a much larger number of samples to be accurate. Repeating the above procedure for a range of values of expected cost gives corresponding estimates for the optimal values of $\delta$ and the MSE as a function of expected cost. We expect the optimal $\delta$ and the MSE to depend on the cost like $x=A \cdot \mathrm{cost}^B$. To validate the statements of theorem~\ref{t:rate} we numerically estimate the exponent~$B$ from simulated data. The result of such a simulation is shown in figure~\ref{fig:rates}. The data are roughly on straight lines, as expected, and the gradients are close to the theoretical gradients, shown as smaller lines. The numerical results for estimating the exponent~$B$ are given in the following table. \begin{center} \begin{tabular}{r|c|c|c} Plot&Gradient&Standard error&Theoretical gradient \\ \hline $\delta$ &$-0.167$ &$0.0036$ &$-1/6 \approx -0.167$ \\ MSE &$-0.671$ &$0.0119$ &$-2/3 \approx -0.667$ \end{tabular} \end{center} The table shows an an excellent fit between empirical and theoretically predicted values. \begin{figure} \begin{center} \includegraphics{fig3.pdf} \end{center} \caption{\label{fig:rates}Estimated dependency of the optimal $\delta$ and of the corresponding MSE on the computational cost. The computational cost is given in arbitrary units, chosen such that the smallest sample size under consideration has cost~$1$. For comparison, the additional line above the fit has the gradient expected from the theoretical results.} \end{figure} \section{Discussion} \label{S:discussion} While the work of this article is mostly theoretical in nature, the results can be used to guide the design of an analysis using ABC methods. Our results can be applied directly in cases where a pilot run is used to tune the algorithm. The performance of the ABC algorithm depends both on the number~$n$ of samples accepted and on the tolerance~$\delta$. From theorem~\ref{t:rate} we know that optimality is achieved by choosing $\delta$ proportional to $n^{-1/4}$. Consequently, if the full run is performed by increasing the number of accepted ABC samples by a factor of~$k$, then the tolerance~$\delta$ for the full run should be divided by~$k^{1/4}$. In this case the expected running time satisfies \begin{equation*} \mathrm{cost} \sim n\delta^{-q} \sim n^{(q+4)/4} \end{equation*} and, using lemma~\ref{l:cost}, we have \begin{equation*} \mathrm{error} \sim \mathrm{cost}^{-2 / (q+4)} \sim n^{-1/2}. \end{equation*} There are two possible scenarios: \begin{itemize} \item If we want to reduce the root mean-squared error by a factor of~$\alpha$, we should increase the number $n$ of accepted ABC samples by $\alpha^2$ and reduce the tolerance~$\delta$ by a factor of~$(a^2)^{1/4} = \sqrt{\alpha}$. These changes will increase the expected running time of the algorithm by a factor of~$\alpha^{(q+4)/2}$. \item If we plan to increase the (expected) running time by a factor of~$\beta$, we should increase the number of accepted samples by a factor of~$\beta^{4/(q+4)}$ and divide $\delta$ by~$\beta^{1/(q+4)}$. These changes will reduce the root mean squared error by a factor of~$\beta^{2/(q+4)}$. \end{itemize} These guidelines will lead to a choice of parameters which, at least asymptotically, maximises the efficiency of the analysis. \vskip2\baselineskip \noindent \textbf{Acknowledgements.} The authors thank Alexander Veretennikov for pointing out that the Lebesgue differentiation theorem can be used in the proof of proposition~\ref{p:convergence}. MW was supported by an EPSRC Doctoral Training Grant at the School of Mathematics, University of Leeds. \bibliographystyle{plainnat}
\section{Introduction} The Smoluchowski model is widely used in the theory of diffusion-influenced reactions \cite{smoluchowski:1917, Rice:1985}. According to this picture, a pair of molecules separated by a distance $r$ may react when they encounter each other at a critical distance $r=a$ via their diffusive motion. Hence, reactive molecules can be modeled by solutions of the diffusion equation that satisfy certain types of boundary conditions (BC) at the encounter distance $r=a$. In the case of an isolated pair, exact expressions for Green's functions (GF) in the time domain, describing irreversible and reversible reactions in one, two and three space dimensions, have been obtained \cite{carslaw1986conduction, Agmon:1984, TPMMS_2012JCP, kimShin:1999}. However, there are alternative approaches to describe the reversible diffusion-influenced reaction of an isolated pair. Ref.~\cite{Khokhlova:2012} discussed the so-called volume reactivity model that eliminates the distinct role of the encounter radius $r=a$ and instead postulates that the reaction can happen throughout the spherical volume $r\leq a$. In the present manuscript, we discuss the corresponding model in two dimensions (2D) and hence refer to it as the "area reactivity" model. Diffusion in 2D is special from both a conceptual and technical point of view. Conceptually, it is the critical dimension regarding recurrence and transience of random walks \cite{Toussaint_Wilczek_1983}. Technically, the mathematical treatment appears to be more involved than in 1D and 3D \cite{TPMMS_2012JCP}. A system of two molecules $A$ and $B$ with diffusion constants $D_{A}$ and $D_{B}$, respectively, can also be described as the diffusion of a point-like molecule with diffusion constant $D=D_{A}+D_{B}$ around a static disk. More precisely, the area-reactivity model assumes that the molecule undergoes free diffusion apart from inside the static "reaction disk" of radius $r = a$, where it may react reversibly. Without loss of generality, we assume that the disk's center is located at the origin. A central notion is the probability density function (PDF) $p(r,t\vert r_{0})$ that gives the probability to find the molecule unbound at a distance equal to $r$ at time $t$, given that the distance was initially $r_{0}$ at time $t=0$. Note that in contrast to the contact reactivity model, $p(r,t\vert r_{0})$ is also defined for $r<a$. Moreover, because the molecule may bind anywhere within the disk $r<a$, it makes sense to define another PDF $q(r,t\vert r_{0})$, which yields the probability to find the molecule bound at a distance equal to $r<a$ at time $t$, given that the distance was initially $r_{0}$ at time $t=0$. The rates for association and dissociation are $\kappa_{r} \Theta(a-r)p(r,t\vert r_{0})$ and $\kappa_{d}q(r,t\vert r_{0})$, respectively, where $\Theta(x)$ refers to the Heaviside step-function that vanishes for $x<0$ and assumes unity otherwise. Furthermore, it is assumed that the dissociated molecule is released at the same point where it assumed its bound state. The equations of motion for the PDF $p(r,t\vert r_{0})$ and $q(r,t\vert r_{0})$ are coupled and read \cite{Khokhlova:2012} \begin{eqnarray} \frac{\partial p(r,t\vert r_{0})}{\partial t} &=& \mathcal{L}_{r}p(r , t\vert r_{0}) - \kappa_{r}\Theta(a-r)p(r,t\vert r_{0}) + \kappa_{d}q(r,t\vert r_{0}), \quad\quad\label{eqMotion1}\\ \frac{\partial q(r,t\vert r_{0})}{\partial t} &=& \kappa_{r}\Theta(a-r)p(r,t\vert r_{0}) -\kappa_{d}q(r,t\vert r_{0}), \label{eqMotion2} \end{eqnarray} where \begin{equation} \mathcal{L}_{r} = D\biggl(\frac{\partial^{2}}{\partial r^{2}}+\frac{1}{r}\frac{\partial}{\partial r}\biggr). \end{equation} The equations of motion have to be supplemented by BC at the origin and at infinity, respectively, \begin{eqnarray} \lim_{r\rightarrow\infty}p(r,t\vert r_{0}) &=& 0,\label{BC_Infinity}\\ \lim_{r\rightarrow 0} r \frac{\partial p(r,t\vert r_{0})}{\partial r} &=& 0 \label{BC_Zero}. \end{eqnarray} In the present manuscript, we focus on the case of the initially unbound molecule. Therefore, the initial conditions are \begin{eqnarray} 2\pi r_{0} p(r, 0\vert r_{0})&=&\delta(r-r_{0}), \label{initial_bc}\\ q(r, 0\vert r_{0}) &=& 0. \label{initial_bc_q} \end{eqnarray} \section{Exact Green's function in the Laplace domain} By applying the Laplace transform, Eqs.~(\ref{eqMotion1})-(\ref{eqMotion2}) become \begin{eqnarray} s \tilde{p}(r,s\vert r_{0}) - p(r,0\vert r_{0}) &=& \mathcal{L}_{r}\tilde{p}(r , s \vert r_{0}) - \kappa_{r}\Theta(a-r)\tilde{p}(r,s\vert r_{0}) \label{p_eqMotionLaplace}\nonumber\\ &&+ \kappa_{d}\tilde{q}(r,s\vert r_{0}),\\ s \tilde{q}(r,s\vert r_{0}) - q(r,0\vert r_{0})&=& \kappa_{r}\Theta(a-r)\tilde{p}(r,s\vert r_{0}) -\kappa_{d}\tilde{q}(r,s\vert r_{0}), \label{q_eqMotionLaplace} \end{eqnarray} where $s$ denotes the Laplace space variable. We use Eq.~(\ref{initial_bc_q}) to obtain from Eq.~(\ref{q_eqMotionLaplace}) \begin{equation}\label{q_p_Laplace} \tilde{q}(r,s\vert r_{0}) = \frac{\kappa_{r}}{s+\kappa_{d}}\Theta(a-r)\tilde{p}(r,s\vert r_{0}). \end{equation} Now we can eliminate $\tilde{q}(r,s\vert r_{0})$ from Eq.~(\ref{p_eqMotionLaplace}) \begin{equation}\label{eqMotionLaplace} \bigg[\mathcal{L}_{r} - s - \frac{s\kappa_{r}}{s+\kappa_{d}}\Theta(a-r) \bigg]\tilde{p}(r,s\vert r_{0}) = -\frac{\delta(r-r_{0})}{2\pi r}, \end{equation} where we used Eq.~(\ref{initial_bc}). In the following, we will calculate the GF separately on the two different domains defined by $r > a$ and $r<a$. The two obtained solutions will still contain unknown constants. The GF can then be completely determined by matching both expressions upon continuity requirements at $r=a$. Henceforth, we will denote the GF within $r<a$ and outside $r>a$ the reactive disk by $p^{<}(r,t\vert r_{0})$ and $p^{>}(r,t\vert r_{0})$, respectively. Also, throughout this manuscript we assume that the molecule was initially located outside the reaction area $r_{0} > a$. Then, we make the following ansatz for the Laplace transform of the GF $p^{>}(r,t\vert r_{0})$ outside the disk $r > a$, \begin{equation}\label{laplaceAnsatz} \tilde{p}^{>}(r, s \vert r_{0}) = \tilde{p}_{0}(r, s \vert r_{0}) +\tilde{f}(r, s \vert r_{0}), \end{equation} where \begin{equation}\label{laplaceFree} \tilde{p}_{0}(r, s \vert r_{0}) = \frac{1}{2\pi D} \biggl\{\begin{array}{lr} I_{0}(vr_{0}) K_{0}(vr),&\text{$ r > r_{0} $} \\ I_{0}(vr) K_{0}(qr_{0}), &\text{$r < r_{0}$} \end{array} \end{equation} is the Laplace transform of the free-space GF, cf. \cite[Ch.~14.8, Eq.~(2)]{carslaw1986conduction}. $I_{0}(x), K_{0}(x)$ denote the modified Bessel functions of first and second kind, respectively, and zero order \cite[Sect.~9.6]{abramowitz1964handbook}. The variable $v$ is defined by \begin{equation} v:=\sqrt{s/D}. \end{equation} Note that the free GF takes into account the $\delta$ function term in Eq.~(\ref{eqMotionLaplace}) and therefore, the function $\tilde{f}(r, s \vert r_{0})$ in Eq.~(\ref{laplaceAnsatz}) satisfies the Laplace transformed 2D diffusion equation \cite[Ch.~14.8, Eq.~(3)]{carslaw1986conduction} \begin{equation}\label{laplaceDE} \frac{d^{2}\tilde{f}}{dr^{2}} + \frac{1}{r}\frac{d\tilde{f}}{dr} - v^{2} \tilde{f} = 0. \end{equation} The general solution to Eq.~(\ref{laplaceDE}) is given by \begin{equation} \tilde{f}(r, v) = B(s, r_{0})I_{0}(vr) + C(s, r_{0})K_{0}(vr), \end{equation} where $B(s, r_{0}), C(s, r_{0})$ are "constants" that may depend on $s$ and $r_{0}$. Because we require the BC Eq.~(\ref{BC_Infinity}) and $\lim_{x\rightarrow\infty}I_{0}(x)\rightarrow \infty$, the coefficient $B(s, r_{0})$ has to vanish and the solution becomes, \begin{equation}\label{laplaceBR} \tilde{f}(r, v\vert r_{0}) = C(v, r_{0}) K_{0}(vr). \end{equation} Next, turning to the case $r<a$, the GF satisfies \begin{equation}\label{laplaceDE2} \frac{d^{2}\tilde{p}^{<}}{dr^{2}} + \frac{1}{r}\frac{d\tilde{p}^{<}}{dr} - w^{2} \tilde{p}^{<} = 0, \end{equation} where $w$ is defined by \begin{equation} w:=v\sqrt{\frac{s + \kappa_{r} + \kappa_{d}}{s+\kappa_{d}}}. \end{equation} Therefore, the general solution, which takes into account the BC Eq.~(\ref{BC_Zero}) is \begin{equation}\label{generalSolution<} p^{<}(r, w\vert r_{0}) = A(s, r_{0}) I_{0}(wr), \end{equation} because $\lim_{x\rightarrow 0}xK_{1}(x)\neq 0$. The two "constants" $A(s, r_{0})$ and $C(s, r_{0})$ can be determined by the requirement that the GF and its derivative have to be continuous at $r=a$ \begin{eqnarray} \tilde{p}^{<}(r=a, s\vert r_{0}) &=& \tilde{p}^{>}(r=a, s\vert r_{0})\\ \frac{\partial\tilde{p}^{<}(r, s\vert r_{0})}{\partial r}\bigg\vert_{r=a} &=& \frac{\partial\tilde{p}^{>}(r, s\vert r_{0})}{\partial r}\bigg\vert_{r=a} \end{eqnarray} Using Eqs.~(\ref{laplaceAnsatz}),~(\ref{laplaceFree}),~(\ref{laplaceBR}), ~(\ref{generalSolution<}) as well as \begin{eqnarray} I^{\prime}_{0}(x) &=& I_{1}(x), \\ K^{\prime}_{0}(x) &=& -K_{1}(x),\\ x^{-1}&=&I_{0}(x)K_{1}(x)+I_{1}(x)K_{0}(x), \end{eqnarray} \cite[Eqs.~(9.6.27),~(9.6.15)]{abramowitz1964handbook}, we obtain \begin{eqnarray} A(s,r_{0}) &=& \frac{K_{0}(vr_{0})}{2\pi a D \mathcal{N}}, \\ C(s,r_{0}) &=& \frac{K_{0}(vr_{0})}{2\pi a D K_{0}(va)}\bigg[ \frac{I_{0}(wa)}{\mathcal{N}} - aI_{0}(va)\bigg], \end{eqnarray} where we introduced \begin{eqnarray} \mathcal{N} = v I_{0}(wa)K_{1}(va) + w I_{1}(wa)K_{0}(va). \end{eqnarray} \section{Exact Green's function in the time domain} To find the corresponding expressions for $p^{<}(r,t\vert r_{0}), p^{>}(r,t\vert r_{0})$ in the time domain, we apply the inversion theorem for the Laplace transformation \begin{equation}\label{inversionFormula} p^{<}(r, t \vert r_{0}) = \frac{1}{2\pi i} \int^{c+i\infty}_{c-i\infty} e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds. \end{equation} We note that $\tilde{p}^{<}(r, s\vert r_{0} )$ has three branch points at $s=0, -\kappa_{d}$ and $s = -\kappa_{r}-\kappa_{d} \equiv -\varphi$. Therefore, to calculate the Bromwich integral, we use the contour of Fig.~\ref{fig:contour} with a branch cut along the negative real axis, cf. \cite[Ch. 12.3, FIG. 40]{carslaw1986conduction}. We arrive at \begin{eqnarray}\label{cauchy} \int^{c+i\infty}_{c-i\infty} e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds =&-&\int_{\mathcal{C}_{2}} e^{ps}\,\tilde{p}^{<}(r, s\vert r_{0} )ds \nonumber\\ &-& \int_{\mathcal{C}_{4}} e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds. \end{eqnarray} To calculate the integral $\int_{\mathcal{C}_{2}}$, we choose $ s = D x^{2} e^{i \pi }. $ Then, \begin{eqnarray} v &=& ix, \quad\text{for}\,\, s\in ]-\infty, 0[\\ w &=& ix\sqrt{\frac{ Dx^{2} - \varphi}{Dx^{2} - \kappa_{d} }}\equiv ix\xi_{1} \quad\text{for}\,\, s\in ]-\infty, -\varphi[,\\ w &=& x\sqrt{\frac{\varphi - Dx^{2}}{Dx^{2} - \kappa_{d}}} \equiv x\xi_{2} \quad\text{for}\,\, s\in]-\varphi, -\kappa_{d}[,\\ w &=& ix\sqrt{\frac{\varphi - Dx^{2}}{\kappa_{d} - Dx^{2}}}= ix\xi_{1} \quad\text{for}\,\, s\in ]-\kappa_{d}, 0[,\\ \end{eqnarray} We now make use of \cite[Append.~3, Eqs.~(25), (26))]{carslaw1986conduction} \begin{eqnarray} I_{n}(xe^{\pm \pi i/2}) &=& e^{\pm n\pi i/2} J_{n}(x), \\ K_{n}(xe^{\pm \pi i/2}) &=& \pm\frac{1}{2}\pi i e^{\mp n\pi i/2} [-J_{n}(x) \pm i Y_{n}(x)]. \end{eqnarray} $J_{n}(x), Y_{n}(x)$ refer to the Bessel functions of first and second kind, respectively \cite[Sect.~9.1]{abramowitz1964handbook}. It follows that \begin{eqnarray} &&\int_{\mathcal{C}_{2}}e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds = \frac{1}{\pi a}\bigg[\int^{\sqrt{\frac{\varphi}{D}}}_{\sqrt{\frac{\kappa_{d}}{D}}}e^{-Dx^{2}t}g^{(2)}(r,r_{0}, x) dx \nonumber\\ && - \int^{\sqrt{\frac{\kappa_{d}}{D}}}_{0}e^{-Dx^{2}t}g^{(1)}(r,r_{0}, x) dx - \int^{\infty}_{\sqrt{\frac{\varphi}{D}}}e^{-Dx^{2}t}g^{(1)}(r,r_{0}, x) dx\bigg],\quad\quad\quad \end{eqnarray} where we introduced \begin{eqnarray} g^{(2)}(r, r_{0}, x) &\equiv& g^{(2)}_{R}(r, r_{0}, x) + i g^{(2)}_{I}(r, r_{0}, x), \nonumber \\ & = & I_{0}(x\xi_{2}r)\frac{\eta(r_{0}) + i \lambda(r_{0})}{\alpha^{2} + \beta^{2}},\quad\quad \\ g^{(1)}(r, r_{0}, x) &\equiv& g^{(1)}_{R}(r, r_{0}, x) + i g^{(1)}_{I}(r, r_{0}, x), \nonumber \\ & = & J_{0}(x\xi_{1}r)\frac{\omega(r_{0}) + i \varkappa(r_{0})}{\gamma^{2} + \delta^{2}},\quad\quad \end{eqnarray} and \begin{eqnarray} \eta(r_{0}) &=& \alpha Y_{0}(xr_{0})+ \beta J_{0}(xr_{0}),\\ \lambda(r_{0}) &=& \alpha J_{0}(xr_{0}) - \beta Y_{0}(xr_{0}),\\ \alpha &=& \xi_{2} I_{1}(\xi_{2} xa)Y_{0}(xa) + I_{0}(\xi_{2} xa)Y_{1}(xa),\\ \beta &=& \xi_{2} I_{1}(\xi_{2} xa)J_{0}(xa) + I_{0}(\xi_{2} xa)J_{1}(xa),\\ \omega(r_{0}) &=& \gamma Y_{0}(xr_{0})+ \delta J_{0}(xr_{0}),\\ \varkappa(r_{0}) &=& \gamma J_{0}(xr_{0}) -\delta Y_{0}(xr_{0}),\\ \gamma &=& \xi_{1} J_{1}(\xi_{1} xa)Y_{0}(xa) - J_{0}(\xi_{1} xa)Y_{1}(xa), \\ \delta &=& \xi_{1} J_{1}(\xi_{1} xa)J_{0}(xa) - J_{0}(\xi_{1} xa)J_{1}(xa). \end{eqnarray} Now, to calculate the integral along the contour $\mathcal{C}_{4}$, we choose $s = Dx^{2}e^{-i\pi}$ and after an analogous calculation one finds that \begin{equation} \int_{\mathcal{C}_{2}}e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds=-\bigg(\int_{\mathcal{C}_{4}}e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds\bigg)^{\ast}, \end{equation} where $\ast$ denotes complex conjugation. Thus, one obtains for the GF $p^{<}(r, t \vert r_{0})$ on the domain $r<a$ \begin{eqnarray}\label{GF<} p^{<}(r, t \vert r_{0}) &=& -\frac{1}{\pi} \mathfrak{Im}\bigg(\int_{\mathcal{C}_{2}}e^{st}\,\tilde{p}^{<}(r, s\vert r_{0} )ds\bigg)\nonumber \\ &=&-\frac{1}{\pi^{2} a}\bigg[\int^{\sqrt{\frac{\varphi}{D}}}_{\sqrt{\frac{\kappa_{d}}{D}}}e^{-Dx^{2}t}g^{(2)}_{I}(r,r_{0}, x) dx \nonumber\\ &-& \int^{\sqrt{\frac{\kappa_{d}}{D}}}_{0}e^{-Dx^{2}t}g^{(1)}_{I}(r,r_{0}, x) dx - \int^{\infty}_{\sqrt{\frac{\varphi}{D}}}e^{-Dx^{2}t}g^{(1)}_{I}(r,r_{0}, x) dx\bigg],\quad\quad\quad \end{eqnarray} Analogously, we can proceed to compute the GF for the region $r>a$. Therefore, we only give the result \begin{eqnarray}\label{GF>} &&p^{>}(r, t \vert r_{0}) = \frac{1}{4\pi D t} e^{-(r^{2}+r^{2}_{0})/4Dt}I_{0}\bigg(\frac{rr_{0}}{2Dt}\bigg)\nonumber \\ && +\frac{1}{\pi^{2}a}\bigg[\int^{\sqrt{\frac{\kappa_{d}}{D}}}_{0} e^{-Dx^{2}t} h^{(1)}(r,r_{0}, x) dx + \int^{\infty}_{\sqrt{\frac{\varphi}{D}}} e^{-Dx^{2}t} h^{(1)}(r,r_{0}, x) dx\quad\quad\quad\nonumber\\ && -\int^{\sqrt{\frac{\varphi}{D}}}_{\sqrt{\frac{\kappa_{d}}{D}}}e^{-Dx^{2}t} h^{(2)}(r,r_{0}, x) dx\bigg] -\frac{1}{2\pi}\int^{\infty}_{0}e^{-Dx^{2}t} h^{(3)}(r,r_{0}, x)xdx, \end{eqnarray} where we defined \begin{eqnarray} h^{(1)}(r,r_{0}, x) &=& J_{0}(x\xi_{1} a)\frac{\rho(r) \omega(r_{0})+ \psi(r) \varkappa(r_{0})}{[\gamma^{2}+\delta^{2}] [J_{0}^{2}(xa)+Y_{0}^{2}(xa)]} ,\\ h^{(2)}(r,r_{0}, x) &=& I_{0}(x\xi_{2} a)\frac{\rho(r) \eta(r_{0})+ \psi(r) \lambda(r_{0})}{[\alpha^{2}+\beta^{2}] [J_{0}^{2}(xa)+Y_{0}^{2}(xa)]},\\ h^{(3)}(r,r_{0}, x) &=& J_{0}(xa)\frac{\Pi(r, r_{0}) Y_{0}(xa)+ \Omega(r, r_{0}) J_{0}(xa)}{J_{0}^{2}(xa)+Y_{0}^{2}(xa)}, \end{eqnarray} and \begin{eqnarray} \rho(r) &=& J_{0}(xr)Y_{0}(xa) - Y_{0}(xr)J_{0}(xa),\\ \psi(r) &=& J_{0}(xr)J_{0}(xa) + Y_{0}(xr)Y_{0}(xa), \\ \Omega(r, r_{0}) &=& J_{0}(xr)J_{0}(xr_{0}) - Y_{0}(xr)Y_{0}(xr_{0}), \\ \Pi(r, r_{0}) &=& Y_{0}(xr)J_{0}(xr_{0}) + J_{0}(xr)Y_{0}(xr_{0}). \end{eqnarray} Note that the first term appearing on the rhs of Eq.~(\ref{GF>}) is the inverse Laplace transform of Eq.~(\ref{laplaceFree}), cf.~\cite[Ch.~14.8, Eq.~(2)]{carslaw1986conduction}. Finally, we can compute an exact expression for $q(r,t\vert r_{0})$ by virtue of Eq.~(\ref{q_p_Laplace}) and the convolution theorem of the Laplace transform. We obtain for $r < a$ \begin{eqnarray} q(r, t \vert r_{0}) &=& -\frac{1}{\pi^{2} a}\bigg[\int^{\sqrt{\frac{\varphi}{D}}}_{\sqrt{\frac{\kappa_{d}}{D}}}\bigg(\frac{e^{-Dx^{2}t}-e^{-\kappa_{d}x^{2}t}}{\kappa_{d}-Dx^{2}}\bigg)g^{(2)}_{I}(r,r_{0}, x) dx \nonumber\\ &-& \int^{\sqrt{\frac{\kappa_{d}}{D}}}_{0}\bigg(\frac{e^{-Dx^{2}t}-e^{-\kappa_{d}t}}{\kappa_{d}-Dx^{2}}\bigg)g^{(1)}_{I}(r,r_{0}, x) dx \nonumber\\ &-& \int^{\infty}_{\sqrt{\frac{\varphi}{D}}}\bigg(\frac{e^{-Dx^{2}t}-e^{-\kappa_{d}t}}{\kappa_{d}-Dx^{2}}\bigg)g^{(1)}_{I}(r,r_{0}, x) dx\bigg]. \quad\quad\quad \end{eqnarray} Clearly, $q(r, t \vert r_{0})$ vanishes for $r > a$. The case of an initially unbound molecule with $r_{0} < a$ and the case of the initially bound molecule will be considered in a forthcoming manuscript. \\ \begin{figure} \includegraphics[scale=0.45]{Fig1} \caption{Integration contour used for calculating the GF in the time domain, Eq.~(\ref{cauchy}).}\label{fig:contour} \end{figure} \section*{Acknowledgments} This research was supported by the Intramural Research Program of the NIH, National Institute of Allergy and Infectious Diseases.
\section{Preliminaries}\label{sec:prelim} The set of natural numbers is $\Nat=\{0,1,2,\dots\}$, and $\Nat_+=\{1,2,\dots\}$. For $k\in\Nat$ we define $[k]=\{1,\dots,k\}$; in particular, $[0]=\emptyset$. The domain of a partial function $f$ is denoted $\dom(f)$. For a set $A$, we denote by $A^*$ the set of sequences, or strings, of elements of $A$. A string $(a_1,\dots,a_n)\in A^*$ will be denoted $a_1\cdots a_n$, unless there is a danger of confusion. The concatenation of two strings $u$ and $v$ is denoted $u\cdot v$ or just $uv$, and the empty string is denoted $\varepsilon$. A string $u$ is a prefix (postfix) of a string $v$ if there exists a string $w$ such that $v=uw$ ($v=wu$); it is a proper prefix (postfix) if $w\neq\varepsilon$. The length of a string $u$ is denoted $|u|$. The cardinality of a set $A$ is denoted $|A|$. \smallskip We assume the reader to be familiar with top-down tree transducers, which work on ranked trees. This means that the number of children of a node of a tree is determined by the symbol at that node. A ranked alphabet $\Sigma$ is a finite set of symbols such that each symbol $a\in\Sigma$ is implicitly equipped with a rank $\rk(a)\in\Nat$. For $k\in\Nat$ we define $\Sigma^{(k)}=\{a\in\Sigma\mid\rk(a)=k\}$. To avoid trivialities, we assume that $\Sigma^{(0)}\neq\emptyset$. To indicate that $\sigma\in\Sigma$ has rank $k$, we also write it as $\sigma^{(k)}$. The set $\T_{\Sigma}$ of (finite, ordered, ranked) trees over the ranked alphabet $\Sigma$ is the smallest set (of terms) such that $a(t_1,\dots,t_k)\in\T_\Sigma$ if $k\in\Nat$, $a\in\Sigma^{(k)}$ and $t_1,\dots,t_k\in\T_\Sigma$. If $a\in\Sigma^{(0)}$, then we also write $a$ for the tree $a()$; if $a\in\Sigma^{(1)}$ and $t\in\T_\Sigma$, then we also write $at$ for $a(t)$. We represent the nodes of a tree in Dewey notation, i.e., by strings of positive natural numbers. The empty string $\varepsilon$ represents the root node and, for $i\in\Nat_+$, $vi$ represents the $i$th child of the node $v$ (and $v$ is the parent of $vi$). Every node $v$ of a tree $t$ has a label in $\Sigma$, denoted $\lab(t,v)$. Formally, the set $V(t)\subseteq\Nat_+^*$ of nodes (together with their labels) of the tree $t$ is inductively defined as: $V(t)=\{\varepsilon\}\cup\{iv\mid i\in[k], v\in V(t_i)\}$ if $t=a(t_1,\dots,t_k)$, $a\in\Sigma^{(k)}$ and $t_1,\dots,t_k\in\T_\Sigma$; moreover, $\lab(t,\varepsilon)=a$ and $\lab(t,iv)=\lab(t_i,v)$. A node $u$ is an ancestor of node $v$ (and $v$ is a descendant of $u$) if $u$ is a prefix of $v$; it is a proper ancestor/descendant if it is a proper prefix. For $\Delta\subseteq\Sigma$, we define $V_\Delta(t)=\{v\in V(t)\mid \lab(t,v)\in\Delta\}$; for $a\in\Sigma$, we write $V_a(t)$ instead of $V_{\{a\}}(t)$. The subtree of $t$ rooted at $v\in V(t)$ is denoted by $t/v$; formally, $t/\varepsilon=t$ and if $t=a(t_1,\dots,t_k)$ then $t/iv=t_i/v$. The size of $t$, denoted $\size(t)$, is its number $|V(t)|$ of nodes. The height of $t$, denoted $\height(t)$, is the maximal length of its nodes, i.e., $\Max\{|v|\mid v\in V(t)\}$. As an example, if $t=\sigma(\sigma(a,b),\tau(b))$, then $V(t)=\{\varepsilon,1,(1,1),(1,2),2,(2,1)\}$, $\;\lab(t,(1,2))=b$, $\;V_b(t)=\{(1,2),(2,1)\}$, $\;t/1=\sigma(a,b)$, $\;\size(t)=6$ and $\height(t)=2$. For a set of trees $\T$, we define $\Sigma(\T)$ to be the set of trees $a(t_1,\dots,t_k)$ such that $k\in\Nat$, $a\in\Sigma^{(k)}$ and $t_1,\dots,t_k\in\T$, and we define $\T_\Sigma(\T)$ to be the smallest set of trees $\T'$ such that $\T\cup\Sigma(\T')\subseteq \T'$. Note that $\T_\Sigma(\emptyset)=\T_\Sigma$. A $\Sigma$-\emph{pattern} is an upper portion, or prefix, of a tree in $\T_\Sigma$. Formally the set $\mathcal{P}_\Sigma$ of $\Sigma$-patterns is defined to be the set of trees $\T_{\Sigma}(\{\bot\})$, where $\bot$ is a new symbol of rank zero that is not in $\Sigma$. If $t_0$ is a pattern containing exactly $k$ occurrences of $\bot$, and $t_1,\dots,t_k$ is a sequence of $k$ patterns, then the pattern $t=t_0[t_1,\dots,t_k]$ is obtained from $t_0$ by replacing the $i$th occurrence of $\bot$ (in left-to-right order) by $t_i$. A $\Sigma$-\emph{context} is a $\Sigma$-pattern that contains exactly one occurrence of $\bot$. The set of all $\Sigma$-contexts is denoted $\C_\Sigma$. Thus, for $C\in\C_\Sigma$ and $t\in\T_\Sigma$, the tree $C[t]\in\T_\Sigma$ is obtained from the context $C$ by replacing the unique occurrence of $\bot$ in $C$ by $t$. On the set $\mathcal{P}_\Sigma$ we define a partial order $\sqsubseteq$ as follows: for patterns $t$ and $t'$ in $\mathcal{P}_\Sigma$, $t'$ is a \emph{prefix} of $t$, denoted $t'\sqsubseteq t$, if $t = t'[t_1,\dots,t_k]$ for suitable patterns $t_1,\dots,t_k$; equivalently, $V_a(t')\subseteq V_a(t)$ for every $a\in\Sigma$.\footnote{ In~\cite{DBLP:journals/jcss/EngelfrietMS09} the inverse of the partial order $\sqsubseteq$ is used.} Note that $\bot\sqsubseteq t$ for every pattern $t$. Every nonempty set $\Pi$ of $\Sigma$-patterns has a greatest lower bound $\sqcap \Pi$ in $\mathcal{P}_\Sigma$, called the \emph{largest common prefix} of the patterns in $\Pi$; it is the unique pattern $t'$ such that for every $v\in\Nat_+^*$ and $a\in\Sigma$, $v\in V_a(t')$ if and only if (1) $v\in V_a(t)$ for every $t\in\Pi$ and (2)~every proper ancestor of $v$ is in $V(t')$. This implies the following easy lemma. \begin{lemma}\label{lm:pat} Let $\Pi$ be a nonempty subset of $\T_\Sigma$, and let $v\in\Nat_+^*$. \\ Then, $v\in V_\bot(\sqcap \Pi)$ if and only if \begin{enumerate} \item[(1)] $v\in V(t)$ for every $t\in\Pi$, \item[(2)] $\lab(t_1,\hat{v})=\lab(t_2,\hat{v})$ for every proper ancestor $\hat{v}$ of $v$ and all $t_1,t_2\in\Pi$, and \item[(3)] there exist $t_1,t_2\in \Pi$ such that $\lab(t_1,v)\neq\lab(t_2,v)$. \end{enumerate} \end{lemma} For instance, $\sqcap\{\sigma(\tau(a),b),\sigma(b,b)\}= \sigma(\tau(a),b)\sqcap \sigma(b,b)= \sigma(\bot,b)$. For $t,t'\in\T_\Sigma$ and $v\in V(t)$, we denote by $t[v\leftarrow t']$ the tree that is obtained from $t$ by replacing its subtree $t/v$ by $t'$. More precisely, if $C$ is the unique context in $\C_\Sigma$ such that $C\sqsubseteq t$ and $C/v=\bot$, then $t[v\leftarrow t']=C[t']$. Let $\cS$ be a subset of $\T_\Sigma$ such that no $s\in\cS$ is a subtree of $s'\in\cS$ with $s\neq s'$. For a tree $t\in\T_\Sigma$ and a partial function $\psi: \cS\to \T_\Sigma$, we define $t[s \leftarrow \psi(s)\mid s\in \cS]$ to be the result of replacing every subtree $s$ of $t$ by $\psi(s)$, for every $s\in\cS$. More precisely, $t[s \leftarrow \psi(s)\mid s\in \cS]= t[v_1\leftarrow \psi(t/v_1)]\cdots[v_k\leftarrow \psi(t/v_k)]$ where $\{v_1,\dots,v_k\}=\{v\in V(t)\mid t/v\in\cS\}$. Note that $v_i$ is not an ancestor of $v_j$, for $i\neq j$, and hence the order of the substitutions $[v_i\leftarrow \psi(t/v_i)]$ is irrelevant. Note also that $t[s \leftarrow \psi(s)\mid s\in \cS]$ is defined if and only if $\psi(t/v_i)$ is defined for every $i\in[k]$. To formulate the rules of top-down tree transducers, we use variables $x_i$, with $i\in\Nat$, which are assumed to have rank~0. The set $\{x_0,x_1,x_2,\dots\}$ of all such variables is denoted $X$. For $k\in\Nat$, we denote $\{x_1,\dots,x_k\}$ by $X_k$; note that $X_0=\emptyset$. \newpage \section{Deterministic Top-Down Tree Transducers} \label{sec:dtop} A \emph{deterministic top-down tree transducer with regular look-ahead} (\emph{dtla} for short) is a tuple $M=(Q,\Sigma,\Delta,R,A,P,\delta)$ where $Q$ is a finite set of states of rank $1$, $\Sigma$ and $\Delta$ are the ranked input and output alphabets, respectively, and $P$ is a finite nonempty set of look-ahead states. The function $A$ maps look-ahead states to trees in $\T_\Delta(Q(\{x_0\}))$; for $p\in P$, the tree $A(p)$ is called the $p$-axiom of $M$. The finite set $R$ provides at most one rule \[ q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta \] for every state $q$, every input symbol $a$ of rank $k\geq 0$ and every sequence $p_1,\dots,p_k$ of look-ahead states. The right-hand side $\zeta$ of the rule is a tree in $\T_\Delta(Q(X_k))$, i.e., $\zeta = t[q_1(x_{i_1}),\ldots, q_r(x_{i_r})]$ for some pattern $t\in\mathcal{P}_{\Delta}$, $r=|V_\bot(t)|$, $q_j\in Q$ and $x_{i_j}\in\{x_1,\ldots,x_k\}$ for $j\in[r]$; we will denote $\zeta$ also by $\rhs(q,a,p_1,\dots,p_k)$. Finally, $\delta$ is the transition function of the (total deterministic bottom-up) look-ahead automaton $(P,\delta)$. That means that $\delta(a,p_1,\dots,p_k)\in P$ for every $k\geq 0$, $a\in\Sigma^{(k)}$ and $p_1,\dots,p_k\in P$. Examples of dtlas are given in the next section. Whenever we consider a dtla with the name $M$, it will be understood that its components are named $(Q,\Sigma,\Delta,R,A,P,\delta)$. When necessary we provide the components of a dtla $M$ with the subscript $M$. Then we have $Q_M$, $\Sigma_M$, $\Delta_M$, $R_M$, $\rhs_M$, etc. We denote by $\maxrhs(M)$ the maximal height of the axioms and the right-hand sides of the rules of $M$. We now define the semantics of the dtla $M$, starting with the semantics of its look-ahead automaton $(P,\delta)$. The transition function $\delta$ gives rise to a function $\delta^*$ that maps $\T_\Sigma$ to $P$. It is defined by $\delta^*(a(s_1,\ldots,s_k)) = \delta(a,\delta^*(s_1),\ldots,\delta^*(s_k))$ for $a\in\Sigma^{(k)}$ and $s_1,\dots,s_k\in\T_\Sigma$. For convenience, we denote the function $\delta^*$ as well with $\delta$. For $p\in P$ we denote by $\sem{p}_M$ the set of trees $s\in \T_\Sigma$ that have look-ahead state $p$, i.e., $\delta(s)=p$; we drop the subscript $M$ from $\sem{p}_M$ whenever it is clear from the context. Note that $\{\sem{p}\mid p\in P\}$ is a partition of $\T_\Sigma$. For a node $u$ of an input tree $s\in \T_\Sigma$, we also say that $\delta(s/u)$ is the look-ahead state at $u$. For $q\in Q$, $s\in\T_\Sigma$ and $u\in V(s)$ we define $\rhs(q,s,u)=\rhs(q,a,p_1,\dots,p_k)$ where $\lab(s,u)=a\in\Sigma^{(k)}$ and $p_i=\delta(s/ui)$ for every $i\in[k]$. Intuitively, $\rhs(q,s,u)$ is the right-hand side of the rule that is applied when $M$ arrives at node $u$ in state $q$ (if that rule exists); it is uniquely determined by the label of $u$ and the look-ahead states at its children. A \emph{sentential form} of $M$ for $s\in\T_\Sigma$ is a tree in $\T_\Delta(Q(V(s)))$, where the nodes in $V(s)$ are viewed as symbols of rank~0. For sentential forms $\xi,\xi'$ we write $\xi \Rightarrow_s \xi'$ if there exist $v\in V(\xi)$, $q\in Q$ and $u\in V(s)$ such that $$\xi/v=q(u) \mbox{ and } \xi'=\xi[v\leftarrow \rhs(q,s,u)[x_i\leftarrow ui\mid i\in\Nat_+]].$$ This will be called a computation step of $M$ in state $q$ at nodes $u$ and $v$. It is easy to see that the rewriting in computations is confluent.\footnote{Confluence means that if $\xi\Rightarrow^*_s\xi_1$ and $\xi\Rightarrow^*_s\xi_2$, then there exists a sentential form $\bar{\xi}$ such that $\xi_1\Rightarrow^*_s\bar{\xi}$ and $\xi_2\Rightarrow^*_s\bar{\xi}$.} Hence, if $\xi\Rightarrow^*_s t\in\T_\Delta$ and $\xi\Rightarrow^*_s\xi'$, then $\xi'\Rightarrow^*_s t$; thus, computations that start with a given sentential form lead to a unique tree in $\T_\Delta$ (if it exists). The dtla $M$ \emph{realizes} a partial function $\sem{M}: \T_\Sigma\to \T_\Delta$, called its \emph{translation}. Let $s\in \T_\Sigma$ and $\delta(s)=p\in P$. The output tree $\sem{M}(s)$ of the transducer $M$ for the input tree $s$ is the unique tree $t\in\T_\Delta$ such that $A(p)[x_0\leftarrow\varepsilon] \Rightarrow_s^* t$ (if it exists). For readability, we will write $M(s)$ instead of $\sem{M}(s)$. Two dtlas $M_1$ and $M_2$ are \emph{equivalent} if they realize the same translation, i.e., if $\Sigma_{M_1}=\Sigma_{M_2}$, $\Delta_{M_1}=\Delta_{M_2}$ and $\sem{M_1}=\sem{M_2}$. Intuitively, a sentential form $\xi$ consists of output that has already been produced by $M$; moreover, $\xi/v=q(u)$ means that $M$ has arrived at node $u$ of the input tree $s$ in state $q$ and, starting in that state, will translate the input subtree $s/u$ into the output subtree $M(s)/v$. Note that several parallel copies of $M$ can arrive at $u$ for different nodes of $\xi$, i.e., there may exist nodes $v'\neq v$ such that $\xi/v'=q'(u)$, where $q'$ may also be equal to $q$. A sentential form $\xi$ for $s$ is \emph{reachable} if $A(p)[x_0\leftarrow\varepsilon] \Rightarrow_s^* \xi$ where $p=\delta(s)$. Thus, if $M(s)$ is defined and $\xi$ is a reachable sentential form for $s$, then $\xi\Rightarrow_s^* M(s)$. We also define the semantics of every state $q$ of $M$ as a partial function $\sem{q}_M: \T_\Sigma\to \T_\Delta$ as follows. For $s\in\T_\Sigma$, $\sem{q}_M(s)$ is the unique tree $t\in\T_\Delta$ such that $q(\varepsilon)\Rightarrow_s^* t$ (if it exists). For readability, we will write $q_M(s)$ instead of $\sem{q}_M(s)$. The following lemma is easy to prove. \begin{lemma}\label{lm:elem} Let $s\in\T_\Sigma$ and $t\in\T_\Delta$. \begin{enumerate} \item[(1)] For every $q\in Q$ and $u\in V(s)$, $$q(u)\Rightarrow_s^* t \;\mbox{ if and only if }\; q_M(s/u)=t.$$ \item[(2)] For every sentential form $\xi$ for $s$, $$\xi\Rightarrow_s^* t \;\mbox{ if and only if }\; t=\xi[q(u)\leftarrow q_M(s/u)\mid q\in Q, u\in V(s)].$$ \item[(3)] If $\delta(s)=p$, then $$M(s)=A(p)[q(x_0)\leftarrow q_M(s)\mid q\in Q].$$ \item[(4)] For every $\bar{q}\in Q$, if $s=a(s_1,\dots,s_k)$, then $$\bar{q}_M(s)=\rhs(\bar{q},a,\delta(s_1),\dots,\delta(s_k)) [q(x_i)\leftarrow q_M(s_i)\mid q\in Q,\,i\in[k]].$$ \end{enumerate} \end{lemma} \begin{proof} (1) follows from the obvious bijection between the nodes of $s/u$ and the nodes of $s$ with prefix $u$ (i.e., the descendants of $u$ in $s$); (2) is obvious from (1) and the fact that the computation steps $\xi \Rightarrow_s \xi'$ of $M$ are context-free\footnote{In fact, the computations of $M$ on $s$ can be viewed as derivations of a context-free grammar with set of nonterminals $Q(V(s))$ and rules $q(u)\to \rhs(q,s,u)[x_i\leftarrow ui\mid i\in\Nat_+]$.}; (3) and (4) follow from (2), taking $\xi=A(p)[x_0\leftarrow\varepsilon]$ and $\xi=\rhs(\bar{q},s,\varepsilon)[x_i\leftarrow i\mid i\in[k]]$, respectively. \qed \end{proof} Note that (3) and (4) of Lemma~\ref{lm:elem} form an alternative way of defining the semantics of $M$ (recursively). \medskip {\bf Convention.} For a given dtla $M$ it can (and will, from now on) be assumed that all its states and look-ahead states are \emph{reachable} in the following sense. A look-ahead state $p$ is reachable if $\sem{p}_M\neq\emptyset$. A state $q$ is reachable if $q$ occurs in an axiom, or if $q$ occurs in the right-hand side of a rule of which the left-hand side starts with a reachable state. \medskip A \emph{deterministic top-down tree transducer} (\emph{dtop} for short) is a dtla $M$ with trivial look-ahead automaton $(P,\delta)$, i.e., $P$ is a singleton. Whenever convenient, we drop $(P,\delta)$ from the tuple defining $M$, we identify $A$ with the unique axiom $A(p)$, we write a rule as $q(a(x_1,\dots,x_k))\to \zeta$ rather than $q(a(x_1\ot p,\dots,x_k\ot p))\to \zeta$ (where $p$ is the unique look-ahead state of $M$) and we denote $\zeta$ by $\rhs(q,a)$. A dtla $M$ is \emph{proper} (a \emph{dtpla} for short) if it is not a dtop, i.e., if $|P|\geq 2$. Obviously, to decide whether $M$ is equivalent to a dtop, we may assume that $M$ is proper. A dtla $M$ is \emph{total} if $\dom(\sem{M})=\T_\Sigma$, i.e., if its translation $\sem{M}$ is a total function. Note that it is decidable whether $M$ is total, because $\dom(\sem{M})$ is effectively a regular tree language (cf.~\cite[Corollary~2.7]{DBLP:journals/mst/Engelfriet77}). From now on we only consider total dtlas. A dtla $M$ is \emph{complete} if $\rhs(q,a,p_1,\dots,p_k)$ is defined for every $q\in Q$, $k\in\Nat$, $a\in\Sigma^{(k)}$ and $p_1,\dots,p_k \in P$. By Lemma~\ref{lm:elem}(4), this means that $q_M(s)$ is defined for every $q\in Q$ and $s\in\T_\Sigma$. Thus, by Lemma~\ref{lm:elem}(3), if $M$ is complete, then $M$ is total. A dtla $M$ is \emph{linear} if for every rule $q(a(x_1\ot p_1,\ldots,x_k\ot p_k)) \to \zeta$, each variable $x_i$ occurs at most once in $\zeta$. A dtla $M$ is \emph{ultralinear} if there is a mapping $\mu: Q\to\Nat$ such that for every rule $q(a(x_1\ot p_1,\ldots,x_k\ot p_k)) \to \zeta$ the following two properties hold for every $\bar{q}(x_i)$ that occurs in $\zeta$: (1) $\mu(\bar{q})\geq \mu(q)$, and (2) if $\mu(\bar{q})= \mu(q)$, then $x_i$ occurs exactly once in $\zeta$. Obviously, every linear dtla is ultralinear. A dtla $M$ is \emph{nonerasing} if it does not have erasing rules. A rule of $M$ is an \emph{erasing rule} if its right-hand side is in $Q(X)$, i.e., contains no symbols from $\Delta$. A dtla $M$ is \emph{bounded erasing} (for short, \emph{b-erasing}) if there is no cycle in the directed graph $E_M$ with the set of nodes $Q$ and an edge from $q$ to $q'$ if there is an erasing rule of the form $q(a(x_1\ot p_1,\ldots,x_k\ot p_k)) \to q'(x_j)$. Obviously, every nonerasing dtla is b-erasing. \subsubsection{Look-ahead states in input trees.} Let $M=(Q,\Sigma,\Delta,R,A,P,\delta)$ be a total dtla. To analyze the behaviour of $M$ for different look-ahead states, we will consider input trees $\bar{s}$ with occurrences of $p\in P$, viewed as input symbol of rank zero, representing an absent subtree $s$ with $\delta(s)=p$. If $M$ arrives in state $q$ at a $p$-labeled leaf of $\bar{s}$, then $M$ will output the new symbol $\angl{q,p}$ of rank zero, representing the absent output tree $q_M(s)$. In this way, $M$ translates input trees $\bar{s}\in\T_\Sigma(P)$ to output trees in $\T_\Delta(Q\times P)$.\footnote{Without loss of generality we assume that $P$ and $\Sigma$ are disjoint, and so are $Q\times P$ and $\Delta$.} Formally, we extend $M$ to a dtla $M^\circ=(Q,\Sigma^\circ,\Delta^\circ,R^\circ,A,P,\delta^\circ)$ where $\Sigma^\circ=\Sigma\cup P$ such that every element of $P$ has rank zero, $\Delta^\circ=\Delta\cup (Q\times P)$ such that every element of $Q\times P$ has rank zero, $R^\circ$ is obtained from $R$ by adding the rules $q(p)\to \angl{q,p}$ for all $q\in Q$ and $p\in P$ such that $q_M(s)$ is defined for some $s\in\sem{p}_M$, and $\delta^\circ$ is the extension of $\delta$ such that $\delta^\circ(p)=p$ for every $p\in P$. For notational simplicity, we will denote $\delta^\circ(\bar{s})$, $M^\circ(\bar{s})$ and $q_{M^\circ}(\bar{s})$ by $\delta(\bar{s})$, $M(\bar{s})$ and $q_M(\bar{s})$, respectively, for every input tree $\bar{s}\in\T_\Sigma(P)$. But note that we do \emph{not} drop $^\circ$ from $\sem{p}_{M^\circ}$, $\sem{M^\circ}$ and $\sem{q}_{M^\circ}$, i.e., $\sem{p}_M$, $\sem{M}$ and $\sem{q}_M$ keep their meaning. We will use the following elementary lemma, which expresses the above intuition. \begin{lemma}\label{lm:MM} Let $M$ be a total dtla. Let $\bar{s}$ be a tree in $\T_\Sigma(P)$, and for every $p\in P$, let $s_p$ be a tree in $\T_\Sigma(P)$ such that $\delta(s_p)=p$. Then $\delta(\bar{s}[p\leftarrow s_p\mid p\in P])=\delta(\bar{s})$ and \begin{equation}\label{eq:MM} M(\bar{s}[p\leftarrow s_p\mid p\in P])= M(\bar{s})\big[\angl{q,p}\leftarrow q_M(s_p)\mid q\in Q, p\in P \big]. \end{equation} \end{lemma} \begin{proof} Let $s'=\bar{s}[p\leftarrow s_p\mid p\in P]$. It should be clear that $\delta(s'/v)=\delta(\bar{s}/v)$ for every node $v$ of $\bar{s}$. Let $p'=\delta(s')=\delta(\bar{s})$. We first assume that $s_p\in\T_\Sigma$ for every $p\in P$. Thus $s'\in\T_\Sigma$, and so $M(s')$ is defined. Let $U$ be the set of nodes $u$ of $\bar{s}$ such that $\lab(\bar{s},u)\in P$, and for $u\in U$, let $p_u=\lab(\bar{s},u)$. Now consider a computation $A(p')[x_0\leftarrow\varepsilon]\Rightarrow_{s'}^*\xi$ of $M$ such that $\xi\in\T_\Delta(Q(U))$ and none of the computation steps is at a node $u\in U$ (and hence not at a descendant of $u$). Then, by the observation above, also $A(p')[x_0\leftarrow\varepsilon]\Rightarrow_{\bar{s}}^*\xi$. By Lemma~\ref{lm:elem}(2), $M(s')=\xi[q(u)\leftarrow q_M(s_{p_u})\mid q\in Q, u\in U]$. Thus, $q_M(s_{p_u})$ is defined for every $q(u)$ that occurs in $\xi$, and so $q(p_u)\to \angl{q,p_u}$ is a rule of $M^\circ$. Hence $\xi\Rightarrow_{\bar{s}}^*\xi[q(u)\leftarrow \angl{q,p_u}\mid q\in Q, u\in U]$ and so $M(\bar{s})=\xi[q(u)\leftarrow \angl{q,p_u}\mid q\in Q, u\in U]$. This proves Equation~(\ref{eq:MM}) for the case where $s_p\in\T_\Sigma$. It also implies that $M^\circ$ is total, because for every $\bar{s}\in\T_\Sigma(P)$ one can choose some $s_p\in\sem{p}_M$ for each $p\in P$, and then $M(\bar{s}[p\leftarrow s_p\mid p\in P])$ is defined and hence so is $M(\bar{s})$. Since $M^\circ$ is total, the previous argument also proves Equation~(\ref{eq:MM}) for the general case where $s_p\in\T_\Sigma(P)$. \qed \end{proof} Note that the proof of Lemma~\ref{lm:MM} shows that if $M$ is total, then so is $M^\circ$. From Lemma~\ref{lm:MM} we immediately obtain the next lemma for $\Sigma$-contexts. Note that for every $C\in\C_\Sigma$ and $p\in P$, the tree $M(C[p])$ is in $\T_\Delta(Q\times\{p\})$. \begin{lemma}\label{lm:semantics} Let $M$ be a total dtla. Let $C\in\C_\Sigma$, $s\in\T_\Sigma(P)$ and $p\in P$ such that $\delta(s)=p$. Then $\delta(C[s])=\delta(C[p])$ and $M(C[s])=M(C[p])\big[\angl{q,p}\leftarrow q_M(s)\mid q\in Q \big]$. \end{lemma} \begin{proof} Apply Lemma~\ref{lm:MM} with $\bar{s}=C[p]$ and $s_p=s$. \qed \end{proof} In Section~\ref{sec:diff} the next lemma will be needed. \begin{lemma}\label{lm:reach} Let $M$ be a total dtop. Then $M$ is complete, and for every $q\in Q$ there exists $C\in\C_\Sigma$ such that $\angl{q,p}$ occurs in $M(C[p])$, where $P=\{p\}$. \end{lemma} \begin{proof} It is convenient to assume that $p=\bot$ (and so $C[p]=C$). We first prove the second statement. Since we assume, by convention, that every state $q$ of $M$ is reachable, we proceed by induction on the definition of reachability. If $q(x_0)$ occurs in the axiom $A$ of $M$, then $C=\bot$ satisfies the requirement because $M(\bot)=A[\bar{q}(x_0)\leftarrow\angl{\bar{q},\bot}\mid \bar{q}\in Q]$. Now let $q(a(x_1,\dots,x_k))\to \zeta$ be a rule of $M$ such that $q$ is reachable, and let $\zeta/z=q'(x_j)$. By induction there exist $C\in\C_\Sigma$ and $v\in V(M(C))$ such that $M(C)/v=\angl{q,\bot}$. Let $s_j=\bot$ and choose $s_m\in\T_\Sigma$ for $m\in[k]-\{j\}$. Let $C'$ be the $\Sigma$-context $C[a(s_1,\dots,s_k)]$. By Lemma~\ref{lm:semantics}, $M(C')=M(C)[\angl{\bar{q},\bot}\leftarrow \bar{q}_M(a(s_1,\dots,s_k))\mid \bar{q}\in Q]$ and by Lemma~\ref{lm:elem}(4), $q_M(a(s_1,\dots,s_k))=\zeta[\bar{q}(x_i)\leftarrow \bar{q}_M(s_i)\mid \bar{q}\in Q, i\in[k]]$. Then $M(C')/vz=q_M(a(s_1,\dots,s_k))/z=q'_M(s_j)=\angl{q',\bot}$. So, $\angl{q',\bot}$ occurs in $M(C')$. To show that $M$ is complete, let $q\in Q$ and $a\in\Sigma^{(k)}$. We have shown that there exist $C\in\C_\Sigma$ and $v\in V(M(C))$ such that $M(C)/v=\angl{q,\bot}$. Let $C'$ be as above. Then $M(C')/v=q_M(a(s_1,\dots,s_k))$. Hence $\rhs(q,a)$ is defined by Lemma~\ref{lm:elem}(4). \qed \end{proof} \newpage \section{Difference Trees} \label{sec:difftree} Let $M$ be a total dtla. We wish to decide whether $M$ is equivalent to a dtop. Let $C$ be a $\Sigma$-context and let $p,p'\in P$. As explained in the Introduction, we are interested in the difference between the output of $M$ on input $C[p]$ and its output on input $C[p']$, see also Lemma~\ref{lm:semantics}. Intuitively, a dtop $N$ that is equivalent to $M$ does not know whether the subtree $s$ of an input tree $C[s]$ has look-ahead state $p$ or $p'$, and hence, when reading the context $C$, it can output at most the largest common prefix $M(C[p])\sqcap M(C[p'])$ of the output trees $M(C[p])$ and $M(C[p'])$.\footnote{Recall that $M(C[p])$ denotes $M^\circ(C[p])$, which is defined because $M^\circ$ is total (as shown in the proof of Lemma~\ref{lm:MM}).} Let $v$ be a node of $M(C[p])\sqcap M(C[p'])$ with label $\bot$. Then we say that $M(C[p])/v$ is a \emph{difference tree} of $M$ (and hence, by symmetry, so is $M(C[p'])/v$). Thus, a difference tree is a part of the output that can be produced by $M$ because it knows that $s$ has look-ahead state $p$ (or $p'$). Intuitively, to simulate $M$, the dtop $N$ must store the difference trees in its state. Hence, for $N$ to exist, there should be finitely many difference trees (as will be proved in Corollary~\ref{co:difffin}). We denote the set of all difference trees of $M$ by $\diff(M)$, for varying $C$, $p$, $p'$ and $v$. Thus we define \begin{eqnarray*} \lefteqn{\diff(M)=} \\ & & \hspace{0.4cm} \{M(C[p])/v \mid C\in\C_\Sigma,\,p\in P,\,\exists p'\in P: v\in V_\bot(M(C[p]) \sqcap M(C[p']))\}, \end{eqnarray*} which is a subset of $\T_\Delta(Q\times P)$. We define the number $\maxdiff(M)\in\Nat\cup\{\infty\}$ to be the maximal height of all difference trees of $M$, i.e.,\footnote{For a nonempty set $S\subseteq \Nat$, $\max S$ is the maximal number in $S$ if $S$ is finite, and $\max S = \infty$ otherwise. } $$\maxdiff(M)=\max\{\height(t)\mid t\in\diff(M)\}.$$ Intuitively, $\maxdiff(M)$ gives a measure of how much the transducer $M$ makes use of its look-ahead information. Obviously, $\maxdiff(M)$ is finite (i.e., in $\Nat$) if and only if $\diff(M)$ is finite. We will say that a number $h(M)\in\Nat$ is a \emph{difference bound} for $M$ if the following holds: if $\diff(M)$ is finite, then $\maxdiff(M)\leq h(M)$. Our first main result (Theorem~\ref{th:alg}) is that if a difference bound for $M$ is known, then we can decide whether $M$ is equivalent to a dtop, and if so, construct such a dtop from $M$. Our second main result (Theorem~\ref{th:ultra}) is that a difference bound can be computed for every total dtla $M$ that is ultralinear and b-erasing. A node $v\in V_\bot(M(C[p]) \sqcap M(C[p']))$ will be called a \emph{difference node} of $M(C[p])$ and $M(C[p'])$. It is characterized in the next lemma, which follows immediately from Lemma~\ref{lm:pat}. \begin{lemma}\label{lm:diff_char} Let $C\in\C_\Sigma$, $p,p'\in P$ and $v\in\Nat_+^*$. Then, \\ $v$ is a difference node of $M(C[p])$ and $M(C[p'])$ if and only if \begin{enumerate} \item[(1)] $v\in V(M(C[p]))\cap V(M(C[p']))$, \item[(2)] $\lab(M(C[p]),\hat{v})=\lab(M(C[p']),\hat{v})$ for every proper ancestor $\hat{v}$ of $v$, and \item[(3)] $\lab(M(C[p]),v)\neq\lab(M(C[p']),v)$. \end{enumerate} \end{lemma} Note that if $v$ is a difference node of $M(C[p])$ and $M(C[p'])$, then $p\neq p'$. Thus, in the definition of $\diff(M)$ we can assume that $p\neq p'$. Hence, if $M$ is a dtop then $\diff(M)=\emptyset$ and so $\maxdiff(M)=0$. Note also that, to compute $\maxdiff(M)$ for a dtla $M$, it suffices to consider difference trees of non-zero height, i.e., difference nodes $v$ that are not leaves of $M(C[p])$. We now give some examples of dtlas with their sets of difference trees. \begin{example}\label{ex:Mex} Let $\Sigma=\Delta=\{\sigma^{(1)},a^{(0)},b^{(0)}\}$, which means that $\Sigma$ and $\Delta$ are the ranked alphabet $\{\sigma,a,b\}$ with $\rk(\sigma)=1$ and $\rk(a)=\rk(b)=0$. We consider the following total dtla $M=(Q,\Sigma,\Delta,R,A,P,\delta)$ with $M(\sigma^na)=a$ and $M(\sigma^nb)=\sigma^nb$ for every $n\in\Nat$. It is, in fact, the dtla $M_{\rm ex}$ of the Introduction, for this particular input alphabet. Its set of look-ahead states is $P=\{p_a,p_b\}$ with transition function $\delta$ defined by $\delta(a)=p_a$, $\delta(b)=p_b$, $\delta(\sigma,p_a)=p_a$ and $\delta(\sigma,p_b)=p_b$. Its set of states is $Q=\{q\}$, its two axioms are $A(p_a)=a$ and $A(p_b)=q(x_0)$, and its set $R$ of rules contains the two rules $q(\sigma(x_1\ot p_b))\to \sigma(q(x_1))$ and $q(b)\to b$. Clearly, $\C_\Sigma=\{\sigma^n\bot\mid n\in\Nat\}$ and for $C=\sigma^n\bot$ we have $M(C[p_a])=a$ and $M(C[p_b])=\sigma^n\angl{q,p_b}$. Since $M(C[p_a])\sqcap M(C[p_b])=\bot$, the only difference node of $M(C[p_a])$ and $M(C[p_b])$ is $\varepsilon$, and we obtain the difference trees $M(C[p_a])$ and $M(C[p_b])$. Hence, $\diff(M)=\{a\}\cup\{\sigma^n\angl{q,p_b}\mid n\in\Nat\}$ and $\maxdiff(M)=\infty$. Since $\diff(M)$ is infinite, $M$ is not equivalent to a dtop, as will be shown in Corollary~\ref{co:difffin}. \qed \end{example} \begin{example}\label{ex:Mthree} Let $\Sigma=\Delta=\{\sigma^{(1)},\tau^{(1)},a^{(0)},b^{(0)}\}$ and consider the following total dtla $M$. For an input tree $s$ with leaf $a$, $M$ outputs the top-most 3 unary symbols of $s$ and the leaf $a$ if $\size(s)\geq 4$, and it outputs $s$ if $\size(s)\leq 3$. For an input tree with leaf $b$, $M$ outputs $b$. The look-ahead automaton of $M$ is similar to the one of the previous example: $P=\{p_a,p_b\}$ with $\delta(a)=p_a$, $\delta(b)=p_b$ and $\delta(\gamma,p)=p$ for $\gamma\in\{\sigma,\tau\}$ and $p\in P$. The set of states is $Q=\{q_0,q_1,q_2\}$. The axioms are $A(p_a)=q_0(x_0)$ and $A(p_b)=b$. For $q\in Q$, $M$ has the rules $q(a)\to a$, and for $\gamma\in\{\sigma,\tau\}$, the rules $q_0(\gamma(x_1\ot p_a))\to \gamma(q_1(x_1))$, $\;q_1(\gamma(x_1\ot p_a))\to \gamma(q_2(x_1))$, and $q_2(\gamma(x_1\ot p_a))\to \gamma(a)$. As in the previous example, $M(C[p_a])\sqcap M(C[p_b])=\bot$ for every $\Sigma$-context $C$, and so $\diff(M)$ consists of all trees $M(C[p_a])$ and $M(C[p_b])$, i.e., $\diff(M)$ is the finite set containing the trees $b$ and $\angl{q_0,p_a}$, and all trees $\gamma_1(\angl{q_1,p_a})$, $\;\gamma_1(\gamma_2(\angl{q_1,p_a}))$ and $\gamma_1(\gamma_2(\gamma_3(a)))$ for $\gamma_1,\gamma_2,\gamma_3\in\{\sigma,\tau\}$. Hence, $\maxdiff(M)=3$. It is not difficult to see that there exists a dtop $N$ equivalent to $M$. It stores in its state the top-most $\leq 3$ unary symbols of the input tree $s$, and depending on the leaf label of $s$, it outputs these symbols and $a$ or it outputs $b$. \qed \end{example} \begin{example}\label{ex:Mleaves} Let $\Sigma=\{\sigma^{(2)},aa^{(0)},ab^{(0)},ba^{(0)},bb^{(0)}\}$ where we view $aa$, $ab$, $ba$ and $bb$ as symbols, and let $\Delta=\{\sigma^{(3)},\#^{(2)},a^{(0)},b^{(0)}\}\cup\Sigma^{(0)}$. We consider the following total dtla $M$ such that $M(aa)=aa$, $M(ab)=ab$, $M(ba)=ba$, $M(bb)=bb$ and for every $s_1,s_2\in\T_\Sigma$, $M(\sigma(s_1,s_2))=\sigma(M(s_1),M(s_2),\#(y,z))$ where $y\in\{a,b\}$ is the first letter of the label of the left-most leaf of $\sigma(s_1,s_2)$ and $z\in\{a,b\}$ is the second letter of the label of its right-most leaf. Its look-ahead automaton has four states $p_{yz}$ with $y,z\in\{a,b\}$, such that $\delta(yz)=p_{yz}$ and $\delta(\sigma,p_{wx},p_{yz})=p_{wz}$ for all $w,x,y,z\in\{a,b\}$. It has one state $q$, its axioms are $A(p_{yz})=q(x_0)$, and its rules are $q(yz)\to yz$ and $$q(\sigma(x_1\ot p_{wx},x_2\ot p_{yz}))\to \sigma(q(x_1),q(x_2),\#(w,z))$$ for all $w,x,y,z\in\{a,b\}$. Consider a $\Sigma$-context $C$ and the trees $M(C[p_{aa}])$ and $M(C[p_{ba}])$. Let $u$ be the node of $C$ with $C/u=\bot$. It is easy to see that the difference nodes of $M(C[p_{aa}])$ and $M(C[p_{ba}])$ are the node $u$ and all nodes $v\cdot(3,1)$ such that $v\neq u$ is a node of $C$ and $u$ is the left-most leaf of $C/v$. That gives the difference trees $M(C[p_{aa}])/u=\angl{q,p_{aa}}$, $M(C[p_{ba}])/u=\angl{q,p_{ba}}$, $M(C[p_{aa}])/v\cdot(3,1)=a$ and $M(C[p_{ba}])/v\cdot(3,1)=b$. In this way we obtain that $\diff(M)=\{a,b\}\cup\{\angl{q,p_{yz}}\mid y,z\in\{a,b\}\}$. Thus, $\maxdiff(M)=0$. Clearly, there is a dtop $N$ equivalent to $M$. It has three states $q,q_1,q_2$, axiom $q(x_0)$, and rules $q(yz)\to yz$, $$q(\sigma(x_1,x_2))\to \sigma(q(x_1),q(x_2),\#(q_1(x_1),q_2(x_2))),$$ $q_i(\sigma(x_1,x_2))\to q_i(x_i)$ for $i=1,2$, $\;q_1(yz)\to y$ and $q_2(yz)\to z$ for $y,z\in\{a,b\}$. \qed \end{example} \begin{example}\label{ex:Mcounter} Let $\Sigma=\{\sigma^{(2)},a^{(0)}\}$ and $\Delta=\{e^{(0)},o^{(0)}\}$, and consider the following total dtla $M$ that translates every tree $s\in\T_\Sigma$ into $e$ if $\size(s)$ is even and into $o$ if it is odd. Its look-ahead automaton has two states $p_e$ and $p_o$ with $\delta(a)=p_o$, $\delta(\sigma,p_e,p_e)=\delta(\sigma,p_o,p_o)=p_e$ and $\delta(\sigma,p_e,p_o)=\delta(\sigma,p_o,p_e)=p_o$. It set of states is empty, and its axioms are $A(p_e)=e$ and $A(p_o)=o$. For every $\Sigma$-context $C$, $\{M(C[p_e]),M(C[p_o])\}=\{e,o\}$. Hence $\diff(M)=\{e,o\}$ and $\maxdiff(M)=0$. Although $\diff(M)$ is finite, there is obviously no dtop equivalent to $M$. \qed \end{example} \section{Conclusion} In Theorem~\ref{th:alg} we have given a general algorithm to decide for a given total dtla $M$ with a given difference bound, whether $M$ is equivalent to a dtop, and if so, to construct such a dtop. In Sections~\ref{sec:upper} and~\ref{sec:graphs} we have shown that a difference bound can be computed for total dtlas that are ultralinear and bounded erasing, or output-monadic, or initialized and depth-uniform. We would like to extend our results to the non-total case where a dtla realizes a partial function, to the case where the dtla and the dtop are restricted to a given regular tree language, and to more general dtlas (preferably to all dtlas, of course). Even more generally, we would like to have an algorithm that for a given dtla constructs an equivalent dtla with a minimal number of look-ahead states. As observed at the end of the Introduction, regular look-ahead can always be removed from a macro tree transducer \cite{DBLP:journals/jcss/EngelfrietV85}. Hence another more general question is: Is it decidable for a given macro tree transducer whether it is equivalent to a top-down tree transducer? \section{Difference Tuples} \label{sec:diff} Let $M$ be a total dtpla and let $P=\{\hat{p}_1,\dots,\hat{p}_n\}$, where the order of the look-ahead states is fixed as indicated. Recall that a dtpla is a proper dtla, i.e., a dtla that is not a dtop, hence $n\geq 2$. For a given context $C$ consider the trees $M(C[\hat{p}_1]),\dots,M(C[\hat{p}_n])$. Intuitively, the largest common prefix of all these trees does \emph{not} depend on the look-ahead. In contrast, the subtrees of the above trees that are not part of the largest common prefix, \emph{do} depend on the look-ahead information. For trees $t_1,\dots,t_n\in\T_\Delta(Q\times P)$ we define \[ \diftup(t_1,\dots,t_n):=\{(t_1/v,\dots,t_n/v)\mid v\in V_\bot(\sqcap\{t_1,\dots,t_n\})\}, \] which is a set of $n$-tuples in $\T_\Delta(Q\times P)^n$. We define the \emph{set of difference tuples} of $M$ as \[ \diftup(M):=\bigcup_{C\in \C_\Sigma}\diftup(M(C[\hat{p}_1]),\dots,M(C[\hat{p}_n])). \] For a $\Sigma$-context $C$, we define $\pref(M,C)\in \mathcal{P}_\Delta$ as \[ \pref(M,C):=\sqcap\{M(C[p])\mid p\in P\}=\sqcap\{M(C[\hat{p}_1]),\dots,M(C[\hat{p}_n])\}. \] Note that $\pref(M,C)$ is a $\Delta$-pattern because a node with label $\angl{q,\hat{p}_i}$ in $M(C[\hat{p}_i])$ cannot have the same label in $M(C[\hat{p}_j])$ for $i\neq j$. Note also that $$\diftup(M)=\{(M(C[\hat{p}_1])/v,\dots,M(C[\hat{p}_n])/v)\mid C\in \C_\Sigma,\,v\in V_\bot(\pref(M,C))\}.$$ Difference tuples are introduced for the following reason, cf. Lemma~\ref{lm:semantics}. We wish to decide whether $M$ is equivalent to a dtop. If there exists a dtop $N$ that is equivalent to $M$, then we expect intuitively for any $s\in\T_\Sigma$, that $N(C[s])=t[q_{1N}(s),\dots,q_{rN}(s)]$ where $t=\pref(M,C)=\sqcap\{M(C[\hat{p}_1]),\dots,M(C[\hat{p}_n])\}$ and $r=|V_\bot(t)|$; in other words, since $N$ does not know the look-ahead state $\delta_M(s)$ of $s$, it translates $C$ into the largest common prefix of the output trees $M(C[\hat{p}_1]),\dots,M(C[\hat{p}_n])$. Moreover, if the $i$th occurrence of $\bot$ is at node $v_i$ of $t$ for $i\in[r]$, then we expect the difference tuple $(M(C[\hat{p}_1])/v_i,\dots,M(C[\hat{p}_n])/v_i)$ to be stored in the state $q_i$ of $N$; in this way $N$ is prepared to continue its simulation of $M$ on the subtree $s$. This will be proved in Lemma~\ref{lm:dtladtopprop2}, under the condition that $M$ is canonical and $N$ is earliest. If $N$ is also canonical, then its states are in one-to-one correspondence with the difference tuples of $M$, as will be proved in Lemma~\ref{lm:Qdiff}. Before giving some examples, we show that the maximal height of the components of the difference tuples of $M$ is $\maxdiff(M)$, defined in Section~\ref{sec:difftree}. This implies that $\diftup(M)$ is finite if and only if $\diff(M)$ is finite. \begin{lemma}\label{lm:twodiff} Let $M$ be a total dtpla. Then $$\maxdiff(M)= \max\{\height(M(C[p])/v)\mid C\in\C_\Sigma,p\in P, v\in V_\bot(\pref(M,C))\}.$$ \end{lemma} \begin{proof} ($\leq$) We show that every difference tree is a subtree of a component of a difference tuple. Consider a difference tree $M(C[p])/v$ with $C\in\C_\Sigma$, $p\in P$ and $v$ a difference node of $M(C[p])$ and $M(C[p'])$ where $p'\in P$, i.e., $v\in V_\bot(M(C[p]) \sqcap M(C[p']))$. Since $\pref(M,C) \sqsubseteq M(C[p]) \sqcap M(C[p'])$, there is an ancestor $\hat{v}$ of $v$ such that $\hat{v}\in V_\bot(\pref(M,C))$. Thus, $M(C[p])/\hat{v}$ is a component of a difference tuple, and $M(C[p])/v$ is one of its subtrees. ($\geq$) We show that every component of a difference tuple is a difference tree. Consider $M(C[p])/v$ with $C\in\C_\Sigma$, $p\in P$ and $v\in V_\bot(\pref(M,C))$. By Lemma~\ref{lm:pat}, each proper ancestor of $v$ has the same label in all $M(C[\bar{p}])$, $\bar{p}\in P$, but $v$ does not have the same label in all $M(C[\bar{p}])$. Thus, there exists $p'\in P$ such that $v$ has different labels in $M(C[p])$ and $M(C[p'])$. Then $v$ is a difference node of $M(C[p])$ and $M(C[p'])$ by Lemma~\ref{lm:diff_char}. \qed \end{proof} \begin{example}\label{ex:diftup} For the dtla $M$ of Example~\ref{ex:Mex}, with the order $P=\{p_a,p_b\}$, we obtain that $\diftup(M)=\{(a,\sigma^n\angl{q,p_b})\mid n\in\Nat\}$. For the dtla $M$ of Example~\ref{ex:Mthree}, also with the order $P=\{p_a,p_b\}$, we obtain that $\diftup(M)$ consists of all pairs $(\angl{q_0,p_a},b)$, $\;(\gamma_1(\angl{q_1,p_a}),b)$, $\;(\gamma_1(\gamma_2(\angl{q_1,p_a})),b)$ and $(\gamma_1(\gamma_2(\gamma_3(a))),b)$ for $\gamma_1,\gamma_2,\gamma_3\in\{\sigma,\tau\}$. For the dtla $M$ of Example~\ref{ex:la-uniform} (which is the la-uniform version of the dtla of Example~\ref{ex:Mleaves}) it is not difficult to see that $\diff(M)=\{a,b\}\cup\{\angl{q_{yz},p_{yz}}\mid y,z\in\{a,b\}\}$, and that the set $\diftup(M)$ consists of the three 4-tuples $(a,a,b,b)$, $\;(a,b,a,b)$ and $(\angl{q_{aa},p_{aa}},\angl{q_{ab},p_{ab}},\angl{q_{ba},p_{ba}},\angl{q_{bb},p_{bb}})$, where we have taken the order $P=\{p_{aa},p_{ab},p_{ba},p_{bb}\}$. For the dtla $M$ of Example~\ref{ex:Mcounter}, $\diftup(M)=\{(e,o),(o,e)\}$. In the above examples, the components of the difference tuples are exactly the difference trees. As another example, let $\Sigma=\Delta=\{\sigma^{(1)},a^{(0)},b^{(0)},c^{(0)}\}$ and consider the dtla $M$ with $P=\{p_a,p_b,p_c\}$, $\delta(y)=p_y$ and $\delta(\sigma,p_y)=p_y$ for $y\in\{a,b,c\}$, $Q=\emptyset$, $A(p_a)=a$, $A(p_b)=\sigma(b)$ and $A(p_c)=\sigma(\sigma(c))$. Thus, for every $n\in\Nat$, $M$ translates $\sigma^na$ into $a$, $\sigma^nb$ into $\sigma b$, and $\sigma^nc$ into $\sigma\sigma c$. Since $a\sqcap\sigma b=\bot$, $\;a\sqcap\sigma\sigma c=\bot$ and $\sigma b\sqcap\sigma\sigma c=\sigma\bot$, we obtain that $\diff(M)=\{a,\sigma b,\sigma\sigma c,b,\sigma c\}$. Since $a\sqcap\sigma b\sqcap\sigma\sigma c=\bot$, we obtain that $\diftup(M)=\{(a,\sigma b,\sigma\sigma c)\}$. Thus, $b$ and $\sigma c$ are difference trees that are not components of a difference tuple (but are subtrees of such components). Note that there is a dtop with one state that is equivalent to $M$. \qed \end{example} In the next lemmas $N$ is a total dtop, equivalent to $M$. We assume that the unique look-ahead state of $N$ is $\bot$. So, $N^\circ$ translates input trees in $\T_\Sigma(\{\bot\})$, in particular $\Sigma$-contexts, into output trees in $\T_\Delta(Q_N\times\{\bot\})$; for a $\Sigma$-context $C$ we of course write $C$ instead of $C[\bot]$. The unique axiom $A_N(\bot)$ is denoted by $A_N$, a rule $q(a(x_1\ot\bot,\dots,x_k\ot\bot))\to\zeta$ is written $q(a(x_1,\dots,x_k))\to\zeta$ and $\zeta$ is denoted $\rhs_N(q,a)$. For a tree $t\in\T_\Delta(Q_N\times \{\bot\})$ we define the pattern $t\Phi\in \mathcal{P}_\Delta$ by $t\Phi=t[\angl{q,\bot}\leftarrow\bot\mid q\in Q_N]$; similarly, for $t\in\T_\Delta(Q_N(X))$ we define $t\Phi=t[q(x_i)\leftarrow\bot\mid q\in Q_N,i\in\Nat]$. Let $M$ be a canonical dtpla and $N$ a dtop such that $\sem{M}=\sem{N}$. We first show that the translation of an input tree by $M$ is always ahead of its translation by $N$, in a uniform way. An \emph{aheadness mapping} from $N$ to $M$ is a function $\varphi: Q_N\times P_M\to \T_\Delta(Q_M\times P_M)$ such that for every $C\in\C_\Sigma$ and $p\in P_M$, \begin{equation}\label{eq:ahead} M(C[p]) = N(C)[\angl{q,\bot}\leftarrow\varphi(q,p)\mid q\in Q_N]. \end{equation} Note that $\varphi(q,p)$ must be in $\T_\Delta(\{\angl{\bar{q},p}\mid\bar{q}\in Q_M,\rho_M(\bar{q})=p\})$. Intuitively, $\varphi$ defines the exact amount in which $M$ is ahead of $N$, which is independent of $C$. \begin{lemma}\label{lm:aheadness-mapping} Let $M$ be a canonical dtpla and $N$ a dtop such that $\sem{M}=\sem{N}$. Then there is a unique aheadness mapping from $N$ to~$M$. \end{lemma} \begin{proof} We first show that $M$ is ahead of $N$, i.e., that all output symbols produced by $N$ on a given input context are also produced by $M$. Let $p\in P_M$ and let $C$ be a $\Sigma$-context. \smallskip {\bf Claim~1.}\quad $V_d(N(C))\subseteq V_d(M(C[p]))$ for every $d\in\Delta$. \\ Equivalently, $N(C)\Phi\sqsubseteq M(C[p])$. \smallskip Proof: By induction on the length of the nodes of $N(C)$. Let $v$ be a node of $N(C)$ with label $d\in\Delta$. Since $v$'s proper ancestors are in $V_\Delta(N(C))$ it follows by induction that $v$ is a node of $M(C[p])$. Consider an arbitrary $s\in \sem{p}_M$. By Lemma~\ref{lm:semantics}, $v$~has label $d$ in $N(C[s])$. Since $\sem{M}=\sem{N}$, $M(C[s])=N(C[s])$ and so $v$ has label $d$ in $M(C[s])$. Suppose that $v$ does not have label $d$ in $M(C[p])$. Then, again by Lemma~\ref{lm:semantics}, $v$ must have some label $\angl{q,p}$ in $M(C[p])$ such that $q_M(s)$ has root label $d$. Since this holds for every $s\in\sem{p}_M$, we obtain that $\rlabs_M(q)=\{d\}$ contradicting the fact that $M$ is earliest. Note that, since $M$ is la-uniform, $\rho_M(q)=p$ by Lemma~\ref{lm:prop_la-uniform}(3) and hence $\sem{p}_M=\dom(\sem{q}_M)$ by Lemma~\ref{lm:prop_la-uniform}(1). This proves the claim. \smallskip Next we show that the amount in which $M$ is ahead of $N$, is independent of $C$. Let $C_1,C_2$ be $\Sigma$-contexts, $v_1,v_2\in\Nat_+^*$ and $q\in Q_N$. \smallskip {\bf Claim~2.}\quad If $N(C_1)/v_1=N(C_2)/v_2=\angl{q,\bot}$, \\ then $M(C_1[p])/v_1=M(C_2[p])/v_2$. \smallskip Proof: By Claim~1, $v_i$ is a node of $M(C_i[p])$. Let $t_i\in \T_\Delta(Q_M\times\{p\})$ denote the tree $M(C_i[p])/v_i$. For every $s\in\sem{p}_M$, $N(C_1[s])/v_1=N(C_2[s])/v_2=q_N(s)$ by Lemma~\ref{lm:semantics}, and so $M(C_1[s])/v_1=M(C_2[s])/v_2$. Hence, again by Lemma~\ref{lm:semantics}, $t_1\Psi_s=t_2\Psi_s$ for all $s\in \sem{p}_M$, where $\Psi_s=[\angl{q,p}\leftarrow q_M(s)\mid q\in Q_M]$. Suppose that $t_1\not=t_2$. Then there is a leaf $v$ of, e.g., $t_1$ with label $\angl{q_1,p}$ such that $v$ is a node of $t_2$ with $t_2/v\not= \angl{q_1,p}$. If the root label of $t_2/v$ is $d\in\Delta$, then $q_{1M}(s)$ has root label $d$ for all $s\in\sem{p}_M$, contradicting the fact that $M$ is earliest. If $t_2/v$ equals $\angl{q_2,p}$ with $q_1\neq q_2$, then $q_{1M}(s)=q_{2M}(s)$ for all $s\in\sem{p}_M$. Since, as observed in the proof of Claim~1, $\sem{p}_M$ is the domain of both $\sem{q_1}_M$ and $\sem{q_2}_M$ by Lemma~\ref{lm:prop_la-uniform}, we obtain that $\sem{q_1}_M=\sem{q_2}_M$, contradicting the fact that $M$ is canonical. This proves the claim. \smallskip An aheadness mapping from $N$ to $M$ can now be defined as follows. Let $q\in Q_N$ and $p\in P_M$. By Lemma~\ref{lm:reach}, there is a $\Sigma$-context $C$ such that $N(C)$ has a node $v$ labeled $\angl{q,\bot}$. By Claim~1, $v$ is a node of $M(C[p])$ and we define $\varphi(q,p)=M(C[p])/v$. By Claim~2, the definition of $\varphi$ does not depend on $C$ and $v$. It follows from Claim~1 that if $N(C)=t[\angl{q_1,\bot},\dots,\angl{q_r,\bot}]$ with $t\in\mathcal{P}_\Delta$ and $r=|V_\bot(t)|$, then $M(C[p])=t[M(C[p])/v_1,\dots,M(C[p])/v_r]$ where $v_i$ is the $i$th occurrence of $\bot$ in $t$, and hence $M(C[p])=t[\varphi(q_1,p),\dots,\varphi(q_r,p)]$. Thus, $\varphi$ satisfies Equation~(\ref{eq:ahead}): the requirement for an aheadness mapping. Obviously, if $\varphi'$ is an aheadness mapping from $N$ to $M$, and $C$ is a $\Sigma$-context such that $N(C)$ has a node $v$ labeled $\angl{q,\bot}$, then $M(C[p])/v=\varphi'(q,p)$ for every $p\in P_M$, by Equation~(\ref{eq:ahead}). Thus $\varphi'=\varphi$, which shows that $\varphi$ is the unique aheadness mapping from $N$ to $M$. \qed \end{proof} \begin{lemma}\label{lm:dtladtopprop1} Let $M$ be a canonical dtpla and $N$ a dtop such that $\sem{M}=\sem{N}$. Let $\varphi$ be the aheadness mapping from $N$ to $M$. For every $s\in\T_\Sigma$ and $q\in Q_N$, if $\delta_M(s)=p$, then $$q_N(s)=\varphi(q,p)[\angl{\bar{q},p}\leftarrow {\bar{q}}_M(s)\mid \bar{q}\in Q_M].$$ \end{lemma} \begin{proof} By Lemma~\ref{lm:reach}, there are $C,v$ such that $N(C)/v=\angl{q,\bot}$. It follows from Equation~(\ref{eq:ahead}) that $M(C[p])/v=\varphi(q,p)$. Since $M$ and $N$ are equivalent, $N(C[s])=M(C[s])$. Applying Lemma~\ref{lm:semantics} twice, we obtain $q_N(s)=N(C[s])/v=M(C[s])/v= (M(C[p])/v)[\angl{\bar{q},p}\leftarrow {\bar{q}}_M(s)\mid \bar{q}\in Q_M]$, which proves the equation. \qed \end{proof} In the next lemma we prove our basic intuition that the output of $N$ on input $C$ is the largest common prefix of the outputs of $M$ on all inputs $C[p]$, $p\in P$, such that the difference tuples of $M$ are stored in the states of $N$. \begin{lemma}\label{lm:dtladtopprop2} Let $M$ be a canonical dtpla and $N$ an earliest dtop such that $\sem{M}=\sem{N}$. Let $C\in\C_\Sigma$ and let $\varphi$ be the aheadness mapping from $N$ to $M$. Then \begin{enumerate} \item[(1)] $N(C)\Phi = \pref(M,C)$ and, for every $v\in\Nat^*_+$, $q\in Q_N$ and $p\in P_M$, \item[(2)] if $N(C)/v=\angl{q,\bot}$ then $\varphi(q,p)=M(C[p])/v$. \end{enumerate} \end{lemma} \begin{proof} By the definition of aheadness mapping (Equation~(\ref{eq:ahead})), $N(C)\Phi\sqsubseteq M(C[p])$ for every $p\in P_M$ (cf. Claim~1 in the proof of Lemma~\ref{lm:aheadness-mapping}), and so $N(C)\Phi\sqsubseteq\pref(M,C)$ by the definition of greatest lower bound. To show equality, we prove for every node $v$ of $N(C)\Phi$ with label $\bot$ that $v$ has label $\bot$ in $\pref(M,C)$ too. Let $N(C)/v=\angl{q,\bot}$ for $q\in Q_N$. Then, by Equation~(\ref{eq:ahead}), $M(C[p])/v=\varphi(q,p)$ for every $p\in P_M$ (which proves statement~(2) of this lemma). Suppose now that $v$ has label $d\in\Delta$ in $\pref(M,C)$. Then $v$ has label $d$ in every $M(C[p])$, and so the root symbol of $\varphi(q,p)$ is $d$ for every $p\in P_M$. It then follows from Lemma~\ref{lm:dtladtopprop1} that $q_N(s)$ has root label $d$ for every $s\in\T_\Sigma$, contradicting the fact that $N$ is earliest. Hence $v$ has label $\bot$ in $\pref(M,C)$. \qed \end{proof} If a dtpla $M$ is equivalent to a dtop, then it is equivalent to a canonical dtop by Theorem~\ref{th:uniform_earliest}. By~\cite[Theorem~15]{DBLP:journals/jcss/EngelfrietMS09}, two canonical dtops are equivalent if and only if they are the same (modulo a renaming of their states). Thus, if a dtpla $M$ is equivalent to a dtop, then it is equivalent to a \emph{unique} canonical dtop $\td(M)$. In the next three lemmas we give another proof of this, and we show that the dtop $\td(M)$ can be constructed from $M$ and $\diftup(M)$. We start by showing that $Q_{\td(M)}$ can be identified with $\diftup(M)$, assuming $M$ to be canonical too. \begin{lemma}\label{lm:Qdiff} Let $M,N$ be equivalent canonical dtlas such that $M$ is a dtpla and $N$ is a dtop. Let $\varphi: Q_N\times P_M\to T_\Delta(Q_M\times P_M)$ be the aheadness mapping from $N$ to $M$, and let for $q\in Q_N$ $$\psi(q)=(\varphi(q,\hat{p}_1),\dots,\varphi(q,\hat{p}_n))$$ where $P_M=\{\hat{p}_1,\dots,\hat{p}_n\}$. Then $\psi$ is a bijection between $Q_N$ and $\diftup(M)$. \end{lemma} \begin{proof} The proof is in three parts. \smallskip (i) For all $q\in Q_N$, $\psi(q)\in\diftup(M)$. \smallskip \noindent Proof: Let $C,v$ be such that $N(C)/v=\angl{q,\bot}$, by Lemma~\ref{lm:reach}. By Lemma~\ref{lm:dtladtopprop2}, $v\in V_\bot(\pref(M,C))$ and $M(C[\hat{p}_i])/v=\varphi(q,\hat{p}_i)$ for every $i\in[n]$. That shows that $\psi(q)\in\diftup(M)$. \smallskip (ii) For every $\bar{t}\in\diftup(M)$ there exists $q\in Q_N$ such that $\psi(q)=\bar{t}$. \smallskip \noindent Proof: If $(t_1,\dots,t_n)\in\diftup(M)$ then there are $C,v$ such that $\pref(M,C)/v=\bot$ and $M(C[\hat{p}_i])/v=t_i$ for $i\in[n]$. By Lemma~\ref{lm:dtladtopprop2}, $N(C)/v=\angl{q,\bot}$ for some $q\in Q_N$, and $M(C[\hat{p}_i])/v=\varphi(q,\hat{p}_i)$. Thus, $t_i=\varphi(q,\hat{p}_i)$ for $i\in[n]$. \smallskip (iii) If $\psi(q_1)=\psi(q_2)$ then $q_1=q_2$. \smallskip \noindent Proof: Let $\psi(q_1)=\psi(q_2)$. By Lemma~\ref{lm:dtladtopprop1}, $q_{1N}(s)=q_{2N}(s)$ for all $s\in\T_\Sigma$. In other words, $\sem{q_1}_N=\sem{q_2}_N$ and hence $q_1=q_2$ because $N$ is canonical. \qed \end{proof} \begin{corollary}\label{co:difffin} Let $M$ be a total dtla. If $M$ is equivalent to a dtop, then $\diff(M)$ is finite. \end{corollary} \begin{proof} If $M$ is a dtop, then $\diff(M)=\emptyset$. Now let $M$ be a dtpla. By Lemma~\ref{lm:make_uniform} we may assume that $M$ is la-uniform. Let $\can(M)$ be a canonical dtpla equivalent to~$M$, which exists by Theorem~\ref{th:uniform_earliest}. If $M$, and hence $\can(M)$, is equivalent to a (total) dtop, then $\can(M)$ is equivalent to a canonical dtop, by Theorem~\ref{th:uniform_earliest}. By Lemmas~\ref{lm:aheadness-mapping} and~\ref{lm:Qdiff}, $\diftup(\can(M))$ is finite, and so $\diff(\can(M))$ is finite by Lemma~\ref{lm:twodiff}. Hence $\diff(M)$ is finite by Theorem~\ref{th:uniform_earliest}, because $\maxdiff(M)\leq\maxdiff(\can(M))+\sumfix(M)$. \qed \end{proof} The reverse of this corollary does not hold, see Example~\ref{ex:Mcounter}. For an la-uniform dtla $M$ and a tree $t\in\T_\Delta(Q(X))$ we define $t\Omega\in\T_\Delta(Q\times P)$ by $t\Omega=t[q(x_i)\leftarrow\angl{q,\rho(q)}\mid q\in Q,i\in\Nat]$. In particular, for a total dtop $N$, $t\Omega=t[q(x_i)\leftarrow\angl{q,\bot}\mid q\in Q_N,i\in\Nat]$. Recall also the definition of the pattern $t\Phi\in\mathcal{P}_\Delta$ for $t\in\T_\Delta(Q_N(X))$ or $t\in\T_\Delta(Q_N\times \{\bot\})$ after Example~\ref{ex:diftup}. Next we show, for a canonical dtpla $M$, how to compute the axiom of $\td(M)$, representing the states of $\td(M)$ by difference tuples. \begin{lemma}\label{lm:dtladtopprop2a} Let $M,N$ be equivalent canonical dtlas such that $M$ is a dtpla and $N$ is a dtop. Let $\varphi$ be the aheadness mapping from $N$ to $M$. Then \begin{enumerate} \item[(1)] $A_N\Phi=\sqcap\{A_M(p)\Omega\mid p\in P_M\}$, and for every $v\in\Nat^*_+$, $q\in Q_N$ and $p\in P_M$, \item[(2)] if $A_N/v=q(x_0)$ then $\varphi(q,p)=A_M(p)\Omega/v$. \end{enumerate} \end{lemma} \begin{proof} By the semantics of $N^\circ$, $N(\bot)=A_N\Omega$ and by the semantics of $M^\circ$, $M(p)=A_M(p)\Omega$ for every $p\in P_M$. Hence by Lemma~\ref{lm:dtladtopprop2}(1), with $C=\bot$, $A_N\Phi=A_N\Omega\Phi=N(\bot)\Phi=\pref(M,\bot)=\sqcap\{M(p)\mid p\in P_M\}= \sqcap\{A_M(p)\Omega\mid p\in P_M\}$. If $A_N/v=q(x_0)$ then $N(\bot)/v=A_N\Omega/v=\angl{q,\bot}$ and so by Lemma~\ref{lm:dtladtopprop2}(2), with $C=\bot$, $\varphi(q,p)=M(p)/v=A_M(p)\Omega/v$ for every $p\in P_M$. \qed \end{proof} We note that it is easy to see that $\sqcap\{A_M(p)\Omega\mid p\in P_M\}=\sqcap\{A_M(p)\mid p\in P_M\}$, and hence $A_N\Phi=\sqcap\{A_M(p)\mid p\in P_M\}.$ Let $M$ be an la-uniform dtla, $Q_N$ be a finite set and $\varphi: Q_N\times P_M\to \T_\Delta(Q_M\times P_M)$ be a mapping such that $\varphi(q,p)\in\T_\Delta(\{\angl{\bar{q},p}\mid\bar{q}\in Q_M,\rho_M(\bar{q})=p\})$ for every $q\in Q_N$ and $p\in P_M$. Then we define for every $q\in Q_N$, $a\in\Sigma^{(k)}$ and $p_1,\dots,p_k\in P_M$, the tree $$\rhs_{M,\varphi}(q,a,p_1,\dots,p_k) = \varphi(q,p)[\angl{\bar{q},p}\leftarrow\rhs_M(\bar{q},a,p_1,\dots,p_k)\mid\bar{q}\in Q_M]$$ where $p=\delta_M(a,p_1,\dots,p_k)$. Note that by the above condition on $\varphi(q,p)$, the right-hand side of the equation is always defined and is a tree in $\T_\Delta(Q_M(X_k))$; moreover, if it has a subtree $q'(x_i)$, then $\rho_M(q')=p_i$. Finally we show how to compute the rules of $\td(M)$. \begin{lemma}\label{lm:dtladtopprop2b} Let $M,N$ be equivalent canonical dtlas such that $M$ is a dtpla and $N$ is a dtop. Let $\varphi$ be the aheadness mapping from $N$ to $M$. \begin{enumerate} \item[$(1)$] For every $q\in Q_N$ and $a\in \Sigma^{(k)}$, $$\rhs_N(q,a)\Phi=\sqcap\{\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Omega\mid p_1,\dots,p_k\in P_M\}.$$ \item[$(2)$] Let $q\in Q_N$, $a\in \Sigma^{(k)}$ and $i\in[k]$. For $j\in[k]$, $j\neq i$, let $s_j\in\T_\Sigma$ and $p_j=\delta_M(s_j)$. Let $\Psi_{iM}= [\bar{q}(x_j)\leftarrow\bar{q}_M(s_j)\mid\bar{q}\in Q_M,j\in[k],j\neq i]\Omega$.\footnote{Note that $\Psi_{iM}$ depends on the trees $s_j$. Note also that, through $\Omega$, it replaces every $\bar{q}(x_i)$ by $\angl{\bar{q},\rho_M(\bar{q})}$.} \\ For every $v\in V_{\bot}(\rhs_N(q,a)\Phi)$, \begin{enumerate} \item[$(a)$] $\rhs_N(q,a)/v\in Q_N(\{x_i\})$ if and only if $$v\in V_{\bot}(\sqcap\{\rhs_{M,\varphi}(q,a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)\Psi_{iM} \mid p\in P_M\}),$$ and \item[$(b)$] for every $\bar{q}\in Q_N$ and $p\in P_M$, if $\rhs_N(q,a)/v=\bar{q}(x_i)$, then $$\varphi(\bar{q},p)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)\Psi_{iM}/v.$$ \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} (1) We first show that $\rhs_N(q,a)\Phi\sqsubseteq\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$ for all $p_1,\dots,p_k$ $\in P_M$. Consider arbitrary trees $s_1,\dots,s_k\in\T_\Sigma$ with $s_i\in\sem{p_i}_M$ for $i\in[k]$, and let $s=a(s_1,\dots,s_k)$. By Lemma~\ref{lm:elem}(4), $$q_N(s)=\rhs_N(q,a)\Psi_N$$ where $\Psi_N=[\bar{q}(x_i)\leftarrow {\bar{q}}_N(s_i)\mid \bar{q}\in Q_N, i\in[k]]$. On the other hand, by Lemma~\ref{lm:dtladtopprop1} and Lemma~\ref{lm:elem}(4), $$q_N(s)=\varphi(q,p)[\angl{\bar{q},p}\leftarrow \rhs_M(\bar{q},a,p_1,\dots,p_k)\mid \bar{q}\in Q_M]\Psi_M$$ where $p=\delta_M(a,p_1,\dots,p_k)$ and $\Psi_M= [\bar{q}(x_i)\leftarrow {\bar{q}}_M(s_i)\mid \bar{q}\in Q_M, i\in[k]]$. Thus \begin{equation}\label{eq:rhss} \rhs_N(q,a)\Psi_N = \rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_M. \end{equation} We have to prove that every node of $\rhs_N(q,a)$ with label $d\in\Delta$ is also a node of $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$, with the same label. We prove this by induction on the length of the nodes of $\rhs_N(q,a)$, similar to the proof of Claim~1 of Lemma~\ref{lm:aheadness-mapping}. Let $v$ be a node of $\rhs_N(q,a)$ with label $d\in\Delta$. By induction, $v$ is a node of $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$. Now suppose that $v$ does not have label $d$ in $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$. Then Equation~(\ref{eq:rhss}) implies that $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)/v=\bar{q}(x_i)$ and that ${\bar{q}}_M(s_i)$ has root label $d$. Since that holds for arbitrary $s_i\in \sem{p_i}_M$, it contradicts the fact that $M$ is earliest. Note that since $M$ is la-uniform, $\sem{p_i}_M$ is the domain of $\sem{\bar{q}}_M$. Now let $\Pi=\{\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Omega\mid p_1,\dots,p_k\in P_M\}$. By the above, $\rhs_N(q,a)\Phi\sqsubseteq\sqcap\Pi$. It remains to prove for every node $v$ of $\rhs_N(q,a)\Phi$ with label $\bot$ that $v$ has label $\bot$ in $\sqcap\Pi$ too. Let $\rhs_N(q,a)/v=\bar{q}(x_i)$. Assume now that $v$ has label $d\in\Delta$ in $\sqcap\Pi$. Then $v$ has label $d$ in every tree $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$. Equation~(\ref{eq:rhss}) then implies that $\bar{q}_N(s_i)$ has root label $d$ for every $s_i\in\sem{p_i}_M$. Since this holds for every $p_i\in P_M$, $\bar{q}_N(s)$ has root label $d$ for every $s\in\T_\Sigma$, which contradicts the fact that $N$ is earliest. (2) Let the substitution $\Psi_{iN}$ be defined similar to $\Psi_{iM}$, i.e., $\Psi_{iN}=[\bar{q}(x_j)\leftarrow\bar{q}_N(s_j)\mid\bar{q}\in Q_N,j\in[k],j\neq i]\Omega$. Let $C,u$ be such that $N(C)/u=\angl{q,\bot}$, by Lemma~\ref{lm:reach}, and consider the context $C_i=C[a(s_1,\dots,s_{i-1},\bot,s_{i+1},\dots,s_k)]$. By Lemma~\ref{lm:semantics} and Lemma~\ref{lm:elem}(4), $N(C_i)=N(C)[\angl{\bar{q},\bot}\leftarrow\rhs_N(\bar{q},a)\mid\bar{q}\in Q_N]\Psi_{iN}$. Hence \begin{equation}\label{eq:rhsN} N(C_i)/u=\rhs_N(q,a)\Psi_{iN}. \end{equation} Note that for every $v\in\Nat_+^*$, $\rhs_N(q,a)/v\in Q_N(\{x_i\})$ if and only if $v$ has label $\bot$ in $\rhs_N(q,a)\Psi_{iN}\Phi$. By Lemma~\ref{lm:dtladtopprop2}(1), $N(C_i)\Phi=\pref(M,C_i)$. Consequently, $\rhs_N(q,a)/v\in Q_N(\{x_i\})$ if and only if $v\in V_{\bot}(\pref(M,C_i)/u)$. It should be clear that $\pref(M,C_i)/u=\sqcap\{M(C_i[p])/u\mid p\in P_M\}$. By Lemma~\ref{lm:semantics} and Lemma~\ref{lm:elem}(4), $M(C_i[p])$ equals $M(C[p'])[\angl{\bar{q},p'}\leftarrow \rhs_M(\bar{q},a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)\mid\bar{q}\in Q_M]\Psi_{iM}$ for every $p\in P_M$, where $p'=\delta_M(a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)$. Hence, since $M(C[p'])/u=\varphi(q,p')$ by Lemma~\ref{lm:dtladtopprop2}(2), we obtain for every $p\in P_M$ that \begin{equation}\label{eq:etaM} M(C_i[p])/u=\rhs_{M,\varphi}(q,a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)\Psi_{iM}, \end{equation} which shows that $\rhs_N(q,a)/v\in Q_N(\{x_i\})$ if and only if $v$ has label $\bot$ in the largest common prefix of all $\rhs_{M,\varphi}(q,a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)\Psi_{iM}$ with $p\in P_M$. If $\rhs_N(q,a)/v=\bar{q}(x_i)$ and $p\in P_M$, then $N(C_i)/uv=\angl{\bar{q},\bot}$ by Equation~(\ref{eq:rhsN}), and so $\varphi(\bar{q},p)=M(C_i[p])/uv$ by Lemma~\ref{lm:dtladtopprop2}(2). Then Equation~(\ref{eq:etaM}) implies the required result. \qed \end{proof} Note that if $a\in\Sigma^{(0)}$, then (1) of this lemma means that $\rhs_N(q,a)=\rhs_{M,\varphi}(q,a)$. Note also that if $a\in\Sigma^{(1)}$, then no $s_j$ need to be chosen in (2) and so $\Psi_{1M}=\Omega$; moreover, (2)(a) need not be checked because it is immediate from~(1). The last three lemmas show that every dtpla $M$ that is equivalent to a dtop, is equivalent to a unique canonical dtop $\td(M)$, modulo a renaming of states. Based on these same lemmas, we can now construct $\td(M)$ from any given canonical dtpla $M$ for which $\diftup(M)$ is a given finite set. The construction returns the answer `no' if $M$ is not equivalent to any dtop. We construct the total dtop $N=\td(M)$, if it exists, by taking $Q_N=\diftup(M)$, by defining the mapping $\varphi: Q_N\times P_M\to \T_\Delta(Q_M\times P_M)$ as $\varphi((t_1,\dots,t_n),\hat{p}_i)=t_i$ for $i\in[n]$ (in accordance with Lemma~\ref{lm:Qdiff}), and by constructing the axiom and rules of $N$ according to Lemmas~\ref{lm:dtladtopprop2a} and~\ref{lm:dtladtopprop2b}, respectively (i.e., by viewing the numbered statements of these lemmas as definitions). In~(2) of Lemma~\ref{lm:dtladtopprop2b} we choose $s_j$ arbitrarily but fixed (e.g., $s_j=a$ for all $j$, with $a\in\Sigma^{(0)}$). It may be that the construction of an axiom or a rule fails when a possible state occurring in it (which is a tuple in $\T_\Delta(Q_M\times P_M)^n$) is not a difference tuple of $M$. Then the construction returns the answer `no'. Also, the construction of a rule can fail when a node $v\in V_{\bot}(\rhs_N(q,a)\Phi)$ is an element of $V_{\bot}(\sqcap\{\rhs_{M,\varphi}(q,a,p_1,\dots,p_{i-1},p,p_{i+1},\dots,p_k)\Psi_{iM} \mid p\in P_M\})$ for several $i$ or for no $i$, see~(2)(a) of Lemma~\ref{lm:dtladtopprop2b}. Again, the construction then returns the answer `no'. If the construction of the dtop $N$ succeeds, then it remains to test whether $M$ and $N$ are equivalent (because, by Lemmas~\ref{lm:Qdiff}, \ref{lm:dtladtopprop2a} and~\ref{lm:dtladtopprop2b}, if $M$ is equivalent to a dtop then it is equivalent to $N$). If they are not equivalent then the answer is `no'. If they are equivalent then the construction returns the dtop $N=\td(M)$. Note that equivalence of dtlas is decidable by~\cite{DBLP:journals/actaC/Esik81} (see also~\cite[Corollary~19]{DBLP:journals/jcss/EngelfrietMS09}). Unfortunately, we do not know whether it is decidable if $\diftup(M)$ is finite, and whether it can be computed when it is finite. In the next theorem we show that, to determine whether a total dtla $M$ is equivalent to a dtop, it suffices to have an upper bound for $\maxdiff(M)$. This is our first main result. Recall from Section~\ref{sec:difftree} that a number $h(M)\in\Nat$ is a \emph{difference bound} for $M$ if the following holds: if $\diff(M)$ is finite, then $\maxdiff(M)\leq h(M)$. \begin{theorem}\label{th:alg} It is decidable for a given total dtla $M$ and a given difference bound for $M$ whether there exists a dtop $N$ such that $\sem{M}=\sem{N}$, and if so, such a dtop $N$ can be constructed. \end{theorem} \begin{proof} Let $M$ be a total dtla and let $h(M)$ be a difference bound for $M$. We may, of course, assume that $M$ is a dtpla. By Lemma~\ref{lm:make_uniform}, we may assume that $M$ is la-uniform. Moreover, by Theorem~\ref{th:uniform_earliest} we may assume that $M$ is canonical. In fact, if $\can(M)$ is the canonical dtpla equivalent to $M$ as constructed in the proof of that theorem, then $h(M)+\sumfix(M)$ is a (computable) difference bound for $\can(M)$, as shown after Theorem~\ref{th:uniform_earliest}. So, let $M$ be a canonical dtpla and let $h(M)$ be a difference bound for $M$. By Lemma~\ref{lm:twodiff}, this means that if $\diftup(M)$ is finite, then the height of the components of the difference tuples of $M$ is at most $h(M)$. We now decide whether $M$ is equivalent to a dtop by constructing $\td(M)$ as described before this theorem. However, since $\diftup(M)$ is not given, we construct $N=\td(M)$ incrementally, using a variable $Q_N$ to accumulate its states (which are all assumed to be reachable). In accordance with Lemma~\ref{lm:Qdiff} we take $Q_N\subseteq \T_\Delta(Q_M\times P_M)^n$ and $\varphi((t_1,\dots,t_n),\hat{p}_i)=t_i$ for $i\in[n]$. We first construct the axiom $A_N$ according to Lemma~\ref{lm:dtladtopprop2a} and initialize the set $Q_N$ with the states, i.e., the tuples in $\T_\Delta(Q_M\times P_M)^n$, that occur in that axiom. If the height of one of the components of one of those tuples is larger than $h(M)$, then either $\diftup(M)$ is infinite or that tuple is not a difference tuple of $M$, and we stop the construction with answer `no', indicating that $M$ is not equivalent to any dtop. Then, repeatedly, for every $q\in Q_N$ and $a\in\Sigma$ we construct $\rhs_N(q,a)$ according to Lemma~\ref{lm:dtladtopprop2b}, and we add to $Q_N$ the states that occur in that right-hand side. If the construction of $\rhs_N(q,a)$ fails or if the height of one of the components of its states is larger than $h(M)$, then the answer is `no'. If the construction of the dtop $N$ succeeds, then it remains to test whether $M$ and $N$ are equivalent, as described before this theorem. We finally observe that, due to Condition~(2) of Lemma~\ref{lm:dtladtopprop2a} and Condition~(2)(b) of Lemma~\ref{lm:dtladtopprop2b}, for every state $q\in Q_N$ and every $p\in P_M$ the tree $\varphi(q,p)$ is in $\T_\Delta(\{\angl{\bar{q},p}\mid\bar{q}\in Q_M,\rho_M(\bar{q})=p\})$, and hence $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$ is always defined and in $\T_\Delta(Q_M(X_k))$. \qed \end{proof} Several examples of the algorithm in the proof of Theorem~\ref{th:alg} will be given in Section~\ref{sec:algex}. The algorithm is useful for the designer of a dtla, cf. Examples~\ref{ex:Mex}--\ref{ex:Mcounter} in Section~\ref{sec:difftree}. If you have designed a dtla $M$ to satisfy a specification of its translation $\sem{M}$, then you usually also know how to specify the output $M(C[p])$, for every context $C$ and every look-ahead state $p$. From this it is probably straightforward for you to obtain a specification of $\diff(M)$. If $\diff(M)$ is infinite (which you probably also can see), then $M$ is not equivalent to a dtop by Corollary~\ref{co:difffin}. If it is finite, then you can determine $\maxdiff(M)$ (or an upper bound for it), and hence you have determined a difference bound for $M$. So now you can use the algorithm of Theorem~\ref{th:alg} to find out whether $M$ is equivalent to a dtop, and if so, to construct such a dtop. On the other hand, if you are \emph{not} able to determine a difference bound for $M$, then you can still use the algorithm of Theorem~\ref{th:alg}, without the tests on the height of the components of the states of $N$. If $M$ is equivalent to a dtop, then the algorithm will construct such a dtop; otherwise, the algorithm may not halt (as will be shown in Example~\ref{ex:alg_non_halt}).\footnote{The existence of such an algorithm is obvious from the fact that the equivalence of dtlas is decidable: one can just enumerate all dtops $N$ and test the equivalence of $M$ and $N$. Obviously, our algorithm is more efficient.} From Theorem~\ref{th:alg} we immediately obtain the following result. \begin{corollary}\label{co:goal} Let $\U$ be a class of total dtlas with the following property. \begin{enumerate} \item [(H)] There is a computable mapping $h: \U\to\Nat$ such that, for every $M\in\U$, \\ $h(M)$ is a difference bound for $M$. \end{enumerate} Then it is decidable for a given dtla $M\in\U$ whether there exists a dtop $N$ such that $\sem{M}=\sem{N}$, and if so, such a dtop $N$ can be constructed. \end{corollary} \begin{proof} For a dtla $M$ in $\U$, compute the difference bound $h(M)$ and run the algorithm of Theorem~\ref{th:alg}. \qed \end{proof} Let $\U$ be the class of total ultralinear b-erasing dtlas. Our goal in Sections~\ref{sec:origins}--\ref{sec:graphs} is to prove that $\U$ has Property~(H), i.e., to compute a difference bound for the dtlas in the class $\U$. \subsection{Avoiding the final equivalence test} \label{sec:avoid-eq} This subsection can be skipped by the reader who is satisfied with the algorithm in the proof of Theorem~\ref{th:alg}. We will show that the equivalence test of $M$ and $N$ at the end of the algorithm can be realized by a simple direct test. This possibility is based on the following lemma. \begin{lemma}\label{lm:NequiM} Let $M$ be a canonical dptla and let $N$ be a total dtop such that $Q_N\subseteq\T_\Delta(Q_M\times P_M)^n$. Let $\varphi: Q_N\times P_M\to \T_\Delta(Q_M\times P_M)$ be such that $\varphi((t_1,\dots,t_n),\hat{p}_i)=t_i$ for $i\in[n]$. If $N$ satisfies Conditions~(1) and~(2) of Lemma~\ref{lm:dtladtopprop2a} and Conditions~(1) and~(2)(b) of Lemma~\ref{lm:dtladtopprop2b},\footnote{To be precise, $N$ should satisfy Condition~(2)(b) of Lemma~\ref{lm:dtladtopprop2b} for all possible choices of the trees $s_j\in\T_\Sigma$. Note also that, since all states of $N$ are reachable, the tree $\varphi(q,p)$ is in $\T_\Delta(\{\angl{\bar{q},p}\mid\bar{q}\in Q_M,\rho_M(\bar{q})=p\})$ for every state $q\in Q_N$ and every $p\in P_M$, as observed at the end of the proof of Theorem~\ref{th:alg}.} then $N$ is equivalent to $M$. \end{lemma} \begin{proof} We first show, by structural induction on $s\in\T_\Sigma$, that if $\delta_M(s)=p$, then $q_N(s)=\varphi(q,p)[\angl{\tilde{q},p}\leftarrow {\tilde{q}}_M(s)\mid \tilde{q}\in Q_M]$ for all $q\in Q_N$, cf. Lemma~\ref{lm:dtladtopprop1}. Let $s=a(s_1,\dots,s_k)$ with $\delta_M(s_i)=p_i$ for $i\in[k]$ and $\delta_M(a,p_1,\dots,p_k)=p$. From Lemma~\ref{lm:elem}(4) we obtain that $q_N(s)=\rhs_N(q,a)[\bar{q}(x_i)\leftarrow \bar{q}_N(s_i)\mid\bar{q}\in Q_N, i\in[k]]$. By induction, $\bar{q}_N(s_i)= \varphi(\bar{q},p_i)[\angl{\tilde{q},p_i}\leftarrow {\tilde{q}}_M(s_i)\mid \tilde{q}\in Q_M]$. If $\rhs_N(q,a)/v=\bar{q}(x_i)$, then we have $\varphi(\bar{q},p_i)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_{iM}/v$ by Condition~(2)(b) of Lemma~\ref{lm:dtladtopprop2b}, and consequently $\bar{q}_N(s_i)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_M/v$, where $\Psi_M=[\tilde{q}(x_j)\leftarrow \tilde{q}_M(s_j)\mid \tilde{q}\in Q_M, j\in[k]]$. Hence $q_N(s)=t\Psi_M$ with $$t=\rhs_N(q,a)\Phi[\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)/v_1,\dots,\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)/v_r]$$ where $v_1,\dots,v_r$ are the occurrences of $\bot$ in $\rhs_N(q,a)\Phi$, from left to right. By Condition~(1) of Lemma~\ref{lm:dtladtopprop2b}, $\rhs_N(q,a)\Phi\sqsubseteq\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$ from which it follows that $t=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)$. So, $q_N(s)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_M= \varphi(q,p)[\angl{\tilde{q},p}\leftarrow\rhs_M(\tilde{q},a,p_1,\dots,p_k)\mid\tilde{q}\in Q_M]\Psi_M= \varphi(q,p)[\angl{\tilde{q},p}\leftarrow {\tilde{q}}_M(s)\mid \tilde{q}\in Q_M]$ where the last equality is by Lemma~\ref{lm:elem}(4). We now show with a similar proof that $N(s)=M(s)$ for all $s\in\T_\Sigma$, i.e., that $N$ is equivalent to $M$. By Lemma~\ref{lm:elem}(3), $N(s)=A_N[q(x_0)\leftarrow q_N(s)\mid q\in Q_N]$ and $M(s)=A_M(p)[\tilde{q}(x_0)\leftarrow \tilde{q}_M(s)\mid \tilde{q}\in Q_M]$ where $p=\delta_M(s)$. By the above, $q_N(s)=\varphi(q,p)[\angl{\tilde{q},p}\leftarrow {\tilde{q}}_M(s)\mid \tilde{q}\in Q_M]$. If $A_N/v=q(x_0)$ then, by Condition~(2) of Lemma~\ref{lm:dtladtopprop2a}, $\varphi(q,p)=A_M(p)\Omega/v$ and consequently $q_N(s)=A_M(p)[\tilde{q}(x_0)\leftarrow \tilde{q}_M(s)\mid \tilde{q}\in Q_M]/v$. Hence $N(s)=t[\tilde{q}(x_0)\leftarrow \tilde{q}_M(s)\mid \tilde{q}\in Q_M]$ with $t=A_N\Phi[A_M(p)/v_1,\dots,A_M(p)/v_r]$ where $v_1,\dots,v_r$ are the occurrences of $\bot$ in $A_N\Phi$, from left to right. By Condition~(1) of Lemma~\ref{lm:dtladtopprop2a}, $A_N\Phi\sqsubseteq A_M(p)$ and so $t=A_M(p)$ and $N(s)=A_M(p)[\tilde{q}(x_0)\leftarrow \tilde{q}_M(s)\mid \tilde{q}\in Q_M]=M(s)$. \qed \end{proof} By this lemma, it suffices to guarantee that the constructed dtop $N$ satisfies (2)(b) of Lemma~\ref{lm:dtladtopprop2b} for all possible choices of $s_j$ (rather than some fixed choice). To do this we use the following property of the canonical dtop $\td(M)$. \begin{lemma}\label{lm:i-unique} Let $M,N$ be equivalent canonical dtlas such that $M$ is a dtpla and $N$ is a dtop, and let $\varphi$ be the aheadness mapping from $N$ to $M$. Then $N$ has the following property. \begin{enumerate} \item[(A)] For every $q,\bar{q}\in Q_N$, $a\in \Sigma_k$, $i\in[k]$ and $v\in V_\bot(\rhs_N(q,a)\Phi)$, \\ if $\rhs_N(q,a)/v=\bar{q}(x_i)$ then, for all $p_1,\dots,p_k\in P_M$, \begin{enumerate} \item[(1)] $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)/v\in \T_\Delta(Q_M(\{x_i\}))$ and \item[(2)] $\varphi(\bar{q},p_i)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Omega/v$. \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} (1) We choose trees $s_j\in\sem{p_j}_M$ for $j\in[k]$, $j\neq i$. By~(2)(b) of Lemma~\ref{lm:dtladtopprop2b}, $\varphi(\bar{q},p_i)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_{iM}/v$. Suppose that there exists $w\in\Nat_+^*$ such that $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)/vw=\tilde{q}(x_j)$ with $j\neq i$. Let $s_j'\in\sem{p_j}_M$ be such that $\tilde{q}_M(s_j')\neq\tilde{q}_M(s_j)$; it exists because $M$ is earliest (and so $\tilde{q}_M(s_j')$ and $\tilde{q}_M(s_j)$ can even have different root symbols). Then, again by~(2)(b) of Lemma~\ref{lm:dtladtopprop2b}, $\varphi(\bar{q},p_i)=\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_{iM}'/v$ where $\Psi_{iM}'$ is obtained from $\Psi_{iM}$ by changing $s_j$ into $s_j'$. But the trees $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_{iM}/v$ and $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)\Psi_{iM}'/v$ are different: their subtrees at node $w$ are $\tilde{q}_M(s_j)$ and $\tilde{q}_M(s_j')$, respectively. That is a contradiction. (2) follows immediately from~(1) and from~(2)(b) of Lemma~\ref{lm:dtladtopprop2b}. \qed \end{proof} We now change the algorithm that constructs $N=\td(M)$, as described before Theorem~\ref{th:alg} and in its proof, by replacing the equivalence test of $N$ with $M$ (at the end of the algorithm) by the test that $N$ satisfies Property~(A) of Lemma~\ref{lm:i-unique}. If that test is negative, then the construction returns the answer `no'. If it is positive, then the construction returns the dtop $N$. In fact, in that case $N$ is equivalent to $M$ by Lemma~\ref{lm:NequiM} because Property~(A) obviously implies (2)(b)~of Lemma~\ref{lm:dtladtopprop2b} for all choices of $s_j$. Let us finally note that (1) of Property~(A) cannot replace (2)(a) of Lemma~\ref{lm:dtladtopprop2b} in the construction of $N$, because it is possible that $\rhs_{M,\varphi}(q,a,p_1,\dots,p_k)/v\in \T_\Delta$ for all $p_1,\dots,p_k\in P_M$. \subsection{Examples} \label{sec:algex} In this subsection we give examples of the algorithm in the proof of Theorem~\ref{th:alg}, first several examples in which the algorithm returns the answer `no', and at the end an example in which the algorithm successfully returns a dtop equivalent to the given dtla. One might wonder whether the tests on height in the algorithm in the proof of Theorem~\ref{th:alg} are necessary, i.e., whether a difference bound for $M$ is needed at all. In the next example we show that, without the tests on height, the algorithm may not halt. In fact, in such a case it can be viewed as computing an \emph{infinite} dtop equivalent to $M$. \begin{example}\label{ex:alg_non_halt} Consider the dtla $M$ of Example~\ref{ex:Mex} with $M(\sigma^na)=a$ and $M(\sigma^nb)=\sigma^nb$ for every $n\in\Nat$. It has $P_M=\{p_a,p_b\}$ with $\delta(a)=p_a$, $\delta(b)=p_b$, $\delta(\sigma,p_a)=p_a$ and $\delta(\sigma,p_b)=p_b$. It has $Q_M=\{q\}$, its two axioms are $A_M(p_a)=a$ and $A_M(p_b)=q(x_0)$, and its rules are $q(\sigma(x_1\ot p_b))\to \sigma(q(x_1))$ and $q(b)\to b$. In Example~\ref{ex:diftup} we have seen that $\diftup(M)=\{(a,\sigma^n\angl{q,p_b})\mid n\in\Nat\}$. It is easy to see that $M$ is canonical. We now apply to $M$ the construction of $N$ in the proof of Theorem~\ref{th:alg}, without the tests on height. By Lemma~\ref{lm:dtladtopprop2a}, $A_N\Phi=a\sqcap\angl{q,p_b}=\bot$ and so $A_N=q_0(x_0)$ with $\varphi(q_0,p_a)=a$ and $\varphi(q_0,p_b)=\angl{q,p_b}$, i.e., $q_0=(a,\angl{q,p_b})$. Assume now that the algorithm has constructed the state $q_n$ with $\varphi(q_n,p_a)=a$ and $\varphi(q_n,p_b)=\sigma^n\angl{q,p_b}$, i.e., $q_n$ is the difference tuple $(a,\sigma^n\angl{q,p_b})$ of $M$. By (1) of Lemma~\ref{lm:dtladtopprop2b}, $\rhs_N(q_n,b)=\rhs_{M,\varphi}(q_n,b)= \varphi(q_n,p_b)[\angl{\bar{q},p_b}\leftarrow\rhs_M(\bar{q},b)\mid\bar{q}\in Q_M]= \varphi(q_n,p_b)[\angl{q,p_b}\leftarrow b]=\sigma^nb$. Thus, $N$ has the rule $q_n(b)\to \sigma^nb$. Similarly, $\rhs_N(q_n,a)=\rhs_{M,\varphi}(q_n,a)=\varphi(q_n,p_a)=a$ and so $N$ has the rule $q_n(a)\to a$. Next, we compute $\rhs_N(q_n,\sigma)$. To do that we need $\rhs_{M,\varphi}(q_n,\sigma,p)$ for every $p\in P_M$. For $p=p_b$ we have $\rhs_{M,\varphi}(q_n,\sigma,p_b)=\varphi(q_n,p_b)[\angl{q,p_b}\leftarrow\rhs_M(q,\sigma,p_b)]= \sigma^n\sigma q(x_1)=\sigma^{n+1}q(x_1)$, and for $p=p_a$ we have $\rhs_{M,\varphi}(q_n,\sigma,p_a)=\varphi(q_n,p_a)=a$. Thus, by (1) of Lemma~\ref{lm:dtladtopprop2b}, $\rhs_N(q_n,\sigma)\Phi=a\sqcap \sigma^{n+1}\angl{q,p_b}=\bot$. Hence, $\rhs_N(q_n,\sigma)=q(x_1)$ for some state $q$ (because $\sigma\in\Sigma_1$, see the remark after Lemma~\ref{lm:dtladtopprop2b}). By (2)(b) of Lemma~\ref{lm:dtladtopprop2b} (and because $\Psi_{1M}=\Omega$), $\varphi(q,p_y)= \rhs_{M,\varphi}(q_n,\sigma,p_y)\Omega$ for $y=a,b$, and so $\varphi(q,p_a)=a$ and $\varphi(q,p_b)=\sigma^{n+1}\angl{q,p_b}$. In other words, $q=q_{n+1}$ and $N$ has the rule $q_n(\sigma(x_1))\to q_{n+1}(x_1)$. This shows that the construction does not halt. It can be viewed as constructing the \emph{infinite} dtop $N$ with $Q_N=\{q_n\mid n\in\Nat\}=\diftup(M)$, $A_N=q_0(x_0)$ and rules $q_n(a)\to a$, $q_n(b)\to \sigma^nb$ and $q_n(\sigma(x_1))\to q_{n+1}(x_1)$ for every $n\in\Nat$. Clearly, this infinite dtop $N$ is equivalent to $M$. The dtla $M$ is ultralinear and b-erasing (in fact, linear and nonerasing). We will see in the proof of Theorem~\ref{th:ultra} that it has difference bound $h(M)=1+4\cdot\maxrhs(M)\cdot(|Q|+2)^2\cdot|P|^2=1+4\cdot 2\cdot 3^2\cdot 2^2=289$. So, with this difference bound given, the construction in the proof of Theorem~\ref{th:alg} (\emph{with} the tests on height) will halt when constructing $q_{290}$. \qed \end{example} In the next example we show that it is necessary to test the equivalence of $N$ and $M$ at the end of the algorithm in the proof of Theorem~\ref{th:alg}, or to test that $N$ has Property~(A) of Lemma~\ref{lm:i-unique}. \begin{example}\label{ex:eqtest} Let $\Sigma=\{\sigma^{(2)},a^{(0)}\}$ and $\Delta=\{\sigma^{(1)},a^{(1)},e^{(0)}\}$, and consider the following canonical dtla $M$ that translates every tree $\sigma(s_1,\sigma(s_2,\dots,\sigma(s_n,a)\cdots))$ into the tree $r_1r_2\cdots r_ne$, where $r_i\in\{a,\sigma\}$ is the root symbol of $s_i$ for $i\in[n]$. Its look-ahead automaton has two states $p_a$ and $p_\sigma$ with $\delta(a)=p_a$ and $\delta(\sigma,p_y,p_z)=p_\sigma$ for all $y,z\in\{a,\sigma\}$. It has one state $q$ with $\rho(q)=p_\sigma$, its two axioms are $A_M(p_a)=e$ and $A_M(p_\sigma)=q(x_0)$, and its four rules are \[ \begin{array}{lllllll} q(\sigma(x_1\ot p_a,x_2\ot p_\sigma) & \to & a(q(x_2)), & \hspace{0.5cm} & q(\sigma(x_1\ot p_\sigma,x_2\ot p_\sigma) & \to & \sigma(q(x_2)), \\ q(\sigma(x_1\ot p_a,x_2\ot p_a) & \to & a(e), & & q(\sigma(x_1\ot p_\sigma,x_2\ot p_a) & \to & \sigma(e). \end{array} \] We construct $N$ as in the proof of Theorem~\ref{th:alg} (assuming a large difference bound). By Lemma~\ref{lm:dtladtopprop2a}, $A_N=q_0(x_0)$ with $q_0=(e,\angl{q,p_\sigma})$ where $P_M=\{p_a,p_\sigma\}$. By (1) of Lemma~\ref{lm:dtladtopprop2b}, $\rhs_N(q_0,a)=\rhs_{M,\varphi}(q_0,a)=e$ and so $N$ has the rule $q_0(a)\to e$. To compute $\rhs_N(q_0,\sigma)$, we observe that for $y,z\in\{a,\sigma\}$, we have $\rhs_{M,\varphi}(q_0,\sigma,p_y,p_z)=\rhs_M(q,\sigma,p_y,p_z)$. Hence $\rhs_N(q_0,\sigma)\Phi=\bot$, and so $N$ may have a rule of the form $q_0(\sigma(x_1,x_2))\to q_1(x_i)$. To determine $i$ and $q_1$, we choose $s_2=a$ (for $i=1$) and $s_1=a$ (for $i=2$) in (2) of Lemma~\ref{lm:dtladtopprop2b}. By (2)(a) of Lemma~\ref{lm:dtladtopprop2b}, \begin{quote} $i=1$ if and only if $\varepsilon\in V_\bot(\rhs_M(q,\sigma,p_a,p_a)\Psi_{1M}\sqcap\rhs_M(q,\sigma,p_\sigma,p_a)\Psi_{1M})$ \end{quote} if and only if $\varepsilon\in V_\bot(a(e)\sqcap\sigma(e))=V_\bot(\bot)$, which is true, and \begin{quote} $i=2$ if and only if $\varepsilon\in V_\bot(\rhs_M(q,\sigma,p_a,p_a)\Psi_{2M}\sqcap\rhs_M(q,\sigma,p_a,p_\sigma)\Psi_{2M})$ \end{quote} if and only if $\varepsilon\in V_\bot(a(e)\sqcap a(\angl{q,p_\sigma}))=V_\bot(a(\bot))$, which is false. So, $N$ has the rule $q_0(\sigma(x_1,x_2))\to q_1(x_1)$, where $q_1=(a(e),\sigma(e))$ by (2)(b) of Lemma~\ref{lm:dtladtopprop2b}. It now follows from (1) of Lemma~\ref{lm:dtladtopprop2b} that $N$ has the rules $q_1(a)\to a(e)$ and $q_1(\sigma(x_1,x_2))\to \sigma(e)$. So, the construction ends with the dtop $N$ that has axiom $A_N=q_0(x_0)$ and the rules $q_0(a)\to e$, $\;q_0(\sigma(x_1,x_2))\to q_1(x_1)$, $\;q_1(a)\to a(e)$ and $\;q_1(\sigma(x_1,x_2))\to \sigma(e)$. Obviously, $N$ is not equivalent to $M$. And $N$ indeed does not satisfy Property~(A) of Lemma~\ref{lm:i-unique} because, although $\rhs_N(q_0,\sigma)=q_1(x_1)$, we have $\rhs_{M,\varphi}(q_0,\sigma,p_a,p_\sigma)= \rhs_M(q,\sigma,p_a,p_\sigma)=a(q(x_2))\notin\T_\Delta(\{x_1\})$. We note that, in the above construction, the choice of $s_1$ is irrelevant because $x_1$ does not occur in any right-hand side of a rule of $M$. If we choose some $s_2 \neq a$, then we obtain for $N$ the state $q_1 = (a(t_2),\sigma(t_2))$ with the rules $q_1(a) \to a(t_2)$ and $q_1(\sigma(x_1,x_2)) \to \sigma(t_2)$, where $t_2 = q_M(s_2)$. \qed \end{example} Continuing the previous example, we now show that the construction of a rule of the dtop $N$ can fail because it is impossible to determine $i$ for an occurrence of $\bar{q}(x_i)$ in its right-hand side. \begin{example} Consider the same dtla $M$ as in Example~\ref{ex:eqtest}, but change its last rule into $q(\sigma(x_1\ot p_\sigma,x_2\ot p_a)\to a(e)$; so now the last symbol $r_n$ in the output tree is always $a$. As before, $N$ may have a rule of the form $q_0(\sigma(x_1,x_2))\to q_1(x_i)$, but now $i=1$ if and only if $\varepsilon\in V_\bot(a(e)\sqcap a(e))=V_\bot(a(e))$, which is false. So, both $i=1$ and $i=2$ are false, and the construction of $N$ fails. Consider again the same dtla $M$ as in the previous example, but now change its first rule into $q(\sigma(x_1\ot p_a,x_2\ot p_\sigma) \to \sigma(q(x_2))$; so now $r_1,\dots,r_{n-1}$ in the output tree are always $\sigma$. For the possible rule $q_0(\sigma(x_1,x_2))\to q_1(x_i)$ we now obtain that $i=2$ if and only if $\varepsilon\in V_\bot(a(e)\sqcap \sigma(\angl{q,p_\sigma}))=V_\bot(\bot)$, which is true. So, both $i=1$ and $i=2$ are true, and the construction of $N$ fails. \qed \end{example} Finally, we give an example in which the construction of $N$ succeeds and returns a dtop that is equivalent to $M$. \begin{example} Consider the la-uniform version of the dtla $M$ of Example~\ref{ex:Mleaves}, as presented in Example~\ref{ex:la-uniform}. As observed after Example~\ref{ex:la-uniform}, it is easy to see that $M$ is canonical. Recall that its translation is such that $M(aa)=aa$, $M(ab)=ab$, $M(ba)=ba$, $M(bb)=bb$ and $M(\sigma(s_1,s_2))=\sigma(M(s_1),M(s_2),\#(y,z))$ where $y\in\{a,b\}$ is the first letter of the label of the left-most leaf of $\sigma(s_1,s_2)$ and $z\in\{a,b\}$ is the second letter of the label of its right-most leaf. It has $P_M=\{p_{aa},p_{ab},p_{ba},p_{bb}\}$ with $\delta(yz)=p_{yz}$ and $\delta(\sigma,p_{wx},p_{yz})=p_{wz}$ for all $w,x,y,z\in\{a,b\}$. It has $Q_M=\{q_{yz}\mid y,z\in\{a,b\}\}$, its axioms are $A_M(p_{yz})=q_{yz}(x_0)$, and its rules are $q_{yz}(yz)\to yz$ and $$q_{wz}(\sigma(x_1\ot p_{wx},x_2\ot p_{yz}))\to \sigma(q_{wx}(x_1),q_{yz}(x_2),\#(w,z))$$ for all $w,x,y,z\in\{a,b\}$. In Example~\ref{ex:diftup} we have seen that $\diftup(M)$ consists of the three 4-tuples $(a,a,b,b)$, $\;(a,b,a,b)$ and $(\angl{q_{aa},p_{aa}},\angl{q_{ab},p_{ab}},\angl{q_{ba},p_{ba}},\angl{q_{bb},p_{bb}})$. We construct $N$ as in the proof of Theorem~\ref{th:alg}; since $\maxdiff(M)=0$, the construction is the same for every difference bound $h(M)$. By Lemma~\ref{lm:dtladtopprop2a}, $A_N=q_0(x_0)$ with $\varphi(q_0,p_{yz})=\angl{q_{yz},p_{yz}}$ for $y,z\in\{a,b\}$. So, $q_0=(\angl{q_{aa},p_{aa}},\angl{q_{ab},p_{ab}},\angl{q_{ba},p_{ba}},$ $\angl{q_{bb},p_{bb}})$. From (1) of Lemma~\ref{lm:dtladtopprop2b} we obtain that $\rhs_N(q_0,yz)\Phi=\rhs_{M,\varphi}(q_0,yz)= \varphi(q_0,p_{yz})[\angl{q_{yz},p_{yz}}\leftarrow\rhs_M(q_{yz},yz)]=\rhs_M(q_{yz},yz)=yz$, and hence $N$ has the rules $q_0(yz)\to yz$ for all $y,z\in\{a,b\}$. To compute $\rhs_N(q_0,\sigma)$, we first observe that for every $w,x,y,z\in\{a,b\}$, \begin{eqnarray*} \lefteqn{\rhs_{M,\varphi}(q_0,\sigma,p_{wx},p_{yz})} \\ & \hspace{0.5cm} = & \varphi(q_0,p_{wz})[\angl{q_{wz},p_{wz}}\leftarrow\rhs_M(q_{wz},\sigma,p_{wx},p_{yz})] \\ & \hspace{0.5cm} = & \rhs_M(q_{wz},\sigma,p_{wx},p_{yz}) \\ & \hspace{0.5cm} = & \sigma(q_{wx}(x_1),q_{yz}(x_2),\#(w,z)). \end{eqnarray*} So, by (1) of Lemma~\ref{lm:dtladtopprop2b}, $\rhs_N(q_0,\sigma)\Phi=\sigma(\bot,\bot,\#(\bot,\bot))$. Thus, $N$ may have a rule of the form $$q_0(\sigma(x_1,x_2))\to \sigma(q_3(x_{i_3}),q_4(x_{i_4}),\#(q_1(x_{i_1}),q_2(x_{i_2}))).$$ Let $s_1=s_2=aa$. From (2)(a) of Lemma~\ref{lm:dtladtopprop2b} we obtain for $v=(3,1)$ that \[ \begin{array}{lll} i_1=1 & \Longleftrightarrow & v\in V_\bot(\sqcap\{\rhs_{M,\varphi}(q_0,\sigma,p_{wx},p_{aa})\Psi_{1M}\mid w,x\in\{a,b\}\}) \\ & \Longleftrightarrow & v\in V_\bot(\sqcap\{\sigma(\angl{q_{wx},p_{wx}},aa,\#(w,a))\mid w,x\in\{a,b\}\}) \end{array} \] if and only if $v\in V_\bot(\sigma(\bot,aa,\#(\bot,a)))$, which is true, and \[ \begin{array}{lll} i_1=2 & \Longleftrightarrow & v\in V_\bot(\sqcap\{\rhs_{M,\varphi}(q_0,\sigma,p_{aa},p_{yz})\Psi_{2M}\mid y,z\in\{a,b\}\}) \\ & \Longleftrightarrow & v\in V_\bot(\sqcap\{\sigma(aa,\angl{q_{yz},p_{yz}},\#(a,z))\mid y,z\in\{a,b\}\}) \end{array} \] if and only if $v\in V_\bot(\sigma(aa,\bot,\#(a,\bot)))$, which is false. So $i_1=1$ and $\varphi(q_1,p_{wx})=w$ for all $w,x\in\{a,b\}$ by (2)(b) of Lemma~\ref{lm:dtladtopprop2b}. Thus, $q_1=(a,a,b,b)$. Similarly we obtain for $v=(3,2)$ that $i_2=2$ and $\varphi(q_2,p_{yz})=z$, for $v=1$ that $i_3=1$ and $\varphi(q_3,p_{wx})=\angl{q_{wx},p_{wx}}$, and for $v=2$ that $i_4=2$ and $\varphi(q_4,p_{yz})=\angl{q_{yz},p_{yz}}$. Hence $q_2=(a,b,a,b)$, $\;q_3=q_4=q_0$ and $N$ has the rule $$q_0(\sigma(x_1,x_2))\to \sigma(q_0(x_1),q_0(x_2),\#(q_1(x_1),q_2(x_2))).$$ Next we consider $q_2$. Clearly, both $\rhs_{M,\varphi}(q_2,yz)$ and $\rhs_{M,\varphi}(q_2,\sigma,p_{wx},p_{yz})$ equal $z$. Thus, $N$ has the rules $q_2(yz)\to z$ and it may have a rule of the form $q_2(\sigma(x_1,x_2))\to q(x_i)$. Taking again $s_1=s_2=aa$, we get that \[ \begin{array}{lll} i=1 & \Longleftrightarrow & \varepsilon\in V_\bot(\sqcap\{\rhs_{M,\varphi}(q_2,\sigma,p_{wx},p_{aa})\mid w,x\in\{a,b\}\}) \end{array} \] if and only if $\varepsilon\in V_\bot(a)$, which is false, and \[ \begin{array}{lll} i=2 & \Longleftrightarrow & \varepsilon\in V_\bot(\sqcap\{\rhs_{M,\varphi}(q_2,\sigma,p_{aa},p_{yz})\mid y,z\in\{a,b\}\}) \end{array} \] if and only if $\varepsilon\in V_\bot(a\sqcap b)$, which is true. So $i=2$ and $\varphi(q,p_{yz})=z$, which means that $q=q_2$. Hence, $N$ has the rule $q_2(\sigma(x_1,x_2))\to q_2(x_2)$. Similarly it has the rules $q_1(yz)\to y$ and $q_1(\sigma(x_1,x_2))\to q_1(x_1)$. Thus, the construction ends with the dtop $N$ that has axiom $A_N=q_0(x_0)$ and the rules \[ \begin{array}{lll} q_0(yz)\to yz, & \hspace{0.5cm} & q_0(\sigma(x_1,x_2))\to \sigma(q_0(x_1),q_0(x_2),\#(q_1(x_1),q_2(x_2))), \\ q_1(yz)\to y, & & q_1(\sigma(x_1,x_2))\to q_1(x_1), \\ q_2(yz)\to z, & & q_2(\sigma(x_1,x_2))\to q_2(x_2). \end{array} \] It should be clear that $N$ is equivalent to $M$ (cf. the end of Example~\ref{ex:Mleaves}). It is also straightforward to check that $N$ has Property~(A) of Lemma~\ref{lm:i-unique}. We finally note that the original dtla $M$ of Example~\ref{ex:Mleaves} is ultralinear and b-erasing (even linear and nonerasing). The difference bound in the proof of Theorem~\ref{th:ultra} for this $M$ is $1+4\cdot 2\cdot 3^2\cdot 4^2=1153$. By Lemma~\ref{lm:make_uniform}, the la-uniform version of $M$ has the same difference bound. Obviously, this is a rather large bound, in view of the fact that $\maxdiff(M)=0$. \qed \end{example} It is left to the reader to use the construction in the proof of Theorem~\ref{th:alg} to obtain a dtop that is equivalent to the dtla of Example~\ref{ex:Mthree}. \section{A Dependency Graph for Output Branches} \label{sec:graphs} \noindent {\bf Convention.} In this section we assume additionally that the initalized la-uniform dtla $M$ is {\bf ultralinear} with mapping $\mu: Q\to \Nat$ and {\bf b-erasing} with graph $E_M$, see Section~\ref{sec:dtop}. These properties will be used (only) in the proof of Lemma~\ref{lm:cycle}. For this $M$ we will compute an ancestral bound $h_{\rm a}(M)$, as defined in Definition~\ref{df:ancestral_bound}. In view of this definition (and Lemma~\ref{lm:diff_char}), it is technically convenient to combine a node $v$ of an output tree with the sequence of labels of the proper ancestors of $v$, as follows. \medskip For the output alphabet $\Delta$, we define the \emph{branch alphabet} $\Delta_\B$ by: \[ \Delta_\B = \{(d,j)\mid d\in\Delta,\,\rk(d)\geq 1,\,j\in[\rk(d)]\}. \] A string in $\Delta_\B^*$ is called a \emph{branch}. For a branch $v\in\Delta_\B^*$ we define $\nod(v)\in\Nat_+^*$ to be the sequence of numbers obtained from $v$ by changing every $(d,j)$ into $j$. For a tree $t=d(t_1,\ldots,t_k)\in T_\Delta(Q\times P)$ with $k\in\Nat$, $d\in(\Delta\cup (Q\times P))^{(k)}$, and $t_1,\ldots,t_k\in T_\Delta(Q\times P)$, we define the set $B(t)\subseteq \Delta_\B^*$ of \emph{branches of $t$} inductively as follows: \[ B(d(t_1,\ldots,t_k))=\{\varepsilon\}\cup\{(d,j)\,v\mid j\in[k], v\in B(t_j)\}. \] The mapping $\nod$, restricted to $B(t)$, is a bijection from $B(t)$ to $V(t)$. Intuitively, a branch $v$ contains the node $\nod(v)$ and the labels of its proper ancestors (from the root to the node). For example, if $v=(a,2)(b,1)(b,3)$ is a branch of $t$, then it corresponds to the node $\nod(v)=(2,1,3)$ of $t$, and moreover, the root of $t$ has label $a$, node $2$ has label $b$, and node $(2,1)$ has label $b$ too. For a branch $v\in B(t)$, we define $t/v=t/\nod(v)$, $\lab(t,v)=\lab(t,\nod(v))$, $\lk_s(v)=\lk_s(\nod(v))$, and $\orig_s(v)=\orig_s(\nod(v))$. We will need the following lemma on branches. Roughly, it says that if $M(C[p])$ has a branch that is longer than any branch of $M(C[p'])$, then a prefix of that branch corresponds to a difference node of $M(C[p])$ and $M(C[p'])$. In this section, we rename $p,p'$ into $p_1,p_2$. \begin{lemma}\label{lm:height-diff} Let $C\in\C_\Sigma$ and $p_1,p_2\in P$. Let $v$ be a branch of both $M(C[p_1])$ and $M(C[p_2])$, and let $w\in\Delta_\B^*$ be such that \begin{enumerate} \item[(1)] $vw$ is a branch of $M(C[p_1])$, and \item[(2)] $|w|>\height(M(C[p_2])/v)$. \end{enumerate} Then there is a prefix $w'$ of $w$ such that $\nod(vw')$ is a difference node of $M(C[p_1])$ and $M(C[p_2])$. \end{lemma} \begin{proof} Since $|w|>\height(M(C[p_2])/v)$, $vw$ is not a branch of $M(C[p_2])$. Let $w'$ be the longest prefix of $w$ such that $vw'$ is a branch of $M(C[p_2])$. Since $vw'$ is also a branch of $M(C[p_1])$, all proper ancestors of $\nod(vw')$ have the same label in $M(C[p_1])$ and $M(C[p_2])$. By Lemma~\ref{lm:diff_char}, it remains to show that $\nod(vw')$ has different labels in $M(C[p_1])$ and $M(C[p_2])$. Since $w'\neq w$, there exist $k\geq 1$, $d\in\Delta^{(k)}$, and $j\in[k]$ such that $w'(d,j)$ is a prefix of $w$. Then $d=\lab(M(C[p_1]),vw')$. Suppose that also $d=\lab(M(C[p_2]),vw')$. Then $vw'(d,j)$ is a branch of $M(C[p_2])$, contradicting the choice of $w'$. \qed \end{proof} Again in view of Definition~\ref{df:ancestral_bound}, we extend the definition of $\lk$ to branches, and call it $\blk$; we are, however, only interested in triples $(q,u,\#)$. For $s\in\T_\Sigma(P)$ and every $v\in\Delta_\B^*$, the sets $\blk_s(v)$ of triples $(q,u,\#)$ with $q\in Q$ and $u\in V(s)$, are defined inductively as follows: \begin{enumerate} \item[(1)] If $\delta(s)=p$, then $(q_{0,p},\varepsilon,\#)\in\blk_s(\varepsilon)$. \item[(2)] If $(q,u,\#)\in\blk_s(v)$ and $\rhs(q,s,u)=\zeta$, \\ then $(\bar{q},ui,\#)\in\blk_s(vz)$ for every $z\in B(\zeta)$ with $\zeta/z=\bar{q}(x_i)$. \end{enumerate} \begin{lemma}\label{lm:blink} The following two statements are equivalent: \begin{enumerate} \item[(1)] $(q,u,\#)\in\blk_s(v)$; \item[(2)] $(q,u,\#)\in\lk_s(v)$ and $v\in B(M(s))$. \end{enumerate} \end{lemma} \begin{proof} Since the definition of $\blk_s$ closely follows the semantics of $M^\circ$, it is straightforward to show the following statement, similar to Lemma~\ref{lm:linkdef}(1): $(q,u,\#)\in\blk_s(v)$ if and only if there is a sentential form $\xi$ of $M^\circ$ such that $q_{0,\delta(s)}(\varepsilon)\Rightarrow_s^*\xi$, $\;v\in B(\xi)$ and $\xi/v=q(u)$. This proves the equivalence, by Lemmas~\ref{lm:linkdef}(1) and~\ref{lm:elem}(2). \qed \end{proof} With these definitions, we can reformulate Definition~\ref{df:ancestral_bound} as follows. \begin{lemma}\label{lm: ancbound_redef1} A number $h_{\rm a}(M)\in\Nat$ is an ancestral bound for $M$ if and only if either $\diff(M)$ is infinite or the following holds for every $C\in\C_\Sigma$, $p_1,p_2\in P$, $u\in V(C)$, $v_1\in B(M(C[p_1]))$, $v_2\in B(M(C[p_2]))$, and $q_1,q_2\in Q$: \\ if $(q_1,u,\#)\in\blk_{C[p_1]}(v_1)$, $\;(q_2,u,\#)\in\blk_{C[p_2]}(v_2)$, and $v_2$ is a prefix of $v_1$, \\ then $|v_1|-|v_2|\leq h_{\rm a}(M)$. \end{lemma} Thus, to determine an ancestral bound for $M$, we are interested in a node $u$ of a $\Sigma$-context $C$ and in two computations of $M$ on the path from the root to $u$, one with input $C[p_1]$ and the other with input $C[p_2]$. The idea is, roughly, to find two ancestors $\bar{u}$ and $\tilde{u}$ of $u$ in $C$ where each of these computations is in a cycle (i.e., arrives in the same state at $\bar{u}$ and $\tilde{u}$), and then to pump the part of $C$ between $\bar{u}$ and $\tilde{u}$, leading to an infinite $\diff(M)$. For this it is technically helpful to build a graph $G_M$ of which each path captures two such computations: $(q_1,u,\#)\in\blk_{C[p_1]}(v_1)$ and $(q_2,u,\#)\in\blk_{C[p_2]}(v_2)$, cf. Lemma~\ref{lm:path_in_graph}. The pumping of $C$ then corresponds to the repetition of a cycle in the graph $G_M$. In this graph, we have to distinguish between the case that $u$ is on or off the spine of $C$, where the spine of a context $C$ is the path from the root to the unique occurrence of $\bot$. Formally, for $C\in\C_\Sigma$, the \emph{spine} of $C$ is the set ${\sf spi}(C)=\{u\in V(C)\mid C/u\in\C_\Sigma\}$. \medskip For the dtla $M$, we define a finite directed edge-labeled \emph{dependency graph} $G_M$ whose nodes are all 3-tuples $(q_1,q_2,b)$ with $q_1,q_2\in Q$ and $b\in\{0,1\}$, and whose edge labels are in $R\times R\times \Nat_+\times\Delta_\B^*\times\Delta_\B^*$, more precisely, in the finite set of all 5-tuples $(r_1,r_2,j,z_1,z_2)$ such that $r_1,r_2\in R$, $j\in[\rk(a)]$ for some $a\in\Sigma$, and $z_i\in \Delta_\B^*$ is a branch of the right-hand side of $r_i$ for $i=1,2$. The boolean $b$ is called the \emph{type} of the node; intuitively, $b=1$ stands for ``on the spine'', and $b=0$ for ``off the spine''. The edges of the dependency graph $G_M$ are defined as follows, together with its ``entry nodes''. \begin{enumerate} \item[(Ge)] For every $p_1,p_2\in P$, $(q_{0,p_1},q_{0,p_2},1)$ is an \emph{entry node} of $G_M$. \end{enumerate} \begin{enumerate} \item[(G1)] Let $k\geq 1$, $a\in\Sigma^{(k)}$, and $\ell,j\in[k]$. Let $(q_1,q_2,1)$ be a node of $G_M$ and, for $i=1,2$, let $r_i=q_i(a(x_1\ot p_{i,1},\ldots,x_k\ot p_{i,k}))\to \zeta_i$ be a rule of $M$, let $z_i\in\Delta_\B^*$ and let $q_i'\in Q$, such that \begin{enumerate} \item $p_{1,m}=p_{2,m}$ for all $m\in[k]-\{\ell\}$, \item $z_i\in B(\zeta_i)$ and $\zeta_i/z_i=q'_i(x_j)$, for $i=1,2$. \end{enumerate} Then there is an edge labeled $(r_1,r_2,j,z_1,z_2)$ from $(q_1,q_2,1)$ to $(q_1',q_2',b')$, where $b'=1$ if and only if $\ell=j$. \end{enumerate} \begin{enumerate} \item[(G0)] Let $k\geq 1$, $a\in\Sigma^{(k)}$, and $j\in[k]$. Let $(q_1,q_2,0)$ be a node of $G_M$ and, for $i=1,2$, let $r_i=q_i(a(x_1\ot p_{i,1},\ldots,x_k\ot p_{i,k}))\to \zeta_i$ be a rule of $M$, let $z_i\in\Delta_\B^*$ and let $q_i'\in Q$, such that \begin{enumerate} \item $p_{1,m}=p_{2,m}$ for all $m\in[k]$, \item $z_i\in B(\zeta_i)$ and $\zeta_i/z_i=q'_i(x_j)$, for $i=1,2$. \end{enumerate} Then there is an edge labeled $(r_1,r_2,j,z_1,z_2)$ from $(q_1,q_2,0)$ to $(q_1',q_2',0)$. \end{enumerate} \noindent Note that there are no edges from a node of type $0$ to a node of type $1$. Note also that if $(q_1,q_2,b)$ is a node of $G_M$, then so is $(q_2,q_1,b)$; moreover, if there is an edge from $(q_1,q_2,b)$ to $(q'_1,q'_2,b')$ with label $(r_1,r_2,j,z_1,z_2)$, then there is an edge from $(q_2,q_1,b)$ to $(q'_2,q'_1,b')$ with label $(r_2,r_1,j,z_2,z_1)$. A path $\pi$ in $G_M$ is a sequence $e_1\cdots e_n$ of edges, $n\geq 0$, such that for every $m\in[n-1]$ the edge $e_{m+1}$ starts at the node where $e_m$ ends. It has a \emph{label} $(u,v_1,v_2)\in\Nat_+^*\times\Delta_\B^*\times\Delta_\B^*$ obtained by the component-wise concatenation of the last three components of the labels of $e_1,\dots,e_n$. The \emph{output label} of $\pi$ is the pair of branches $(v_1,v_2)$, denoted by $\out(\pi)$. We say that $\pi$ is an \emph{entry path} if it starts at an entry node, and that it is a \emph{$(q_1,q_2,b)$-path} if it is an entry path that ends at the node $(q_1,q_2,b)$. We will only be interested in the entry paths of $G_M$. \begin{example}\label{ex:depgraph} Let $\Sigma=\{\tau^{(2)},\sigma^{(1)},a^{(0)},b^{(0)}\}$ and $\Delta=\{\tau^{(2)},\sigma_a^{(1)},\sigma_b^{(1)},a^{(0)},b^{(0)},e^{(0)}\}$. We consider an initialized la-uniform dtla $M$ such that for all $m,n\in\Nat$ and $y\in\{a,b\}$, $M(\tau(\sigma^m y,\sigma^n a))=\tau(\sigma_y^my,M_y(\sigma^n a))$ where $M_y(a)=a$, $$M_a(\sigma^{n+1} a)=\tau(M_a(\sigma^n a),a) \mbox{\;, and \;} M_b(\sigma^{n+1} a)=\tau(\tau(a,M_b(\sigma^n a)),a).$$ We will not be interested in other input trees, which are translated by $M$ into trees with at least one occurrence of $e$. The look-ahead states of $M$ are $p_a$ and $p_b$, which compute the label of the left-most leaf: $\delta(y)=p_y$, $\delta(\sigma,p_y)=p_y$, and $\delta(\tau,p_y,p_z)=p_y$ for all $y,z\in\{a,b\}$. The states of $M$ are all $q_{iy}$ with $i\in\{0,1,2\}$ and $y\in\{a,b\}$; the la-map of $M$ is $\rho(q_{0y})=\rho(q_{1y})=p_y$ and $\rho(q_{2y})=p_a$. For $y\in\{a,b\}$, the axioms of $M$ are $A(p_y)=q_{0y}(x_0)$, and it has the following rules, where the missing rules all have right-hand side $e$. First, it has the rules \[ \begin{array}{lll} r_{0y} & = & q_{0y}(\tau(x_1\ot p_y,x_2\ot p_a)) \to \tau(q_{1y}(x_1),q_{2y}(x_2)), \\ r_{1y} & = & q_{1y}(\sigma(x_1\ot p_y)) \to \sigma_y(q_{1y}(x_1)), \\ r_{2a} & = & q_{2a}(\sigma(x_1\ot p_a)) \to \tau(q_{2a}(x_1),a) \mbox{, and} \\ r_{2b} & = & q_{2b}(\sigma(x_1\ot p_a)) \to \tau(\tau(a,q_{2b}(x_1)),a). \end{array} \] Second, it has the rules $q_{1y}(y)\to y$ and $q_{2y}(a)\to a$. The dependency graph $G_M$ has the four entry nodes $(q_{0y},q_{0z},1)$ for $y,z\in\{a,b\}$. We will not construct all edges of $G_M$, but just mention four interesting ones. First, by~(G1) with $\ell=j=1$, there is an edge $e_{01}$ from $(q_{0a},q_{0b},1)$ to $(q_{1a},q_{1b},1)$ with label $(r_{0a},r_{0b},1,(\tau,1),(\tau,1))$. Second, by (G1) with $\ell=1$ and $j=2$, there is an edge $e_{02}$ from $(q_{0a},q_{0b},1)$ to $(q_{2a},q_{2b},0)$ with label $(r_{0a},r_{0b},2,(\tau,2),(\tau,2))$. Third, by (G1), there is an edge $e_1$ from $(q_{1a},q_{1b},1)$ to itself with label $(r_{1a},r_{1b},1,(\sigma_a,1),(\sigma_b,1))$. And finally, by~(G0), there is an edge $e_2$ from $(q_{2a},q_{2b},0)$ to itself with the label $(r_{2a},r_{2b},1,(\tau,1),(\tau,1)(\tau,2))$. \qed \end{example} In Lemma~\ref{lm:path_in_graph} we express the meaning of the entry paths in $G_M$. But we will need a stronger version of the implication $(1)\Rightarrow(2)$ of that lemma, which we prove now. Recall from the paragraph before Theorem~\ref{th:uniform_earliest} that $\fix(M)$ is a fixed set of representatives of the equivalence classes $\sem{p}$, $p\in P$. \begin{lemma}\label{lm:path_to_context} Let $\pi=e_1\cdots e_n$, $n\geq 0$, be an entry path in $G_M$, where $e_m$ is an edge of $G_M$ for every $m\in[n]$. For every $m$, $0\leq m\leq n$, let the prefix $e_1\cdots e_m$ of $\pi$ be a $(q_1^{(m)},q_2^{(m)},b^{(m)})$-path with label $(u^{(m)},v_1^{(m)},v_2^{(m)})$. Then there are $C\in\C_\Sigma$ and $p_1,p_2\in P$ such that for every $m$, $0\leq m\leq n$, and $i=1,2$, \begin{enumerate} \item[$(a)$] $(q_i^{(m)},u^{(m)},\#)\in\blk_{C[p_i]}(v_i^{(m)})$, \item[$(b)$] $b^{(m)}=1$ if and only if $u^{(m)}\in{\sf spi}(C)$, \item[$(c)$] if $m\leq n-1$ and $e_{m+1}$ has label $(r_1,r_2,j,z_1,z_2)$, then \begin{enumerate} \item[$(1)$] $\rhs(q_i^{(m)},C[p_i],u^{(m)})=\rhs(r_i)$ and \item[$(2)$] for every $j'\in\Nat_+$ such that $u^{(m)}j'$ is a node of $C$, if $j'\neq j$ or $m=n-1$, then $C/u^{(m)}j'\in\fix(M)\cup\{\bot\}$. \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} We proceed by induction on the length $n$ of $\pi$. If $n=0$, i.e., $\pi$ is empty, then $m=0$, $u^{(0)}=v_1^{(0)}=v_2^{(0)}=\varepsilon$, and $(q_1^{(0)},q_2^{(0)},b^{(0)})=(q_{0,p_1},q_{0,p_2},1)$ for $p_1,p_2\in P$. Let $C=\bot$. It is easy to check requirements (a) and (b) for these $C$, $p_1$ and $p_2$; requirement (c) holds trivially. Now let $\pi' = \pi e$ where $\pi$ is a $(q_1,q_2,b)$-path with label $(u,v_1,v_2)$, and where the last edge $e$ has label $(r_1,r_2,j,z_1,z_2)$ and ends at node $(q_1',q_2',b')$. Then $\pi'$ is a $(q_1',q_2',b')$-path with label $(u',v_1',v_2')$ where $u'=uj$ and $v_i'=v_iz_i$. Let $r_i$ be the rule $q_i(a(x_1\ot \bar{p}_{i,1},\ldots,x_k\ot \bar{p}_{i,k}))\to \zeta_i$ of $M$, hence $k\geq 1$, $a\in\Sigma^{(k)}$, $j\in[k]$, and $z_i\in B(\zeta_i)$ with $\zeta_i/z_i = q_i'(x_j)$ for $i=1,2$. By induction, requirements~(a), (b) and~(c) hold for some $C\in\C_\Sigma$ and $p_1,p_2\in P$, for every prefix of $\pi$. In particular (for $m=n$), $(q_i,u,\#)\in\blk_{C[p_i]}(v_i)$, and $b=1$ if and only if $u\in{\sf spi}(C)$. We first consider the case where $b=1$, i.e., the last edge $e$ is obtained by (G1). Then there exists $\ell\in[k]$ such that $\bar{p}_{1,m}=\bar{p}_{2,m}$ for all $m\in[k]-\{\ell\}$, and $b'=1$ if and only if $\ell=j$. Let $p_1'=\bar{p}_{1,\ell}$ and $p_2'=\bar{p}_{2,\ell}$. Let $s_\ell = \bot$ and for $m\in[k]-\{\ell\}$, let $s_m$ be the unique tree in $\fix(M)$ such that $\delta(s_m)=\bar{p}_{1,m}=\bar{p}_{2,m}$. Define $C'=C[u\leftarrow C_a]$ where $C_a$ is the $\Sigma$-context $a(s_1,\dots,s_k)$; since $u\in{\sf spi}(C)$, $C'$ is a $\Sigma$-context. We now claim that requirements~(a), (b), and~(c) hold for $C'$ and $p_1',p_2'$, for every prefix of $\pi'$. Obviously, $u'=uj$ is a branch of $C'$, and $C'/u\ell=\bot$. Hence, $u'\in{\sf spi}(C')$ if and only if $\ell=j$ if and only if $b'=1$. For every prefix $e_1\cdots e_m$ of the path $\pi$, by the definition of the label of a path, $u^{(m)}$ is a prefix of $u$, i.e., an ancestor of $u$. Hence, $u^{(m)}\in{\sf spi}(C')$ if and only if $u^{(m)}\in{\sf spi}(C)$ if and only if $b^{(m)}=1$ (and in fact, since $b=1$, they are all true). This proves requirement~(b). Since $(q_i,u,\#)\in\blk_{C[p_i]}(v_i)$, Lemmas~\ref{lm:blink} and~\ref{lm:linkdef}(4) imply that $\delta(C[p_i]/u)=\rho(q_i)$ which equals $\delta(C_a[p_i'])$ by definition of $C_a$ (and because $M$ is la-uniform). Hence $\delta(C[p_i]/u)=\delta(C'[p_i']/u)$, and so $\delta(C[p_i])=\delta(C'[p_i'])$ and every proper ancestor of $u$ is similar in $C[p_i]$ and $C'[p_i']$. It now follows from Lemmas~\ref{lm:link_preserve}(1) and~\ref{lm:blink} that $(q_i,u,\#)\in\blk_{C'[p_i']}(v_i)$. Since $\rhs(q_i,C'[p_i'],u)=\rhs(q_i,C_a[p_i'],\varepsilon)=\zeta_i$, the definition of $\blk$ implies that $(q_i',uj,\#)\in\blk_{C'[p_i']}(v_iz_i)$, i.e., that $(q_i',u',\#)\in\blk_{C'[p_i']}(v_i')$. Since $u^{(m)}$ is an ancestor of $u$ for every prefix $e_1\cdots e_m$ of $\pi$, Lemmas~\ref{lm:link_preserve}(1) and~\ref{lm:blink} also imply that $(q_i^{(m)},u^{(m)},\#)\in\blk_{C'[p_i']}(v_i^{(m)})$. This proves requirement~(a). It should be clear from the observations in the previous two paragraphs that requirement~(c) holds for every proper prefix $e_1\cdots e_m$ of $\pi$. In fact, $u^{(m)}$ is a proper ancestor of $u$, and $\rhs(q,C[p_i],\hat{u})=\rhs(q,C'[p_i'],\hat{u})$ for every proper ancestor $\hat{u}$ of $u$ (and every $q$), because $\hat{u}$ is similar in $C[p_i]$ and $C'[p_i']$. Moreover, if $u^{(m)}j'\neq u^{(m+1)}$ then $C/u^{(m)}j'=C'/u^{(m)}j'$. For the proper prefix $\pi$ of $\pi'$ and the edge $e$, requirement (c) holds by definition of the context $C'$. The proof for the case where $e$ is obtained by (G0) is similar, with $b=b'=0$. Since $u$ is not on the spine of $C$, we let $s_m\in \fix(M)$ with $\delta(s_m)=\bar{p}_{1,m}=\bar{p}_{2,m}$ for every $m\in[k]$, and we take $p_1'=p_1$ and $p_2'=p_2$. \qed \end{proof} \begin{example}\label{ex:path_to_context} For $n\in\Nat_+$, consider the entry path $\pi=e_{01}e_1^{n-1}$ in $G_M$, where $M$ is the dtla of Example~\ref{ex:depgraph}. It is a path from $(q_{0a},q_{0b},1)$ to $(q_{1a},q_{1b},1)$, and it has label $(1^n,(\tau,1)(\sigma_a,1)^{n-1},(\tau,1)(\sigma_b,1)^{n-1})$ where $1^n$ is the sequence $(1,\dots,1)$ of length~$n$. Let $\fix(M)=\{a,b\}$. Then the requirements of Lemma~\ref{lm:path_to_context} are fulfilled for $C=\tau(\sigma^{n-1}\bot,a)$ and for $p_1=p_a$ and $p_2=p_b$. Now consider the entry path $\pi=e_{02}e_2^{n-1}$ from $(q_{0a},q_{0b},1)$ to $(q_{2a},q_{2b},0)$ with label $(2\cdot 1^{n-1},(\tau,2)(\tau,1)^{n-1},(\tau,2)((\tau,1)(\tau,2))^{n-1})$. The requirements of Lemma~\ref{lm:path_to_context} for this path are fulfilled for $C=\tau(\bot,\sigma^{n-1}a)$ and for $p_1=p_a$ and $p_2=p_b$. \qed \end{example} In the next lemma we express the meaning of the entry paths in $G_M$. \begin{lemma}\label{lm:path_in_graph} Let $q_1,q_2\in Q$, $b\in\{0,1\}$, $u\in\Nat_+^*$, and $v_1,v_2\in\Delta_\B^*$. \\ Then the following two statements are equivalent: \begin{enumerate} \item[$(1)$] there is a $(q_1,q_2,b)$-path in $G_M$ with label $(u,v_1,v_2)$; \item[$(2)$] there are $C\in\C_\Sigma$ and $p_1,p_2\in P$ such that \begin{enumerate} \item $(q_1,u,\#)\in\blk_{C[p_1]}(v_1)$ and $(q_2,u,\#)\in\blk_{C[p_2]}(v_2)$, and \item $b=1$ if and only if $u\in{\sf spi}(C)$. \end{enumerate} \end{enumerate} \end{lemma} \begin{proof} $(1)\Rightarrow(2)$.\quad Immediate from Lemma~\ref{lm:path_to_context} for $m=n$ (and disregarding (c)). $(2)\Rightarrow(1)$.\quad This proof is similar to the proof of Lemma~\ref{lm:path_to_context}, but it is easier because in the induction step we can use the same $C$, $p_1$, and $p_2$. Here we proceed by induction on the length of $u$. Let $u=\varepsilon$. Then $u\in{\sf spi}(C)$ and so $b=1$. Since $(q_i,\varepsilon,\#)\in\blk_{C[p_i]}(v_i)$, it follows from the definition of $\blk$ that $v_i=\varepsilon$, and that $q_i=q_{0,\bar{p}_i}$ where $\bar{p}_i=\delta(C[p_i])$. Hence $(q_1,q_2,b)=(q_{0,\bar{p}_1},q_{0,\bar{p}_2},1)$ is an entry node. This proves statement~(1) for this case, because the empty path from $(q_1,q_2,b)$ to itself has label $(\varepsilon,\varepsilon,\varepsilon)$. Now let $u'=uj$ with $u\in\Nat_+^*$ and $j\in\Nat_+$. Let $C\in\C_\Sigma$ and $p_1,p_2\in P$ (and consider $q_i'$, $v_i'$, and $b'$) such that $(q_i',u',\#)\in\blk_{C[p_i]}(v_i')$ for $i=1,2$ and such that $b'=1$ if and only if $u'\in{\sf spi}(C)$. By the definition of $\blk$, there exists a rule $r_i$ of the form $q_i(a(x_1\ot \bar{p}_{i,1},\ldots,x_k\ot \bar{p}_{i,k}))\to \zeta_i$ such that $a=\lab(C,u)$ with $k\geq 1$, $a\in\Sigma^{(k)}$, and $j\in[k]$, and such that $\bar{p}_{i,m}=\delta(C[p_i]/um)$ for every $m\in[k]$; moreover, there exist $v_i\in\Delta_\B^*$ and $z_i\in B(\zeta_i)$ such that $v_i'=v_iz_i$, $\;(q_i,u,\#)\in\blk_{C[p_i]}(v_i)$, and $\zeta_i/z_i=q_i'(x_j)$. We first consider the case where $u\in{\sf spi}(C)$. Then, by induction, there is a $(q_1,q_2,1)$-path $\pi$ in $G_M$ with label $(u,v_1,v_2)$. Let $\ell\in[k]$ be such that $u\ell\in{\sf spi}(C)$; so, $b'=1$ if and only if $\ell=j$. We observe that $\bar{p}_{1,m}=\bar{p}_{2,m}$ for all $m\in[k]-\{\ell\}$, because $C/um\in\T_\Sigma$. So, by (G1), there is an edge $e$ labeled $(r_1,r_2,j,z_1,z_2)$ from $(q_1,q_2,1)$ to $(q_1',q_2',b')$. Hence, $\pi e$ is a $(q_1',q_2',b')$-path with label $(uj,v_1z_1,v_2z_2)=(u',v_1',v_2')$. The proof for the case where $u\notin{\sf spi}(C)$ is similar. Then, by induction, there is a $(q_1,q_2,0)$-path $\pi$. Of course $u'\notin{\sf spi}(C)$, and so $b'=0$. Since now $\bar{p}_{1,m}=\bar{p}_{2,m}$ for all $m\in[k]$, there is an edge $e$ from $(q_1,q_2,0)$ to $(q_1',q_2',b')$ by~(G0). \qed \end{proof} For strings $v_1,v_2$, we define $\diff(v_1,v_2)$ to be the pair of strings $(w_1,w_2)$ such that $v_1=vw_1$ and $v_2=vw_2$ where $v$ is the longest common prefix of $v_1$ and $v_2$. Note that if $\diff(v_1,v_2)=(w_1,w_2)$, then $\diff(v_1z_1,v_2z_2)=\diff(w_1z_1,w_2z_2)$ for all strings $z_1$ and $z_2$. For a path $\pi$ in $G_M$, we define $\diff(\pi)=\diff(\out(\pi))$. So, if $\pi=\pi_1\pi_2$ then $\diff(\pi)=\diff(\diff(\pi_1)\cdot\out(\pi_2))$, where $\cdot$ is component-wise concatenation. We say that a pair of strings $(w_1,w_2)\in\Delta_\B^*\times\Delta_\B^*$ is \emph{ancestral} if $w_1=\varepsilon$ or $w_2=\varepsilon$ (or both). A path $\pi$ in $G_M$ is \emph{ancestral} if $\diff(\pi)$ is ancestral. Thus, if $\out(\pi)=(v_1,v_2)$, then $\pi$ is ancestral if and only if $v_1$ is a prefix of $v_2$ or vice versa. Clearly, if $\pi$ is ancestral then every prefix of $\pi$ is ancestral: if $\pi=\pi_1\pi_2$ and $\diff(\pi_1)$ is not ancestral (i.e., both its components are nonempty and their first symbols differ), then $\diff(\pi_1)\cdot\out(\pi_2)$ is not ancestral and $\diff(\pi)=\diff(\pi_1)\cdot\out(\pi_2)$. By Lemmas~\ref{lm: ancbound_redef1} and~\ref{lm:path_in_graph}, and using the above definitions, we can again reformulate the definition of ancestral bound (Definition~\ref{df:ancestral_bound}), as follows. \begin{lemma}\label{lm: ancbound_redef2} A number $h_{\rm a}(M)\in\Nat$ is an ancestral bound for $M$ if and only if either $\diff(M)$ is infinite or the following holds for every ancestral entry path $\pi$ in $G_M$: \\ if $\diff(\pi)=(w,\varepsilon)$, then $|w|\leq h_{\rm a}(M)$. \end{lemma} In the next lemma we show that every dtla $M$ has an ancestral bound. It will be convenient to prove a slightly stronger result, where we consider more entry paths in $G_M$ than the ancestral ones. We say that a pair of strings $(w_1,w_2)\in\Delta_\B^*\times\Delta_\B^*$ \emph{splits} if there exist $d\in\Delta^{(k)}$ and $j_1,j_2\in[k]$ with $j_1\neq j_2$ such that the first symbol of $w_i$ is $(d,j_i)$.\footnote{The fact that $\diff(v_1,v_2)$ splits means that there exists a tree $t\in\T_\Delta$ such that $v_1$ and $v_2$ are branches of $t$ that correspond to distinct leaves of $t$.} A path $\pi$ in $G_M$ is \emph{nonsplitting} if $\diff(\pi)$ does not split, and \emph{splitting} if $\diff(\pi)$ splits. Note that an ancestral pair of strings does not split, and hence every ancestral path in $G_M$ is nonsplitting. Recall again from the paragraph before Theorem~\ref{th:uniform_earliest} that $\fix(M)$ is a fixed set of representatives of the equivalence classes $\sem{p}$, $p\in P$. We define $$\maxfix(M)=\max\{\height(q_M(s))\mid q\in Q,\,s\in\fix(M),\,\rho(q)=\delta(s)\}.$$ Note that since $\fix(M)$ is finite, $\maxfix(M)\in\Nat$. The next lemma implies that $\maxfix(M)+\maxdiff(M)$ is an ancestral bound for $M$ when $\diff(M)$ is finite, and so every dtla has an ancestral bound. \begin{lemma}\label{lm:no_split} Let $\pi$ be a nonsplitting entry path in $G_M$. If $\diff(\pi)=(w_1,w_2)$, then $|w_i|\leq \maxfix(M)+\maxdiff(M)$ for $i=1,2$. \end{lemma} \begin{proof} Let $\pi$ be a $(q_1,q_2,b)$-path with label $(u,v_1,v_2)$ and let $\diff(v_1,v_2)=(w_1,w_2)$. For reasons of symmetry, it suffices to show that $|w_1|\leq \maxfix(M)+\maxdiff(M)$. Thus, we assume that $w_1\neq\varepsilon$ (and hence $\pi\neq\varepsilon$). We consider two cases, depending on the value of $b$. \smallskip \emph{Case 1}: $b=1$. In this case we show that $|w_1|\leq \maxdiff(M)$. By Lemma~\ref{lm:path_to_context} there are $C\in\C_\Sigma$ and $p_1,p_2\in P$ such that $(q_i,u,\#)\in\blk_{C[p_i]}(v_i)$ for $i=1,2$, and $C/u=\bot$. Hence, cf. the proof of Lemma~\ref{lm:blink}, there is a reachable sentential form $\xi$ for $C[p_i]$ such that $v_i\in B(\xi)$ and $\xi/v_i=q_i(u)$. Since $C[p_i]/u=p_i$, Lemma~\ref{lm:elem}(2) implies that $v_i\in B(M(C[p_i]))$ and $M(C[p_i])/v_i=\angl{q_i,p_i}$. Let $v$ be the longest common prefix of $v_1$ and $v_2$. So $v_i=vw_i$. Since $(w_1,w_2)$ does not split, $\nod(v)$ is a difference node of $M(C[p_1])$ and $M(C[p_2])$ by Lemma~\ref{lm:diff_char}: each of its proper ancestors has the same label in $M(C[p_1])$ and $M(C[p_2])$, but the node itself does not have the same label in $M(C[p_1])$ and $M(C[p_2])$.\footnote{In fact, if $w_1$ and $w_2$ are both nonempty, then their first symbols are $(d_1,j_1),(d_2,j_2)$ with $d_1\neq d_2$, and $\nod(v)$ has label $d_i$ in $M(C[p_i])$. If $w_2=\varepsilon$, then $v=v_2$ and hence $\nod(v)$ has label $\angl{q_2,p_2}$ in $M(C[p_2])$; also, $\nod(vw_1)$ has label $\angl{q_1,p_1}$ in $M(C[p_1])$ and hence the label of $\nod(v)$ is in $\Delta$ (because $w_1\neq\varepsilon$).} Hence $|w_1|\leq \height(M(C[p_1])/v)\leq \maxdiff(M)$. \smallskip \emph{Case 0}: $b=0$. By Lemma~\ref{lm:path_to_context}, there exist $C\in\C_\Sigma$ and $p_1,p_2\in P$ such that $(q_i,u,\#)\in\blk_{C[p_i]}(v_i)$ and $C/u\in\fix(M)$. Note that $v_i$ is a branch of $M(C[p_i])$ by Lemma~\ref{lm:blink}, and that $M(C[p_i])/v_i=q_{iM}(C/u)$ by Lemma~\ref{lm:orig_prop1}(1). First assume that $w_2\neq\varepsilon$. This is similar to (1) above. Let $v$ be the longest common prefix of $v_1$ and $v_2$. So $v_i=vw_i$. Since $(w_1,w_2)$ does not split and both $w_1$ and $w_2$ are nonempty, $\nod(v)$ is a difference node of $M(C[p_1])$ and $M(C[p_2])$. So $|w_1|\leq \height(M(C[p_1])/v)\leq\maxdiff(M)$. Now assume that $w_2=\varepsilon$, and so $v_1=v_2w_1$. If $|w_1|\leq\maxfix(M)$ then we are ready. If $|w_1|>\maxfix(M)$, then $|w_1|>\height(q_{2M}(C/u))$ because $C/u\in\fix(M)$, and so $|w_1|>\height(M(C[p_2])/v_2)$. Thus, the branch $v_2w_1$ of $M(C[p_1])$ is longer than any branch of $M(C[p_2])$ with prefix $v_2$. By Lemma~\ref{lm:height-diff} (with $v:=v_2$ and $w:=w_1$) there is a difference node $\nod(v_2w_1')$ of $M(C[p_1])$ and $M(C[p_2])$ where $w_1'$ is a prefix of $w_1$. Since $\nod(v_2w_1')$ is a node of $M(C[p_2])$, $\;|w_1'|\leq\height(M(C[p_2])/v_2)=\height(q_{2M}(C/u))$. Hence, $|w_1|\leq |w_1'|+\height(M(C[p_1])/v_2w_1')\leq \height(q_{2M}(C/u))+\maxdiff(M)$. This shows that $|w_1|\leq\maxfix(M)+\maxdiff(M)$ by the definition of $\maxfix(M)$. \qed \end{proof} \begin{corollary}\label{cor:no_split_fin} If $\diff(M)$ is finite, then the set $\{\diff(\pi)\mid \pi$ is a nonsplitting entry path in $G_M\}$ is finite. \end{corollary} We now turn to the pumping of $\Sigma$-contexts, as discussed before. In terms of $G_M$ that corresponds to the repetition of a cycle. In the next two lemmas we show that pumping (i.e., repeating) a cycle in an ancestral entry path produces again an ancestral path, and in Lemma~\ref{lm:bound_graph} we show that this leads to infinitely many $\diff(\pi)$ values and hence to an infinite $\diff(M)$ by Corollary~\ref{cor:no_split_fin}. A path $\pi$ in $G_M$ is a cycle, and in particular a \emph{$(q_1,q_2,b)$-cycle}, if it is nonempty and leads from the node $(q_1,q_2,b)$ to itself; note that every node of $\pi$ is of type $b$. For every $i\geq 1$, we denote by $\pi^i$ the $i$-fold concatenation $\pi\cdots \pi$. \begin{lemma}\label{lm:drop_prime} Let $\diff(M)$ be finite. Let $\pi_0$ be a $(q_1,q_2,b)$-path in $G_M$, and $\pi$ be a $(q_1,q_2,b)$-cycle in $G_M$ with output label $(v_1,v_2)$ such that $v_1v_2\neq\varepsilon$. If $\pi_0\pi$ is nonsplitting, then it is ancestral. \end{lemma} \begin{proof} The idea of the proof is that if $\pi_0\pi$ is not ancestral, then we can pump $\pi$ and thus obtain nonsplitting paths with arbitrarily large $\diff$ values, contradicting Corollary~\ref{cor:no_split_fin}. Let $\diff(\pi_0\pi)=(w_1,w_2)$. Suppose that $\pi_0\pi$ is not ancestral, i.e., both $w_1$ and $w_2$ are nonempty. Since $(w_1,w_2)$ is nonsplitting, $(w_1,w_2)=((d_1,j_1)w_1',(d_2,j_2)w_2')$ with $d_1\neq d_2$. For every $k\geq 1$, consider the $(q_1,q_2,b)$-path $\pi_0\pi^{k+1}$. Now $\diff(\pi_0\pi^{k+1})=\diff(\diff(\pi_0\pi)\cdot\out(\pi^k))= \diff(w_1v_1^k,w_2v_2^k)=((d_1,j_1)w_1'v_1^k,(d_2,j_2)w_2'v_2^k)$. Hence $\pi_0\pi^{k+1}$ is nonsplitting. Also since $v_1v_2\neq\varepsilon$, we obtain that $|(d_i,j_i)w_i'v_i^k|>k$ for $i=1$ or $i=2$. Since this holds for every $k$, the set $\{\diff(\pi_0\pi^{k+1})\mid k\geq 1\}$ is infinite, contradicting Corollary~\ref{cor:no_split_fin}. \qed \end{proof} \begin{example}\label{ex:drop_prime} Let $M$ be the dtla of Example~\ref{ex:depgraph} and let $\pi_0=e_{01}$ and $\pi=e_1$. Then $\pi$ has output label $((\sigma_a,1),(\sigma_b,1))$ and $(\sigma_a,1)(\sigma_b,1)\neq\varepsilon$. The path $\pi_0\pi$ has output label $((\tau,1)(\sigma_a,1),(\tau,1)(\sigma_b,1))$ and so $\diff(\pi_0\pi)=((\sigma_a,1),(\sigma_b,1))$, which means that $\pi_0\pi$ is nonsplitting but not ancestral. Indeed, consider the pumped path $\pi_0\pi^{k+1}$. By Example~\ref{ex:path_to_context}, it has label $(1^{k+2},(\tau,1)(\sigma_a,1)^{k+1},(\tau,1)(\sigma_b,1)^{k+1})$ and satisfies the requirements of Lemma~\ref{lm:path_to_context} for $C=\tau(\sigma^{k+1}\bot,a)$ and for $p_1=p_a$ and $p_2=p_b$. By the proof of Lemma~\ref{lm:no_split}, that implies that the node $\nod((\tau,1))=1$ is a difference node of $M(C[p_a])$ and $M(C[p_b])$, and that $M(C[p_a])/1=(\sigma_a,1)^{k+1}\angl{q_{1a},p_a}$ is a difference tree of $M$. Hence $\diff(M)$ is infinite. \qed \end{example} In the proof of the next, technical pumping lemma we use that $M$ is ultralinear (with mapping $\mu: Q\to \Nat$) and b-erasing (with graph $E_M$). We note here the obvious fact that if $e$ is an edge in $G_M$ from $(q_1,q_2,b)$ to $(q_1',q_2',b')$, then $\mu(q_i)\leq\mu(q_i')$; and hence the same holds for paths in $G_M$. Thus, if $e$ is part of a cycle in $G_M$, then $\mu(q_i)=\mu(q_i')$. Also, if $e$ has output label $(v_1,v_2)$ and $v_i=\varepsilon$, then there is an edge from $q_i$ to $q_i'$ in $E_M$; and hence the same holds for paths (of the same length) in $G_M$ and $E_M$. Thus, if a $(q_1,q_2,b)$-cycle in $G_M$ has output label $(\bar{v}_1,\bar{v}_2)$, then both $\bar{v}_1$ and $\bar{v}_2$ are nonempty. \begin{lemma}\label{lm:cycle} Let $\diff(M)$ be finite. Let $\pi_0$ be a $(q_1,q_2,b)$-path in $G_M$ and $\pi$ be a $(q_1,q_2,b)$-cycle in $G_M$. If $\pi_0\pi$ is nonsplitting, then $\pi_0\pi^k$ is nonsplitting for every $k\geq 1$. \end{lemma} \begin{proof} Clearly, it suffices to show that $\pi_0\pi^2$ is nonsplitting. Suppose that there exist $\pi_0$ and $\pi$ such that $\pi_0\pi^2$ is splitting. Let $\pi=e_1\cdots e_n$ with $n\geq 1$ and $e_m$ an edge of $G_M$ for $m\in[n]$. Without loss of generality, we may assume that $\pi_0\pi e_1$ is splitting. In fact, suppose that $e_j$ is the first edge of $\pi$ that causes splitting, i.e., $\pi_0\pi e_1\cdots e_{j-1}$ is nonsplitting and $\pi_0\pi e_1\cdots e_j$ is splitting. Let $e_{j-1}$ lead to node $(\widetilde{q}_1,\widetilde{q}_2,b)$, and let $\widetilde{\pi}_0=\pi_0 e_1\cdots e_{j-1}$ and $\widetilde{\pi}=e_j\cdots e_ne_1\cdots e_{j-1}$. Then $\widetilde{\pi}_0$ is a $(\widetilde{q}_1,\widetilde{q}_2,b)$-path, $\widetilde{\pi}$ is a $(\widetilde{q}_1,\widetilde{q}_2,b)$-cycle, $\widetilde{\pi}_0\widetilde{\pi}$ is nonsplitting and $\widetilde{\pi}_0\widetilde{\pi}e_j$ is splitting. So, suppose that $\pi_0\pi e$ is splitting, where $e=e_1$. Let $\pi_0\pi$ have label $(u,v_1,v_2)$, let $\diff(\pi_0\pi)=\diff(v_1,v_2)=(w_1,w_2)$, and let $\pi$ have output label $(\bar{v}_1,\bar{v}_2)$, where $\bar{v}_i$ is a postfix of $v_i$. Since $M$ is b-erasing, $\bar{v}_1\neq\varepsilon$, and $\bar{v}_2\neq\varepsilon$ (because if $\bar{v}_i=\varepsilon$, then there is a cycle of length $n$ from $q_i$ to $q_i$ in $E_M$). Consequently, $\pi_0\pi$ is ancestral by Lemma~\ref{lm:drop_prime}, i.e., $w_1=\varepsilon$ or $w_2=\varepsilon$. Let $e$ end at the node $(q_1',q_2',b)$ and let $(r_1,r_2,j,z_1,z_2)$ be its label. Since $\pi_0\pi e$ is splitting, $\diff(\pi_0\pi e)=\diff(v_1z_1,v_2z_2)=\diff(w_1z_1,w_2z_2)=((d,j_1)y_1,(d,j_2)y_2)$ with $j_1\neq j_2$. Assume that $w_1=\varepsilon$, i.e., $v_2=v_1w_2$ (the case where $w_2=\varepsilon$ is analogous). So $\diff(z_1,w_2z_2)=((d,j_1)y_1,(d,j_2)y_2)$. Let $y$ be the longest common prefix of $z_1$ and $w_2z_2$, i.e., $z_1=y(d,j_1)y_1$ and $w_2z_2=y(d,j_2)y_2$. Then $v_1y$ is the longest common prefix of $v_1z_1$ and $v_2z_2$, because $v_1z_1=v_1y(d,j_1)y_1$ and $v_2z_2=v_1w_2z_2=v_1y(d,j_2)y_2$. Now recall that $z_1$ is a branch of the right-hand side $\zeta_1$ of $r_1$ such that $\zeta_1/z_1=q'_1(x_j)$. Since $M$ is ultralinear, we have $\mu(q_1)\leq\mu(q_1')$. We also have $\mu(q_1')\leq\mu(q_1)$, because $e_2\cdots e_n$ is a path in $G_M$ from $(q_1',q_2',b)$ to $(q_1,q_2,b)$. Hence $\mu(q_1)=\mu(q_1')$, and so, again by ultralinearity, there is no other occurrence of $x_j$ in $\zeta_1$. In particular, since $\zeta_1/y(d,j_1)y_1=\zeta_1/z_1=q'_1(x_j)$, the subtree $\zeta_1/y(d,j_2)$ does not contain $x_j$. Let $\pi'=e_2\cdots e_n$, so $\pi=e\pi'$. Let $k\in\Nat$ be such that $k>\maxdiff(M)+\maxrhs(M)+\maxfix(M)$, and let $\pi_k$ be the pumped path $\pi_0\pi^{k+2}=\pi_0\pi e\pi'\pi^k$. By Lemma~\ref{lm:path_to_context}, applied to $\pi_k$, there exist $C\in\C_\Sigma$ and $p_1,p_2\in P$ such that requirements~(a), (b), and~(c) of that lemma hold for the prefix $\pi_0\pi$ of $\pi_k$. By~(a), $(q_1,u,\#)\in\blk_{C[p_1]}(v_1)$. Hence, since $\rhs(q_1,C[p_1],u)=\rhs(r_1)=\zeta_1$ by (c)(1), it follows from Lemma~\ref{lm:blink} and the definition of $\lk$ that $(q_1,u,\nod(y))\in\lk_{C[p_1]}(v_1y)$ and hence $\orig_{C[p_1]}(v_1y)=(q_1,u,\nod(y))$. So, by Lemma~\ref{lm:orig_prop1}(2), $M(C[p_1])/v_1y(d,j_2)= \zeta_1/y(d,j_2)[\bar{q}(x_{j'})\leftarrow \bar{q}_M(C[p_1]/uj')\mid\bar{q}\in Q,\,j'\in\Nat_+]$. For every $j'\in\Nat_+$ with $j'\neq j$ and $uj'\in V(C)$, the node $uj'$ is not on the spine of $C$ because, by~(b), $u$ and $uj$ are both on or both off the spine. Hence, by (c)(2), $C[p_1]/uj'=C/uj'\in\fix(M)$ and so $\height(\bar{q}_M(C[p_1]/uj'))\leq\maxfix(M)$. Since as observed above, $\zeta_1/y(d,j_2)$ does not contain $x_j$, it follows that $\height(M(C[p_1])/v_1y(d,j_2))\leq\maxrhs(M)+\maxfix(M)<k$. Let $\pi'$ have label $(v_1',v_2')$. Then $\out(\pi_k)=(v_1z_1v_1'\bar{v}_1^k,v_2z_2v_2'\bar{v}_2^k)$ and since (a) also holds for the path $\pi_k$ itself, $v_2z_2v_2'\bar{v}_2^k$ is a branch of $M(C[p_2])$ by Lemma~\ref{lm:blink}. Recall that $v_2z_2=v_1w_2z_2=v_1y(d,j_2)y_2$. Hence $y_2v_2'\bar{v}_2^k$ is a branch of $M(C[p_2])/v_1y(d,j_2)$. Since $\bar{v}_2\neq\varepsilon$, $|y_2v_2'\bar{v}_2^k|\geq k$.\footnote{ In the case where $w_2'=\varepsilon$, we need here that $\bar{v}_1\neq\varepsilon$.} By Lemma~\ref{lm:height-diff} (with $v:= v_1y(d,j_2)$ and $w:=y_2v_2'\bar{v}_2^k$, and with $p_1$ and $p_2$ interchanged), there is a prefix $w'$ of $y_2v_2'\bar{v}_2^k$ such that $\nod(v_1y(d,j_2)w')$ is a difference node of $M(C[p_1])$ and $M(C[p_2])$. Since $\nod(w')$ is a node of $M(C[p_1])/v_1y(d,j_2)$, we have $|w'|\leq \maxrhs(M)+\maxfix(M)$. So $\height(M(C[p_2])/v_1y(d,j_2)w')\geq |y_2v_2'\bar{v}_2^k|-|w'|\geq k-(\maxrhs(M)+\maxfix(M))> \maxdiff(M)$, a contradiction. \qed \end{proof} \begin{example}\label{ex:cycle} Let $M$ be the dtla of Example~\ref{ex:depgraph} and let $\pi_0=e_{02}$ and $\pi=e_2$. The path $\pi_0\pi$ has output label $((\tau,2)(\tau,1),(\tau,2)(\tau,1)(\tau,2))$ and so $\diff(\pi_0\pi)=(\varepsilon,(\tau,2))$, which means that $\pi_0\pi$ is ancestral. On the other hand, the path $\pi_0\pi^2$ has output label $((\tau,2)(\tau,1)(\tau,1),(\tau,2)(\tau,1)(\tau,2)(\tau,1)(\tau,2))$ which implies that $\diff(\pi_0\pi^2)=((\tau,1),(\tau,2)(\tau,1)(\tau,2))$, and so $\pi_0\pi^2$ is splitting. Indeed, consider the pumped path $\pi_k=\pi_0\pi^{k+2}$, which has label $(2\cdot 1^{k+2},(\tau,2)(\tau,1)^{k+2},(\tau,2)((\tau,1)(\tau,2))^{k+2})$ and satisfies the requirements of Lemma~\ref{lm:path_to_context} for $C=\tau(\bot,\sigma^{k+2}a)$ and for $p_1=p_a$ and $p_2=p_b$, by Example~\ref{ex:path_to_context}. Now $M(C[p_a])=\tau(\angl{q_{1a},p_a},\tau(\tau(q_{2aM}(\sigma^ka),a),a))$ and $M(C[p_b])=\tau(\angl{q_{1a},p_a},\tau(\tau(a,\tau(\tau(a,q_{2bM}(\sigma^ka)),a)),a))$. Consequently, the node $\nod((\tau,2)(\tau,1)(\tau,2))=(2,1,2)$ is a difference node of $M(C[p_a])$ and $M(C[p_b])$, with $M(C[p_a])/(2,1,2)=a$ and $M(C[p_b])/(2,1,2)=\tau(\tau(a,q_{2bM}(\sigma^ka)),a)$. So, $\tau(\tau(a,q_{2bM}(\sigma^ka)),a)$ is a difference tree of $M$ and has a branch $((\tau,1)(\tau,2))^{k+1}$, for every $k$. Hence $\diff(M)$ is infinite, which, of course, we already knew from Example~\ref{ex:drop_prime}. \qed \end{example} Finally, we prove that $(2\cdot|Q|^2-1)\cdot\maxrhs(M)$ is an ancestral bound for $M$. \begin{lemma}\label{lm:bound_graph} Let $\diff(M)$ be finite. Let $\pi$ be an ancestral entry path in $G_M$. \\ If $\diff(\pi)=(w,\varepsilon)$, then $|w|\leq (2\cdot|Q|^2-1)\cdot\maxrhs(M)$. \end{lemma} \begin{proof} Suppose that $|w|> (2\cdot|Q|^2-1)\cdot\maxrhs(M)$. For a prefix $\pi'$ of $\pi$, we will denote the first component of $\diff(\pi')$ by $\diff_1(\pi')$. So, $|\diff_1(\varepsilon)|=0$ and $|\diff_1(\pi)|> (2\cdot|Q|^2-1)\cdot\maxrhs(M)$. We also observe that, by definition of $G_M$, every edge $e$ of $\pi$ adds at most $\maxrhs(M)$ to $|\diff_1(\pi')|$; formally, if $\pi'e$ is a prefix of $\pi$, then $|\diff_1(\pi'e)| \leq |\diff_1(\pi')|+\maxrhs(M)$, because $\diff(\pi'e)=\diff(\diff(\pi')\cdot\out(e))$ and so $\diff_1(\pi'e)$ is a postfix of $\diff_1(\pi')z_1$ if $\out(e)=(z_1,z_2)$. Let $\pi_0=\varepsilon$ and for $1\leq i\leq 2\cdot|Q|^2$, let $\pi_i$ be the shortest prefix of $\pi$ such that $|\diff_1(\pi_i)|>|\diff_1(\pi_{i-1})|$. Obviously, $\pi_{i-1}$ is a proper prefix of $\pi_i$. Moreover, by the above observation, $|\diff_1(\pi_i)| \leq |\diff_1(\pi_{i-1})|+\maxrhs(M)$ and hence $|\diff_1(\pi_i)| \leq i\cdot\maxrhs(M)$, which shows that $\pi_i$ is well defined for every $i\leq 2\cdot|Q|^2$. Note that $\pi_i$ is ancestral and $\diff(\pi_i)=(w_i,\varepsilon)$ for some branch $w_i$. Since we have defined $2\cdot|Q|^2+1$ prefixes of $\pi$, there exist $\pi_i$ and $\pi_j$ with $i<j$ that are $(q_1,q_2,b)$-paths for the same node $(q_1,q_2,b)$ of $G_M$. So, $\pi_i$ is a prefix of $\pi_j$ and $|w_j|>|w_i|$. Let $\bar{\pi}$ be the $(q_1,q_2,b)$-cycle such that $\pi_j=\pi_i\bar{\pi}$, and let $\out(\bar{\pi})=(v_1,v_2)$. Then $\diff(\pi_j)=\diff(\diff(\pi_i)\cdot\out(\bar{\pi}))$ and so $(w_j,\varepsilon)=\diff(w_iv_1,v_2)$, i.e., $w_iv_1=v_2w_j$. Since $|w_i|<|w_j|$, we obtain that $|v_1|>|v_2|$. We now pump $\bar{\pi}$ and consider the path $\pi_i\bar{\pi}^k$ for every $k\geq 1$. Let $(w_{1,k}',w_{2,k}')=\diff(\pi_i\bar{\pi}^k)=\diff(\diff(\pi_i)\cdot\out(\bar{\pi}^k))= \diff(w_iv_1^k,v_2^k)$. By Lemmas~\ref{lm:cycle} and~\ref{lm:drop_prime}, $\pi_i\bar{\pi}^k$ is ancestral. Since $|w_iv_1^k|>|v_2^k|$, it follows that $v_2^k$ is a prefix of $w_iv_1^k$, and so $w_{2,k}'=\varepsilon$ and $w_iv_1^k=v_2^kw_{1,k}'$. Hence $|w_{1,k}'|=|w_iv_1^k|-|v_2^k|=|w_i|+k(|v_1|-|v_2|)\geq k$. Since this holds for every $k\geq 1$, it contradicts Corollary~\ref{cor:no_split_fin}. \qed \end{proof} \begin{theorem}\label{th:ancestral_bound} The number $(2\cdot|Q|^2-1)\cdot\maxrhs(M)$ is an ancestral bound for $M$. \end{theorem} \begin{proof} Immediate from Lemmas~\ref{lm: ancbound_redef2} and~\ref{lm:bound_graph}. \qed \end{proof} We now present our second main result. \begin{theorem}\label{th:ultra} It is decidable for a total ultralinear bounded erasing dtla $M$ whether there exists a dtop $N$ such that $\sem{M}=\sem{N}$, and if so, such a dtop $N$ can be constructed. \end{theorem} \begin{proof} We use Corollary~\ref{co:goal} for the class $\U$ of total ultralinear b-erasing dtlas. We first consider an initialized and la-uniform $M\in\U$. As observed in Section~\ref{sec:upper}, an output bound $h_{\rm o}(M)$ for $M\in\U$ can be computed from $M$; in fact, by Theorem~\ref{th:output_bound}, $h_{\rm o}(M)=\maxrhs(M)\cdot|Q|\cdot(|P|+2)$ is an output bound for $M$. By Theorem~\ref{th:ancestral_bound}, $h_{\rm a}(M)=(2\cdot|Q|^2-1)\cdot\maxrhs(M)$ is an ancestral bound for $M$ (which, of course, can be computed from $M$). Finally, Theorem~\ref{th:upper_bound} shows that $2\cdot\maxrhs(M)+h_{\rm o}(M)+h_{\rm a}(M)+1= 1+\maxrhs(M)\cdot(|Q|\cdot(|P|+2)+2\cdot|Q|^2+1)$ is a difference bound for $M$, and hence so is the larger number $1+2\cdot\maxrhs(M)\cdot(|Q|+|P|)^2$. It now follows from Lemmas~\ref{lm:make_initialized} and~\ref{lm:make_uniform} that we obtain a difference bound for \emph{every} $M\in\U$ by first changing $\maxrhs(M)$ into $2\cdot\maxrhs(M)$ and $|Q|$ into $(|Q|+1)\cdot|P|$, and then taking the maximum with $\maxrhs(M)$ (cf. the paragraphs after Lemmas~\ref{lm:make_initialized} and~\ref{lm:make_uniform}). Thus, $\U$ has Property~(H) of Corollary~\ref{co:goal} with the computable mapping $h:\U\to\Nat$ such that $h(M)=1+4\cdot\maxrhs(M)\cdot(|Q|+2)^2\cdot|P|^2$. \qed \end{proof} The same result holds for a total output-monadic dtla $M$, where \emph{output-monadic} means that $\rk(d)\leq 1$ for every $d\in\Delta$. The reason is that Lemma~\ref{lm:cycle}, which is the only result that needs the assumption that $M$ is ultralinear and bounded erasing, is trivial if $M$ is output-monadic, because in that case every entry path in $G_M$ is nonsplitting. Note that every output-monadic dtla is (ultra)linear, but not necessarily bounded erasing. This result is a slight generalization of the result of \cite{DBLP:journals/tcs/Choffrut77} for string transducers (more precisely, subsequential functions). The same result of Theorem~\ref{th:ultra} also holds for a total initialized depth-uniform dtla $M$, where \emph{depth-uniform} means that for every $a\in\Sigma^{(k)}$ and every $j\in[k]$ there exists a number $d_{a,j}\in\Nat$ such that for every rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to\zeta$ and every $z\in V(\zeta)$, if $\zeta/z=q'(x_j)$ then $|z|=d_{a,j}$. In other words, every translation of the $j$th child of a node with label $a$ is at the same depth of the right-hand side of every rule for $a$. This property implies that $|z_1|=|z_2|$ for every label $(r_1,r_2,j,z_1,z_2)$ of an edge of $G_M$, hence $|v_1|=|v_2|$ for every output label $(v_1,v_2)$ of an entry path in $G_M$, and hence $\diff(\pi)=(\varepsilon,\varepsilon)$ for every ancestral entry path in $G_M$. Thus, $h_{\rm a}(M)=0$ is an ancestral bound for $M$, by Lemma~\ref{lm: ancbound_redef2}. \begin{example} Let $\Sigma=\{\sigma^{(1)},\tau^{(1)},a^{(0)},b^{(0)}\}$ and $\Delta=\{\sigma_a^{(2)},\sigma_b^{(2)},a^{(0)},b^{(0)}\}$. Consider the initialized la-uniform dtla $M$ that translates every input tree $s$ with $n$ occurrences of $\sigma$ into the full binary tree of height~$n$ over $\{\sigma_y,y\}$, where $y$ is the label of the leaf of $s$. For instance, $M(\sigma\tau\sigma a)= \sigma_a(\sigma_a(a,a),\sigma_a(a,a))$. It has look-ahead states $p_a$ and $p_b$, with the usual transitions, and states $q_a$ and $q_b$. For $y\in\{a,b\}$, it has axioms $A(p_y)=q_y(x_0)$ and rules $q_y(\sigma(x_1\ot p_y))\to \sigma_y(q_y(x_1),q_y(x_1))$, $\;q_y(\tau(x_1\ot p_y))\to q_y(x_1)$, and $q_y(y)\to y$. Clearly $M$ is depth-uniform, with $d_{\sigma,1}=1$ and $d_{\tau,1}=0$. Note that $M$ is neither ultralinear nor b-erasing. \qed \end{example} The depth-uniform property can be weakened by defining it as follows: there is a number $h'_{\rm a}(M)\in\Nat$ such that for every entry path $\pi$ in $G_M$, if $\out(\pi)=(v_1,v_2)$ then $|\,|v_1|-|v_2|\,|\leq h'_{\rm a}(M)$, i.e., the distance between $|v_1|$ and $|v_2|$ is at most $h'_{\rm a}(M)$. Clearly, such a number $h'_{\rm a}(M)$ is an ancestral bound for $M$ by Lemma~\ref{lm: ancbound_redef2}. This weak depth-uniform property is decidable, and if it holds then a bound $h'_{\rm a}(M)$ can be computed. In fact, it is easy to see that $M$ is weak depth-uniform if and only if $|v_1|=|v_2|$ for every output label $(v_1,v_2)$ of a cycle in $G_M$. This can be checked by considering all simple cycles in $G_M$, i.e., cycles in which no node of $G_M$ occurs more than once. The bound $h'_{\rm a}(M)$ can then be computed by considering all entry paths in $G_M$ that do not contain a cycle. \section{Introduction} Many simple tree transformations can be modeled by top-down tree transducers~\cite{DBLP:journals/mst/Rounds70,DBLP:journals/jcss/Thatcher70}. They are recently used in XML database theory (e.g., \cite{DBLP:journals/jcss/EngelfrietMS09,haruo,DBLP:conf/pods/LemayMN10,DBLP:conf/dbpl/ManethN99,DBLP:journals/jcss/MartensN07,DBLP:journals/iandc/MartensNG08}), in computational linguistics (e.g., \cite{DBLP:conf/cicling/KnightG05,DBLP:conf/mol/Maletti11,DBLP:journals/siamcomp/MalettiGHK09}) and in picture generation \cite{drewes}. A \emph{top-down tree transducer} is a finite-state device that scans the input tree in a (parallel) top-down fashion, simultaneously producing the output tree in a (parallel) top-down fashion. A more expressive (but also more complex) model for specifying tree translations is the top-down tree transducer \emph{with regular look-ahead}~\cite{DBLP:journals/mst/Engelfriet77}. It consists of a top-down tree transducer and a finite-state bottom-up tree automaton, called the look-ahead automaton. We may think of its execution in two phases: In a first phase the input tree is relabeled by attaching to each input node the active state of the automaton, called the look-ahead state at that node. In the second phase the top-down tree transducer is executed over the relabeled tree, thus possibly making use of the look-ahead information in the new input labels. As an example, consider a transducer $M_{\text{ex}}$ of which the look-ahead automaton checks whether the input tree contains a leaf labeled $a$. If so, then $M_{\text{ex}}$ outputs $a$, and otherwise it outputs a copy of the input tree. It should be clear that there is no top-down tree transducer (without look-ahead) that realizes the same translation as $M_{\text{ex}}$. The intuitive reason is that in general the complete input tree must be read and buffered in memory, before the appropriate choice of output can be made. How can we formally prove that indeed no top-down tree transducer (without look-ahead) can realize this translation? In general, is there a method to determine for a given top-down tree transducer \emph{with look-ahead}, whether or not its translation can be realized by a top-down tree transducer \emph{without look-ahead}? And if the answer is yes, can such a transducer be constructed from the given one? In this paper we give two partial answers to these questions, where we restrict ourselves to total deterministic transducers (which will not be mentioned any more in what follows). For such transducers we provide a general method as discussed above. However, part of the method is not automatic, but depends on additional knowledge about the given transducer with look-ahead (which can usually be determined by the designer of the transducer). For a restricted type of transducers (where the restrictions concern the capability of the transducer to copy and erase) that knowledge can also be obtained automatically, which means that for a thus restricted transducer with look-ahead it is decidable whether its translation can be realized by a (nonrestricted) transducer without look-ahead, and if so, such a transducer can be constructed from the given transducer. The main notion on which our method is based, is that of a \emph{difference tree} of a top-down tree transducer with regular look-ahead. Consider two trees obtained from one input tree by replacing one of its leaves by two different look-ahead states of the transducer $M$. Compare now the two output trees of $M$ on these input trees, where $M$ treats the look-ahead state as representing an input subtree for which the look-ahead automaton of $M$ arrives in that state at the root of the subtree.\footnote{Since $M$ is total and deterministic, the output trees exist and are unique, respectively. } By removing the largest common prefix of these two output trees (i.e., every node of which every ancestor has the same label in each of the two trees), we obtain a number of output subtrees that we call difference trees of $M$. Intuitively, the largest common prefix is the part of the output that does not depend on the two possible look-ahead states of the subtree, whereas a difference tree is a part of the output that can be produced because $M$ knows the look-ahead state of the subtree. Thus, the set $\diff(M)$ of all difference trees of $M$ can be viewed as a measure of the impact of the look-ahead on the behaviour of $M$. For the example transducer $M_{\text{ex}}$ above, $\diff(M_{\text{ex}})$ consists of the one-node tree $a$ and all trees of which no leaf is labeled $a$ (with one leaf representing a subtree without $a$-labeled leaves); thus, $\diff(M_{\text{ex}})$ is infinite. Now the idea of our method is as follows, where we use \emph{dtla} and \emph{dtop} to abbreviate top-down tree transducer with and without look-ahead, respectively. In~\cite{DBLP:journals/jcss/EngelfrietMS09} it was shown that for every dtop an equivalent canonical earliest dtop can be constructed. Earliest means that each output node is produced as early as possible by the transducer, and canonical means that different states of the transducer are inequivalent. We prove that also for every dtla an equivalent canonical earliest dtla (with the same look-ahead automaton) can be constructed, where the earliest and canonical properties are relativized with respect to each look-ahead state. Thus, to devise our method we may restrict attention to canonical earliest transducers. Assume there exists a (canonical earliest) dtop $N$ equivalent to the (canonical earliest) dtla $M$. Then the dtla $M$ is at least as early as $N$. In other words, at each moment of the translation, $M$ may be ahead of $N$ but not vice versa, i.e., the output of $N$ is a prefix of that of $M$, which is because $M$ has additional information through its look-ahead. The output of $N$ is the part of $M$'s output that does not depend on the look-ahead state. Thus, when removing the output of $N$ from that of $M$, the remaining trees are difference trees of $M$. Since $N$ must be able to simulate $M$, it has to store these difference trees in its states. Hence, $\diff(M)$ must be finite. Moreover, it turns out that the above description of $N$'s behaviour completely determines $N$, and so, roughly speaking, $N$ can be constructed from $M$ and $\diff(M)$, and then tested for equivalence with $M$ (\cite{DBLP:journals/actaC/Esik81}). Note that since $\diff(M_{\text{ex}})$ is infinite, the translation of $M_{\text{ex}}$ cannot be realized by a dtop. A natural number $h$ is a \emph{difference bound} for a dtla $M$ if the following holds: if $M$ has finitely many difference trees, then $h$ is an upper bound on their height; in other words, if a tree in $\diff(M)$ has height $>h$, then $\diff(M)$ is infinite. Our first main result is that it is decidable for a given dtla $M$ for which a difference bound is also given, whether $M$ is equivalent to a dtop $N$, and if so, such a dtop $N$ can be constructed. We do not know whether a difference bound can be computed for every dtla $M$, but the designer of $M$ will usually be able to determine $\diff(M)$ and hence a difference bound for $M$. Our second main result is that a difference bound can be computed for dtlas that are ultralinear and bounded erasing. Ultralinearity means that the transducer cannot copy an input subtree when it is in a cycle (i.e., in a computation that starts and ends in the same state). Thus it is weaker than the linear property (which forbids copying) but stronger than the finite-copying property \cite{DBLP:journals/jcss/EngelfrietRS80,DBLP:journals/iandc/EngelfrietM99}. The latter implies that the size of the output tree of an ultralinear dtla is linear in the size of its input tree. Bounded erasing means that the transducer has no cycle in which no output is produced. The proof that a difference bound can be computed for ultralinear and bounded erasing dtlas, is based on pumping arguments that are technically involved. The paper is structured as follows. Section~\ref{sec:prelim} contains basic terminology, in particular concerning prefixes of trees. Nodes of trees are represented by Dewey notation. Section~\ref{sec:dtop} defines the dtla (deterministic top-down tree transducer with regular look-ahead) and discusses some of its basic properties. It also explains the treatment of look-ahead states that occur in the input tree. In Section~\ref{sec:difftree} we define the notions of difference tree and difference bound, illustrated by some examples. In Section~\ref{sec:normform} we discuss some normal forms for dtlas, in particular \emph{look-ahead uniformity} which is technically convenient. We prove that for every dtla $M$ there is an equivalent canonical earliest dtla $M'$ (which is also look-ahead uniform), and we show how to compute a difference bound for $M'$ from one of $M$. Our first main result is proved in Section~\ref{sec:diff} (Theorem~\ref{th:alg}), refined in Section~\ref{sec:avoid-eq} and illustrated in Section~\ref{sec:algex}. Section~\ref{sec:diff} starts with the definition of a \emph{difference tuple} of a dtla $M$, which generalizes the notion of difference tree by considering all look-ahead states of $M$ rather than just two. If $N$ is a dtop equivalent to $M$, then its states are in one-to-one correspondence with the difference tuples of $M$ (assuming that both $N$ and $M$ are canonical earliest), see Lemma~\ref{lm:Qdiff}. Sections~\ref{sec:origins}--~\ref{sec:graphs} are devoted to the proof of our second main result (Theorem~\ref{th:ultra}). Section~\ref{sec:origins} continues Section~\ref{sec:dtop} by discussing some basic properties of dtlas: the \emph{links} that exist between an input tree and its corresponding output tree, and for each node of the output tree, its \emph{origin} in the input tree. In Section~\ref{sec:upper} the problem of computing a difference bound for a dtla $M$ is reduced to that of computing two related upper bounds: an \emph{output bound} for $M$ and an \emph{ancestral bound} for $M$. An output bound can be computed for every dtla. Finally, in Section~\ref{sec:graphs}, an ancestral bound is computed for every ultralinear and bounded erasing dtla. The computed output bound (in the previous section) and ancestral bound for a dtla $M$ are both based on pumping arguments (simple for the output bound, complicated for the ancestral bound). In both cases a part of the input tree on which $M$ has a cyclic computation, is pumped in such a way that the corresponding output tree contains arbitrarily large difference trees. In the ancestral case the pumping argument is technically based on the fact that $M$ cannot copy and must produce output during its cyclic computations. Since the pumping of trees makes it hard to address nodes by Dewey notation, a \emph{dependency graph} is defined for $M$ such that a cyclic computation of $M$ corresponds to a cycle in its dependency graph; pumping the input tree then corresponds to repeating a cycle in the graph. At the end of Section~\ref{sec:graphs} we consider two other classes of total dtlas for which equivalence to a dtop is decidable (and if so, such a dtop can be constructed): \emph{output-monadic} dtlas and \emph{depth-uniform} dtlas. Output-monadic means that every node of an output tree has at most one child. Depth-uniform means, in its simplest form, that all states in the right-hand sides of the rules of the dtla are at the same depth. \subsubsection*{Related Work.} For deterministic string transducers with regular look-ahead, look-ahead removal is decidable, i.e., it is decidable whether a given transducer with look-ahead is equivalent to a transducer without look-ahead, and if so, such a transducer can be constructed. This was proved in \cite{DBLP:journals/tcs/Choffrut77} (see also \cite[Theorem IV.6.1]{berstel}), for so-called subsequential functions. We extend that result (for the total case) by proving that look-ahead removal is decidable for output-monadic dtlas. Look-ahead has been investigated for other types of tree transducers. For macro tree transducers \cite{DBLP:journals/jcss/EngelfrietV85,DBLP:journals/iandc/EngelfrietM99} and streaming tree transducers \cite{DBLP:conf/icalp/AlurD12}, regular look-ahead can always be removed. The same is true for nondeterministic visibly pushdown transducers \cite{DBLP:conf/sofsem/FiliotS12}. For deterministic visibly pushdown transducers the addition of regular look-ahead increases their power, but the decidability of look-ahead removal for these transducers is not studied in \cite{DBLP:conf/sofsem/FiliotS12}. In~\cite{DBLP:journals/ipl/FulopKV04} the deterministic multi bottom-up tree transducer (dmbot) was introduced and shown to have (effectively) the same expressive power as the dtla. Thus, our results can also be viewed as partial answers to the question whether it is decidable for a given dmbot to be equivalent to a dtop. \subsubsection*{Note.} The results of this paper were first presented at DLT~2014, see \cite{DBLP:conf/dlt/EngelfrietMS14}. \section{Links and Origins} \label{sec:origins} In this section we define two basic concepts for dtlas, and discuss some of their properties. Thus, this section can be viewed as a sequel to Section~\ref{sec:dtop}. \medskip {\bf Convention.} In this section and the next two sections we assume that $M$ is a dtla that is {\bf initialized} with initial states $q_{0,p}$ for $p\in P$ and {\bf la-uniform} with la-map $\rho: Q\to P$, cf. Lemmas~\ref{lm:make_initialized} and~\ref{lm:make_uniform}. Since $M$ is la-uniform, all its initial states $q_{0,p}$ are distinct. \medskip If $M(s)=t$, then every node $v$ of the output tree $t\in\T_\Delta$ is produced (together with its label) at a certain node $u$ of the input tree $s\in\T_\Sigma$ by the application of a rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta$ at $u$. The node $v$ is an instantiation of a node $z$ of $\zeta$ with label in $\Delta$. The triple $(q,u,z)$ will be called the ``origin'' of $v$, cf.~\cite{deursen}. In the special case that $z=\varepsilon$, there exist pairs $(q',u')$ for which there is a reachable sentential form $\xi$ for $s$ such that $\xi/v=q'(u')$. In such a case $v$ is already a node of $\xi$, but does not yet have a label in $\Delta$. In fact, $(q,u)$ is such a pair, but there can be more due to the presence of erasing rules: if $(q',u')$ is such a pair and $\rhs(q',s,u')=q''(x_i)$, then $(q'',u'i)$ is also such a pair. A pair $(q',u')$ will be called a ``link'' to $v$ (cf.~\cite{DBLP:conf/mol/Maletti11}). For convenience it will be denoted as a triple $(q',u',\#)$, and the origin $(q,u,z)$ of $v$ will also be called a link to $v$. We now formally define \emph{links}. For $s\in\T_\Sigma$ and every $v\in\Nat_+^*$, the sets $\lk_s(v)$ of triples $(q,u,z)$ with $q\in Q$, $u\in V(s)$ and $z\in\Nat_+^*\cup\{\#\}$, are defined recursively to be the smallest sets such that: \begin{enumerate} \item[(1)] If $\delta(s)=p$, then $(q_{0,p},\varepsilon,\#)\in\lk_s(\varepsilon)$. \item[(2)] If $(q,u,\#)\in\lk_s(v)$ and $\rhs(q,s,u)=\zeta$, then \begin{enumerate} \item $(q,u,z)\in\lk_s(vz)$ for every $z\in V_\Delta(\zeta)$, and \item $(\bar{q},ui,\#)\in\lk_s(vz)$ for every $z\in V(\zeta)$ with $\zeta/z=\bar{q}(x_i)$. \end{enumerate} \end{enumerate} The above intuition about links is proved in the next two lemmas. \begin{lemma}\label{lm:linkdef} Let $s\in\T_\Sigma$, $q\in Q$, $u\in V(s)$, and $z,v\in\Nat_+^*$. Then \begin{enumerate} \item[(1)] $(q,u,\#)\in\lk_s(v)$ if and only if there is a sentential form $\xi$ of $M$ such that $q_{0,\delta(s)}(\varepsilon)\Rightarrow_s^*\xi$ and $\xi/v=q(u)$, \item[(2)] $(q,u,z)\in\lk_s(v)$ if and only if there are $\hat{v}\in\Nat_+^*$ and a sentential form $\xi$ of $M$ such that $v=\hat{v}z$, $\,q_{0,\delta(s)}(\varepsilon)\Rightarrow_s^*\xi$, $\,\xi/\hat{v}=q(u)$, and $z\in V_\Delta(\rhs(q,s,u))$, \item[(3)] $\lk_s(v)\neq\emptyset$ if and only if $v\in V(M(s))$, and \item[(4)] if $(q,u,\#)\in\lk_s(v)$ or $(q,u,z)\in\lk_s(v)$, then $\rho(q)=\delta(s/u)$. \end{enumerate} \end{lemma} \begin{proof} Since the definition of $\lk_s$ closely follows the semantics of $M$, (1) and~(2) are easy to prove by recursion induction on the definition of $\lk_s$ in one direction, and by induction on the length of the computation $q_{0,\delta(s)}(\varepsilon)\Rightarrow_s^*\xi$ in the other direction. Clearly, the right-hand side of~(1) implies that $v\in V(M(s))$, by Lemma~\ref{lm:elem}(2) applied to $\xi\Rightarrow_s^* M(s)$, and the right-hand side of~(2) also implies that $v\in V(M(s))$ because there is a computation step $\xi \Rightarrow_s \xi'=\xi[\hat{v}\leftarrow \rhs(q,s,u)[x_i\leftarrow ui\mid i\in\Nat_+]]$ with $v\in V(\xi')$. Thus, if $\lk_s(v)\neq\emptyset$ then $v\in V(M(s))$. Now let $v\in V(M(s))$ and consider, in a computation of $M$ that translates $s$ into $M(s)$, the first sentential form $\xi$ such that $v\in V_{\Delta\cup Q}(\xi)$. If $v\in V_Q(\xi)$, then $\lk_s(v)$ contains a triple $(q,u,\#)$ by~(1). If $v\in V_\Delta(\xi)$, then $v$ is produced in the computation step $\xi' \Rightarrow_s \xi$ where $\xi'$ is the previous sentential form, and then $\lk_s(v)$ contains a triple $(q,u,z)$ by~(2) applied to $\xi'$. Thus, $\lk_s(v)\neq\emptyset$. (4) follows from~(1) and~(2) by Lemma~\ref{lm:prop_la-uniform}(4). \qed \end{proof} \begin{lemma}\label{lm:linkform} For every $v\in V(M(s))$, either $\lk_s(v)=\{(q,u,z)\}$ with $z\in\Nat_+^*$ and $z\neq\varepsilon$, or $\lk_s(v)=\{(q_1,u_1,\#),\dots,(q_n,u_n,\#),(q_n,u_n,\varepsilon)\}$ with $n\geq 1$ and for every $j\in[n-1]$ there exists $i_j\in\Nat_+$ such that $\rhs(q_j,s,u_j)=q_{j+1}(x_{i_j})$ and $u_{j+1}=u_ji_j$. \end{lemma} \begin{proof} Let us say that a mapping $\lambda$ from $\Nat_+^*$ to the finite subsets of $Q\times V(s)\times (\Nat^*\cup\{\#\})$ is an \emph{approximation} of $\lk_s$ if it is obtained by a finite number of applications of requirements~(1) and~(2) of the definition of $\lk_s$, starting with $\lambda(v)=\emptyset$ for all $v\in\Nat_+^*$ (and so $\lambda(v)\subseteq\lk_s(v)$ for all $v$). It is straightforward to show by induction on the number of such applications that for every $v\in\Nat_+^*$, either $\lambda(v)=\emptyset$, or $\lambda(v)=\{(q,u,z)\}$ with $z\in\Nat_+^*-\{\varepsilon\}$, or $\lambda(v)=\{(q_1,u_1,\#),\dots,(q_n,u_n,\#),(q_n,u_n,\varepsilon)\}$ with the condition stated in the lemma, or $\lambda(v)=\{(q_1,u_1,\#),\dots,(q_n,u_n,\#)\}$ with that same condition. In the last case, Lemma~\ref{lm:linkdef}(4) implies that $\rhs(q_n,s,u_n)$ is defined, and hence $\lambda(v)$ is properly included in $\lk_s(v)$ by the definition of $\lk_s$. The first case does not occur when $v\in V(M(s))$, by Lemma~\ref{lm:linkdef}(3). Since $\lk_s$ is itself an approximation of $\lk_s$, this proves the lemma. \qed \end{proof} Lemma~\ref{lm:linkform} shows that for every $v\in V(M(s))$ there is exactly one triple $(q,u,z)$ in $\lk_s(v)$ with $z\in\Nat_+^*$. Thus, for $s\in\T_\Sigma$ and $v\in V(M(s))$, we define $\orig_s(v)\in Q\times V(s)\times\Nat_+^*$, called the \emph{origin} of $v$, by $\orig_s(v)=(q,u,z)$ if $(q,u,z)\in\lk_s(v)$. We denote $u$ also by $\orignode_s(v)$ and call it the \emph{origin node} of $v$. In the remainder of this section we state some elementary properties of links and origins. \begin{lemma}\label{lm:orig_prop1} \mbox{} \begin{enumerate} \item[(1)] If $(q,u,\#)\in\lk_s(v)$, then $M(s)/v=q_M(s/u)$. \item[(2)] If $\orig_s(v)=(q,u,z)$, then $z\in V_\Delta(\zeta)$ and \[ M(s)/v=\zeta/z[\bar{q}(x_i)\leftarrow \bar{q}_M(s/ui)\mid\bar{q}\in Q,\,i\in\Nat_+] \] where $\zeta=\rhs(q,s,u)$. \end{enumerate} \end{lemma} \begin{proof} If $(q,u,\#)\in\lk_s(v)$ then, by Lemma~\ref{lm:linkdef}(1), there is a reachable sentential form $\xi$ for $s$ such that $\xi/v=q(u)$. Since $\xi \Rightarrow_s^*M(s)$, we obtain that $\xi/v \Rightarrow_s^*M(s)/v$ and so $M(s)/v=q_M(s/u)$ by Lemma~\ref{lm:elem}(2). If $\orig_s(v)=(q,u,z)$ then, by Lemma~\ref{lm:linkdef}(2), there are $\hat{v}\in\Nat^*$ and a reachable sentential form $\xi$ for $s$ such that $v=\hat{v}z$, $\,\xi/\hat{v}=q(u)$, and $z\in V_\Delta(\zeta)$. Hence $\xi \Rightarrow_s \xi'\Rightarrow_s^*M(s)$, where $\xi'=\xi[\hat{v}\leftarrow \zeta[x_i\leftarrow ui\mid i\in\Nat_+]]$, and $\xi'/v=\zeta/z[x_i\leftarrow ui\mid i\in\Nat_+]\Rightarrow_s^*M(s)/v$. Then Lemma~\ref{lm:elem}(2) implies that $M(s)/v=\zeta/z[x_i\leftarrow ui\mid i\in\Nat_+] [\bar{q}(ui)\leftarrow \bar{q}_M(s/ui)\mid \bar{q}\in Q, i\in\Nat_+]$, which proves the statement. \qed \end{proof} The following four lemmas state relationships between the $\lk$ sets and the ancestor relation. Recall that (since $M$ is initialized) $\maxrhs(M)$ is the maximal height of the right-hand sides of rules of $M$. \begin{lemma}\label{lm:link_anc} Let $(q_1,u_1,z_1)\in\lk_s(v_1)$ and $(q_2,u_2,z_2)\in\lk_s(v_2)$, and let $v_1$ be a proper ancestor of $v_2$. Then: \begin{enumerate} \item[(1)] $u_1$ is an ancestor of $u_2$, and \item[(2)] if $u_1=u_2$, then $z_1,z_2\in\Nat_+^*$ and $|v_2|-|v_1|\leq\maxrhs(M)$. \end{enumerate} \end{lemma} \begin{proof} It is immediate from the definition of $\lk_s$ that if $v_1$ is a proper ancestor of $v_2$, then $u_1$ is an ancestor of $u_2$. Moreover, if $u_1=u_2=u$, then $q_1=q_2=q$, $z_1,z_2\in V_\Delta(\rhs(q,s,u))$, and there exists $v\in\Nat_+^*$ such that $v_1=vz_1$ and $v_2=vz_2$; hence $|v_2|-|v_1|=|z_2|-|z_1|\leq\height(\rhs(q,s,u))\leq\maxrhs(M)$. \qed \end{proof} The next lemma is an immediate corollary of Lemma~\ref{lm:link_anc}. \begin{lemma}\label{lm:orig_anc} Let $s\in\T_\Sigma$ and $v,\hat{v}\in V(M(s))$. \begin{enumerate} \item[(1)] If $\hat{v}$ is an ancestor of $v$, then $\orignode_s(\hat{v})$ is an ancestor of $\orignode_s(v)$. \item[(2)] If $\hat{v}$ is an ancestor of $v$ and $\orignode_s(\hat{v})=\orignode_s(v)$, then $|v|-|\hat{v}|\leq\maxrhs(M)$. \end{enumerate} \end{lemma} \begin{lemma}\label{lm:orig_prop2} If $\orig_s(v)=(q,u,z)$, then there exists $\hat{v}\in\Nat_+^*$ such that $v=\hat{v}z$ and $(q,u,\#)\in\lk_s(\hat{v})$. \end{lemma} \begin{proof} Immediate by (2) and (1) of Lemma~\ref{lm:linkdef}. \qed \end{proof} \begin{lemma}\label{lm:anc_preserve} If $(q,u,z)\in\lk_s(v)$ and $\hat{u}$ is an ancestor of $u$, then there exist an ancestor $\hat{v}$ of $v$ and a state $q'\in Q$ such that $(q',\hat{u},\#)\in\lk_s(\hat{v})$. \end{lemma} \begin{proof} Straightforward by recursion induction on the definition of $\lk_s$, as follows. For $(q_{0,\delta(s)},\varepsilon,\#)\in\lk_s(\varepsilon)$ we have $\hat{u}=\varepsilon$ and we take $\hat{v}=\varepsilon$ and $q'=q_{0,\delta(s)}$. Assume now that the statement holds for $(q,u,\#)\in\lk_s(v)$. For $(q,u,z)\in\lk_s(vz)$ and an ancestor $\hat{u}$ of $u$ we obtain by induction that $(q',\hat{u},\#)\in\lk_s(\hat{v})$ for an ancestor $\hat{v}$ of $v$, which is also an ancestor of $vz$. For $(\bar{q},ui,\#)\in\lk_s(vz)$ and an ancestor $\hat{u}$ of $u$ (and hence of $ui$) the previous argument also holds; for the ancestor $\hat{u}=ui$ of $ui$ we take $\hat{v}=vz$ and $q'=\bar{q}$. \qed \end{proof} Intuitively, a fact such as $(q,u,\#)\in\lk_s(v)$ does not depend on the whole of $s$, but only on the proper ancestors of $u$ and their children. This is proved in the next lemma. Let $s,s'\in\T_\Sigma$ and $u\in V(s)\cap V(s')$. We will say that $u$ is \emph{similar in $s$ and $s'$} if \begin{enumerate} \item[$(1)$] $\lab(s,u)=\lab(s',u)$, and \item[$(2)$] $\delta(s/ui)=\delta(s'/ui)$ for every child $ui$ of $u$. \end{enumerate} This implies that $\rhs(q,s,u)=\rhs(q,s',u)$ for every $q\in Q$. \newpage \begin{lemma}\label{lm:link_preserve} Let $s,s'\in\T_\Sigma$ be such that $\delta(s)=\delta(s')$, and let $u\in V(s)\cap V(s')$ be such that every proper ancestor of $u$ is similar in $s$ and $s'$. Let $q\in Q$ and $v,z\in\Nat_+^*$. \begin{enumerate} \item[$(1)$] If $(q,u,\#)\in\lk_s(v)$ then $(q,u,\#)\in\lk_{s'}(v)$. \item[$(2)$] Let, moreover, $u$ be similar in $s$ and $s'$. \\ If $(q,u,z)\in\lk_s(v)$ then $(q,u,z)\in\lk_{s'}(v)$. \end{enumerate} \end{lemma} \begin{proof} We prove~(1) by induction on $|u|$. By the definition of $\lk_s$, if $(q,\varepsilon,\#)\in\lk_s(v)$, then $q=q_{0,\delta(s)}$ and $v=\varepsilon$, and so $(q,\varepsilon,\#)\in\lk_{s'}(v)$ by the definition of $\lk_{s'}$. For the induction step we consider, by the definition of $\lk_s$, that $(\bar{q},ui,\#)\in\lk_s(vz)$ where $(q,u,\#)\in\lk_s(v)$, $\rhs(q,s,u)=\zeta$ and $\zeta/z=\bar{q}(x_i)$. By induction, $(q,u,\#)\in\lk_{s'}(v)$. Since $u$ is similar in $s$ and $s'$, $\rhs(q,s',u)=\zeta$. Hence $(\bar{q},ui,\#)\in\lk_{s'}(vz)$ by the definition of $\lk_{s'}$. To prove~(2) we consider, by the definition of $\lk_s$, that $(q,u,z)\in\lk_s(vz)$ where $(q,u,\#)\in\lk_s(v)$, $\rhs(q,s,u)=\zeta$, and $z\in V_\Delta(\zeta)$. By~(1), $(q,u,\#)\in\lk_{s'}(v)$, and since $u$ is similar in $s$ and $s'$ by assumption, $\rhs(q,s',u)=\zeta$. Hence $(q,u,z)\in\lk_{s'}(vz)$ by the definition of $\lk_{s'}$. \qed \end{proof} Note that in this lemma, by reasons of symmetry, the implications are actually equivalences. In the next section we wish to change (in fact, pump) parts of the input tree $s$ in such a way that a given node $v$ of $M(s)$ is preserved, together with the labels of all its ancestors (cf. Lemma~\ref{lm:diff_char}). Intuitively, this can be done as long as we do not change the following in $s$: the origin node $u$ of $v$, the labels of all ancestors of $u$, and for each ancestor of $u$ the look-ahead states at its children. This is proved in the next lemma. \begin{lemma}\label{lm:orig_preserve} Let $s\in\T_\Sigma$ and $v\in V(M(s))$, and let $\orignode_s(v)=u\in V(s)$. Moreover, let $s'\in\T_\Sigma$ be such that $u\in V(s')$ and every ancestor of $u$ (including $u$ itself) is similar in $s$ and $s'$. Then $v\in V(M(s'))$ and $\lab(M(s'),\hat{v})=\lab(M(s),\hat{v})$ for every ancestor $\hat{v}$ of $v$ (including $v$ itself). \end{lemma} \begin{proof} Since, by Lemma~\ref{lm:orig_anc}(1), $\orignode_s(\hat{v})$ is an ancestor of $u$, it suffices to prove the lemma for $\hat{v}=v$. Let $\orig_s(v)=(q,u,z)$. Since the root $\varepsilon$ is similar in $s$ and $s'$, $\delta(s)=\delta(s')$. It now follows from Lemma~\ref{lm:link_preserve}(2) that $v\in V(M(s'))$ and $\orig_{s'}(v)=(q,u,z)=\orig_s(v)$. Using this, and the fact that $\rhs(q,s',u)=\rhs(q,s,u)$ (because $u$ is similar in $s$ and $s'$), we obtain from Lemma~\ref{lm:orig_prop1}(2) that $\lab(M(s'),v)=\lab(M(s),v)$. \qed \end{proof} \section{Normal Forms}\label{sec:normform} In this section we prove normal forms for total dtlas. For each of these normal forms we consider its effect on $\maxdiff(M)$. We start with a simple normal form in which each axiom consists of one state, more precisely, is in $Q(\{x_0\})$. A dtla $M$ is \emph{initialized} if for every $p\in P$ there is a state $q_{0,p}$ such that $A(p)=q_{0,p}(x_0)$. The states $q_{0,p}$ are called initial states; they are not necessarily distinct. Note that for an initialized dtla, $M(s)=q_M(s)$ where $q=q_{0,p}$ and $p=\delta(s)$, for every $s\in\T_\Sigma$. Recall that $\maxrhs(M)$ is the maximal height of the axioms and the right-hand sides of the rules of $M$. \begin{lemma}\label{lm:make_initialized} For every total dtla $M$ an equivalent initialized dtla $M'$ can be constructed, with the same look-ahead automaton as $M$, such that \\ $|Q_{M'}|=|Q_M|+1$, $\;\maxrhs(M')\leq 2\cdot\maxrhs(M)$ and $$\maxdiff(M')\leq\maxdiff(M)\leq\Max\{\maxdiff(M'),\maxrhs(M)\}.$$ If $M$ is ultralinear or b-erasing, then so is $M'$. \end{lemma} \begin{proof} To construct $M'$ from $M$, we introduce a new state $q_0$. For every $a\in\Sigma^{(k)}$ and $p_1,\dots,p_k\in P$ we add the rule $$q_0(a(x_1\ot p_1,\dots,x_k\ot p_k))\to A(p)[q(x_0)\leftarrow\rhs(q,a,p_1,\dots,p_k)\mid q\in Q],\footnote{Note that the right-hand side of this rule is defined: since every look-ahead state is reachable, there exist trees $s_i$ with $\delta(s_i)=p_i$; then $\delta(a(s_1,\dots,s_k))=p$ and since $M(a(s_1,\dots,s_k))$ is defined because $M$ is total, $\rhs(q,a,p_1,\dots,p_k)$ is defined for every $q$ that occurs in $A(p)$. }$$ where $p=\delta(a,p_1,\dots,p_k)$. Finally, we change $A(p)$ into $q_0(x_0)$ for every $p\in P$. Then $M'$ is initialized, with $q_{0,p}=q_0$ for all $p\in P$. It should be clear from Lemma~\ref{lm:elem}(3/4) that $M'$ is equivalent to $M$. It should also be clear that for every $C\in\C_\Sigma$ and $p\in P$, if $C\neq\bot$ then $M'(C[p])=M(C[p])$. Moreover, for $C=\bot$ we have $M'(p)= \angl{q_0,p}$ and $M(p)=A(p)[q(x_0)\leftarrow\angl{q,p}\mid q\in Q]$. Since $\height(M'(p)/\varepsilon)=0$ and $\height(M(p)/v)\leq \height(A(p))\leq \maxrhs(M)$ for every $v\in V(M(p))$, this implies the required inequalities for $\maxdiff(M)$ and $\maxdiff(M')$. If $M$ is ultralinear with mapping $\mu$, then we can assume that $\mu(q)>0$ for all $q$, and then $M'$ is ultralinear by extending $\mu$ with $\mu(q_0)=0$. If $M$ is b-erasing then so is $M'$, because a cycle in $E_{M'}$ does not contain $q_0$ and hence is also a cycle in $E_M$. \qed \end{proof} Note that it follows from the inequalities for $\maxdiff(M)$ and $\maxdiff(M')$ that if $h(M')$ is a difference bound for $M'$, then $\Max\{h(M'),\maxrhs(M)\}$ is a difference bound for $M$. In fact, if $\diff(M)$ is finite, then $\diff(M')$ is finite because $\maxdiff(M')\leq\maxdiff(M)$, hence $\maxdiff(M')\leq h(M')$, from which it follows that $\maxdiff(M)\leq\Max\{\maxdiff(M'),\maxrhs(M)\}\leq\Max\{h(M'),\maxrhs(M)\}$. We continue with a basic and technically convenient normal form in which every state of the dtla only translates input trees that have the same look-ahead state; moreover, the rules satisfy a generalized completeness condition. A dtla $M$ is \emph{look-ahead uniform} (for short, \emph{la-uniform})\footnote{This notion is closely related to the notion of a \emph{uniform} i-transducer in~\cite{DBLP:journals/jcss/EngelfrietMS09}.} if there is a mapping $\rho: Q\to P$ (called \emph{la-map}) satisfying the following conditions, for $p\in P$ and $q,\bar{q}\in Q$: \begin{enumerate} \item[(1)] If $q(x_0)$ occurs in $A(p)$, then $\rho(q)=p$. \item[(2)] For every rule $q(a(x_1\ot p_1,\ldots,x_k\ot p_k))\to \zeta$ in $R$, \begin{enumerate} \item[(a)] $\rho(q)=\delta(a,p_1,\ldots,p_k)$ and \item[(b)] if $\bar{q}(x_i)$ occurs in $\zeta$, then $\rho(\bar{q})=p_i$. \end{enumerate} \item[(3)] For every $q\in Q$, $a\in \Sigma^{(k)}$ and $p_1,\dots,p_k\in P$ such that $\delta(a,p_1,\dots,p_k) = \rho(q)$, \\ there is a rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta$ in $R$. \end{enumerate} Note that a dtop is la-uniform if and only if it is complete (if and only if it is total, by Lemma~\ref{lm:reach}). Clearly, the dtla $M$ of Example~\ref{ex:Mex} is la-uniform with $\rho(q)=p_b$, and similarly, the one of Example~\ref{ex:Mthree} is la-uniform with $\rho(q)=p_a$. We will need the following obvious properties of an la-uniform dtla. \begin{lemma}\label{lm:prop_la-uniform} Let $M$ be an la-uniform dtla with la-map $\rho$. \begin{enumerate} \item[(1)] $\dom(\sem{q}_M)=\sem{\rho(q)}_M$ for every $q\in Q$. \item[(2)] $M$ is total. \item[(3)] $M^\circ$ is la-uniform with the same la-map $\rho$ as $M$. \item[(4)] Let $\xi$ be a reachable sentential form for $s\in\T_\Sigma$. \\ For all $v\in V(\xi)$, $q\in Q$ and $u\in V(s)$, if $\xi/v=q(u)$ then $\rho(q)=\delta(s/u)$. \end{enumerate} \end{lemma} \begin{proof} (1) We prove by structural induction on $s\in\T_\Sigma$ that $q_M(s)$ is defined if and only if $\delta(s)=\rho(q)$. Let $s=a(s_1,\dots,s_k)$ and $\delta(s_i)=p_i$ for $i\in[k]$. Then $\delta(s)=\delta(a,p_1,\dots,p_k)$. Thus, by conditions~(2)(a) and~(3) above, $\delta(s)=\rho(q)$ if and only if there is a rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta$ in $R$. For such a rule, by condition (2)(b) above, if $\bar{q}(x_i)$ occurs in $\zeta$, then $\rho(\bar{q})=p_i$ and hence, by induction, $\bar{q}_M(s_i)$ is defined. It now follows from Lemma~\ref{lm:elem}(4) that $\delta(s)=\rho(q)$ if and only if $q_M(s)$ is defined. (2) This is immediate from (1) and Lemma~\ref{lm:elem}(3), by condition~(1) above. (3) By (1), the set $R^\circ$ of rules of $M^\circ$ is obtained from $R$ by adding all the rules $q(p)\to \angl{q,p}$ such that $\rho(q)=p$. Hence $\rho$ also satisfies conditions~(2) and~(3) above for $M^\circ$ (and condition~(1) above, because $M^\circ$ has the same axioms as $M$). (4) The easy proof is by induction on the length of the computation $A(\delta(s))[x_0\leftarrow\varepsilon] \Rightarrow_s^* \xi$, using conditions~(1) and~(2) above. \qed \end{proof} Note that by (3) of this lemma, for every $C\in\C_\Sigma$ and $p\in P$, the tree $M(C[p])$ is in $\T_\Delta(Q_p\times\{p\})$ where $Q_p=\{q\in Q\mid \rho(q)=p\}$. We now prove that la-uniformity is a normal form for total dtlas. \begin{lemma}\label{lm:make_uniform} For every total dtla $M$ an equivalent la-uniform dtla $M'$ can be constructed, with the same look-ahead automaton as $M$, such that $|Q_{M'}|=|Q_M|\cdot|P_M|$, $\maxrhs(M')=\maxrhs(M)$ and $\maxdiff(M')=\maxdiff(M)$. If $M$ is initialized, ultralinear or b-erasing, then so is $M'$. \end{lemma} \begin{proof} We observe that it may be assumed that $M$ is complete: if $\rhs(q,a,p_1,\dots,p_k)$ is undefined, then we add the (dummy) rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to d$ where $d$ is any element of $\Delta^{(0)}$. We construct $M'$ as follows. The state set of $M'$ is $Q_{M'}=Q\times P$. Every axiom $A(p)$ of $M$ is changed into $A(p)[q(x_0)\leftarrow\angl{q,p}(x_0)\mid q\in Q]$, and every rule $$q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to\zeta$$ is changed into the rule $$\angl{q,p}(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta [\bar{q}(x_i)\leftarrow \angl{\bar{q},p_i}(x_i)\mid\bar{q}\in Q,\,i\in[k]]$$ where $p=\delta(a,p_1,\dots,p_k)$. It should be clear that $M'$ satisfies conditions~(1) and~(2) of the definition of la-uniformity with la-map $\rho$ such that $\rho(\angl{q,p})=p$; since $M$ is complete, condition~(3) is also satisfied. Hence $M'$ is la-uniform. Since $M$ and $M'$ are total, so are $M^\circ$ and $(M')^\circ$ (see the proof of Lemma~\ref{lm:MM}). Obviously, for every computation of $(M')^\circ$ on an input tree $\bar{s}\in \T_\Sigma(P)$, one obtains a computation of $M^\circ$ on that input tree by changing every $\angl{q,p}(u)$ that occurs in a sentential form into $q(u)$, and every $\angl{\angl{q,p},p}$ into $\angl{q,p}$. Hence $M'(\bar{s})=M(\bar{s})[\angl{q,p}\leftarrow \angl{\angl{q,p},p}\mid q\in Q, p\in P]$. This implies that $M'$ is equivalent to $M$. It also implies that $\maxdiff(M')=\maxdiff(M)$, as can easily be verified. Obviously, if $M$ is ultralinear with mapping $\mu_M: Q\to\Nat$, then so is $M'$ with the mapping $\mu$ such that $\mu(\angl{q,p})=\mu_M(q)$. Moreover, if there is an edge from $\angl{q,p}$ to $\angl{q',p_j}$ in $E_{M'}$, then there is an edge from $q$ to $q'$ in $E_M$. Hence, if $M$ is b-erasing, then so is $M'$. \qed \end{proof} Note that since $\maxdiff(M')=\maxdiff(M)$, the la-uniform dtla $M'$ has the same difference bounds as $M$. \begin{example}\label{ex:la-uniform} The dtla $M$ of Example~\ref{ex:Mleaves} is not la-uniform. We change it into an la-uniform dtla by the construction in the proof of Lemma~\ref{lm:make_uniform} (but keep calling it $M$). Then it has set of states $Q=\{q_{yz}\mid y,z\in\{a,b\}\}$ where $q_{yz}$ abbreviates $\angl{q,p_{yz}}$, so $\rho(q_{yz})=p_{yz}$. Its axioms are $A(p_{yz})=q_{yz}(x_0)$, and its rules are $q_{yz}(yz)\to yz$ and $$q_{wz}(\sigma(x_1\ot p_{wx},x_2\ot p_{yz}))\to \sigma(q_{wx}(x_1),q_{yz}(x_2),\#(w,z))$$ for all $w,x,y,z\in\{a,b\}$. \qed \end{example} From now on we mainly consider la-uniform dtlas. For an la-uniform dtla $M$, its la-map will be denoted $\rho$ (or $\rho_M$ when necessary). \smallskip Finally we generalize the normal form for dtops in~\cite{DBLP:journals/jcss/EngelfrietMS09} to total dtlas. For this normal form it is essential that dtlas need not be initialized, i.e., that arbitrary axioms are allowed. A dtla $M$ is \emph{earliest} if it is la-uniform and, for every state $q$ of $M$, the set $$\rlabs_M(q):=\{\lab(q_M(s),\varepsilon)\mid s\in\dom(\sem{q}_M)\}\subseteq\Delta$$ is not a singleton.\footnote{This is equivalent with requiring that $\sqcap\{q_M(s)\mid s\in\dom(\sem{q}_M)\}=\bot$, cf. the definition of earliest in~\cite{DBLP:journals/jcss/EngelfrietMS09}.} In other words, $M$ is \emph{not} earliest if it has a state $q$ for which the roots of all output trees $q_M(s)$ have the same label; intuitively, the node with that label could be produced earlier in the computation of $M$. A dtla $M$ is \emph{canonical} if it is earliest and $\sem{q}_M\neq\sem{q'}_M$ for all distinct states $q,q'$ of $M$. Since it is required that $M$ is la-uniform, the earliest and canonical properties are appropriately relativized with respect to each look-ahead state, see Lemma~\ref{lm:prop_la-uniform}(1). It is easy to see that the dtla $M$ of Example~\ref{ex:la-uniform} (which is the la-uniform version of the dtla of Example~\ref{ex:Mleaves}) is canonical: for all $y,z\in\{a,b\}$, $\rlabs_M(q_{yz})=\{yz,\sigma\}$ and $\sem{q_{yz}}_M$ is the restriction of $\sem{M}$ to $\sem{p_{yz}}_M$. For an la-uniform dtla $M$ the sets $\rlabs_M(q)$ can be computed in a standard way. In fact, consider the directed graph with set of nodes $Q\cup\Delta$ and with the following edges: for every rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to\zeta$ of $M$, if $\lab(\zeta,\varepsilon)=d\in\Delta$ then there is an edge $q\to d$, and if $\zeta=q'(x_j)$ then there is an edge $q\to q'$.\footnote{Note that the subgraph induced by $Q$ is the graph $E_M$, as in the definition of a b-erasing dtla in Section~\ref{sec:dtop}.} It is straightforward to show that $\rlabs_M(q)=\{d\in\Delta\mid q\to^* d\}$, as follows: \smallskip $(\subseteq)$ Structural induction on $s=a(s_1,\dots,s_k)$, such that $q_M(s)$ has root label $d$. By Lemma~\ref{lm:elem}(4), $q_M(s)=\zeta[\bar{q}(x_i)\leftarrow \bar{q}_M(s_i)\mid \bar{q}\in Q,\,i\in[k]]$ where $\zeta=\rhs(q,s,\varepsilon)$. If $\zeta$ has root label $d\in\Delta$, then there is an edge $q\to d$. If $\zeta=q'(x_j)$, then $q_M(s)=q'_M(s_j)$ and so $q\to q'\to^*d$ by induction. $(\supseteq)$ Induction on the length of $q\to^* d$. If $q\to d$ then $q_M(a(s_1,\dots,s_k))=\zeta[\bar{q}(x_i)\leftarrow \bar{q}_M(s_i)\mid \bar{q}\in Q,\,i\in[k]]$ by Lemma~\ref{lm:elem}(4), for any $s_i\in\sem{p_i}$, and so $q_M(a(s_1,\dots,s_k))$ has root label $d$. We have used that $M$ is la-uniform: if $\bar{q}(x_i)$ occurs in $\zeta$, then $\rho(\bar{q})=p_i$ and hence $\bar{q}_M(s_i)$ is defined by Lemma~\ref{lm:prop_la-uniform}(1). If $q\to q'\to^*d$ then $q_M(a(s_1,\dots,s_k))=q'_M(s_j)$ has root label $d$, because by induction there exists $s_j$ such that $q'_M(s_j)$ has root label $d$. \smallskip We now prove that canonicalness is a normal form for total dtlas. For an la-uniform dtla $M$, let $\fix(M)$ be a fixed subset of $\T_\Sigma$ such that for every $p\in P$ there is a unique $s\in\fix(M)$ with $\delta(s)=p$. Thus, $\fix(M)$ is a set of representatives of the equivalence classes $\sem{p}$, $p\in P$. Since every $\sem{p}$ is a regular tree language, a particular $\fix(M)$ can be computed from $M$. For every $p\in P$, let $s_p$ be the unique tree in $\fix(M)$ with $\delta(s_p)=p$. We define $$\sumfix(M)=\sum_{q\in Q}\size(q_M(s_{\rho(q)})).$$ Note that $\sumfix(M)$ is in $\Nat$ and can be computed from $M$. \begin{theorem}\label{th:uniform_earliest} For every la-uniform dtla $M$ an equivalent canonical dtla $\can(M)$ can be constructed, with the same look-ahead automaton as $M$, such that $$\maxdiff(M)-\sumfix(M)\leq\maxdiff(\can(M))\leq\maxdiff(M)+\sumfix(M).$$ \end{theorem} \begin{proof} We first prove the statement of this theorem for the case where $M$ is earliest. Since the equivalence of two dtlas is decidable (see~\cite{DBLP:journals/actaC/Esik81} and~\cite[Corollary~19]{DBLP:journals/jcss/EngelfrietMS09}), it is decidable for two states $q,q'$ of $M$ whether or not $\sem{q}_M=\sem{q'}_M$. If this holds, then $q'$ can be replaced by $q$ in every axiom and every right-hand side of a rule, thus making $q'$ unreachable and hence superfluous. Since in $M(C[p])$ every $\angl{q',p}$ is replaced by $\angl{q,p}$, $\maxdiff(M)$ does not change. Thus, repeating this procedure one obtains a canonical dtla $\can(M)$ equivalent to $M$, with $\maxdiff(\can(M))=\maxdiff(M)$. For the interested reader we observe that for an earliest dtla $M$ the equivalence relation $\equiv$ on $Q$ defined by $q\equiv q'$ if and only if $\sem{q}_M=\sem{q'}_M$, can in fact easily be computed by fixpoint iteration, because it is the largest equivalence relation on $Q$ such that if $q\equiv q'$ then (a) $\rho(q)=\rho(q')$ and (b) if $\rhs(q,a,p_1,\dots,p_k)= t[q_1(x_{i_1}),\dots,q_r(x_{i_r})]$ where $t\in\mathcal{P}_\Sigma$ and $r=|V_\bot(t)|$, then $\rhs(q',a,p_1,\dots,p_k)= t[q'_1(x_{i_1}),\dots,q'_r(x_{i_r})]$ with $q_j\equiv q'_j$ for every $j\in[r]$. The straightforward proof of this is left to the reader, cf. the proof of~\cite[Theorem 13]{DBLP:journals/jcss/EngelfrietMS09}. Thus, the full dtla equivalence test of \cite{DBLP:journals/actaC/Esik81,DBLP:journals/jcss/EngelfrietMS09} is not needed. It remains to be proved that every la-uniform dtla $M$ can be transformed into an equivalent earliest dtla~$M'$, with the same look-ahead automaton, such that the distance between $\maxdiff(M')$ and $\maxdiff(M)$ is at most $\sumfix(M)$. If $M$ is not earliest, then we obtain $M'$ by repeatedly applying the following transformation step. \emph{Transformation}. We transform $M$ into a dtla $N$ with the same look-ahead automaton. Let $Q_1$ be the (nonempty) set of states $q\in Q$ such that $\rlabs_M(q)$ is a singleton, and for every $q\in Q_1$ let $\rlabs_M(q)=\{d_q\}$ and $m_q=\rk(d_q)$. The set of states of $N$ is $$Q_N:=(Q-Q_1)\cup \{\angl{q,i}\mid q\in Q_1, i\in[m_q]\}.$$ When $M$ arrives in state $q\in Q_1$ at a node $u$ of an input tree $s$, $N$ will first output the symbol $d_q$ and then arrive at node $u$ in the states $\angl{q,1},\dots,\angl{q,m_q}$, to compute the direct subtrees of the tree $q_M(s/u)$, where $\angl{q,i}$ computes the $i$th direct subtree $q_M(s/u)/i$. So, to describe $N$, we define for every tree $\zeta\in\T_\Delta(Q(\Omega))$ where $\Omega$ is any set of symbols of rank~0, the tree $\zeta\Phi_\Omega=\zeta[q(\omega)\leftarrow d_q(\angl{q,1}(\omega),\dots,\angl{q,m_q}(\omega)) \mid q\in Q_1, \omega\in\Omega]$. For every $p\in P$, the $p$-axiom of $N$ is $A(p)\Phi_{\{x_0\}}$. Every rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta$ is changed into the rule $q(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta\Phi_{X_k}$ if $q\in Q-Q_1$, and into the $m_q$ rules $\angl{q,i}(a(x_1\ot p_1,\dots,x_k\ot p_k))\to \zeta\Phi_{X_k}/i$ if $q\in Q_1$. Note that in the latter case the root of $\zeta\Phi_{X_k}$ has label $d_q$ and so its $i$th direct subtree is well defined.\footnote{This is clear if $\lab(\zeta,\varepsilon)=d_q$. If $\zeta=q'(x_j)$, then $q'\in Q_1$ and $d_{q'}=d_q$; so $\zeta\Phi_{X_k}=d_q(\angl{q',1}(x_j),\dots,\angl{q',m_q}(x_j))$ and one obtains the rules $\angl{q,i}(a(x_1\ot p_1,\dots,x_k\ot p_k))\to\angl{q',i}(x_j)$ for $i\in[m_q]$.} If $M$ has la-map $\rho$, then $N$ is la-uniform with la-map $\rho_N$ such that $\rho_N(q)=\rho(q)$ for $q\in Q-Q_1$ and $\rho_N(\angl{q,i})=\rho(q)$ for $q\in Q_1$ and $i\in[m_q]$. It should be clear intuitively that $N$ is equivalent to $M$. Formally it can easily be shown for every $s\in\T_\Sigma$ and every reachable sentential form $\xi$ of $M$ for $s$, that $\xi\Phi_{V(s)}$ is a reachable sentential form of $N$ for $s$ (where each computation step of $M$ is simulated by one or $m_q$ computation steps of $N$), and hence $N(s)=M(s)$. We will compute $\maxdiff(N)$ below; to do that we need to extend the previous statement to trees $\bar{s}\in\T_\Sigma(P)$. Let $\Psi$ be the substitution $[\angl{q,p}\leftarrow d_q(\angl{\angl{q,1},p},\dots,\angl{\angl{q,m_q},p})\mid q\in Q_1,p\in P]$. Then, for every reachable sentential form $\xi$ of $M^\circ$ for $\bar{s}$, the tree $\xi\Phi_{V(s)}\Psi$ is a reachable sentential form of $N^\circ$ for $\bar{s}$, and hence $N(\bar{s})=M(\bar{s})\Psi$. \emph{Repetition}. Using Lemma~\ref{lm:elem}(3/4), it can easily be shown for every $q\in Q$ and $s\in\sem{\rho(q)}$, that $q_N(s)=q_M(s)$ if $q\notin Q_1$, and that $\angl{q,i}_N(s)=q_M(s)/i$ for every $i\in[m_q]$ if $q\in Q_1$. From this (and assuming that $\fix(N)=\fix(M)$), it should be clear that $\sumfix(N)<\sumfix(M)$, because, for $q\in Q_1$ and $s\in\sem{\rho(q)}$, $$\sum_{i\in[m_q]}\size(\angl{q,i}_N(s))=\sum_{i\in[m_q]}\size(q_M(s)/i)=\size(q_M(s))-1.$$ Hence, the repetition of the above transformation stops after at most $\sumfix(M)$ steps, with an earliest dtla $M'$ equivalent to $M$. \emph{Difference trees}. It remains to prove that the distance between $\maxdiff(N)$ and $\maxdiff(M)$ is at most~1, i.e., that $\maxdiff(N)\leq \maxdiff(M)+1$ and $\maxdiff(M)\leq \maxdiff(N)+1$. Consider $C\in\C_\Sigma$ and $p,p'\in P$ with $p\neq p'$. Recall from above that $N(C[p])=M(C[p])\Psi= M(C[p])[\angl{q,p}\leftarrow d_q(\angl{\angl{q,1},p},\dots,\angl{\angl{q,m_q},p})\mid q\in Q_1]$ and similarly for $p'$. We observe that if $v$ is a node of $M(C[p])$, then each proper ancestor of $v$ has the same label in $M(C[p])$ and $N(C[p])$, and similarly for $p'$. Let $v$ be a difference node of $N(C[p])$ and $N(C[p'])$ that is not a leaf of $N(C[p])$. Then $v\in V(M(C[p]))$. If also $v\in V(M(C[p']))$, then Lemma~\ref{lm:diff_char} implies that $v$ is also a difference node of $M(C[p])$ and $M(C[p'])$ (in fact, by the above observation every proper ancestor of $v$ has the same label in $M(C[p])$ and $M(C[p'])$; if $v$ would have the same label in $M(C[p])$ and $M(C[p'])$, then that label would be in $\Delta$ because $p\neq p'$, and hence $v$ would have the same label in $N(C[p])$ and $N(C[p'])$). Consequently, $\height(N(C[p])/v)=\height(M(C[p])\Psi/v)\leq \height(M(C[p])/v)+1\leq\maxdiff(M)+1$. If $v\notin V(M(C[p']))$, then the parent $\hat{v}$ of $v$ is a difference node of $M(C[p])$ and $M(C[p'])$ (in fact, the label of $\hat{v}$ in $M(C[p'])$ is $\angl{q,p'}$ for some $q\in Q_1$, and so $\hat{v}$ has different labels in $M(C[p])$ and $M(C[p'])$ because $p\neq p'$). Hence, in this case, $\height(N(C[p])/v)\leq \height(N(C[p])/\hat{v})\leq \height(M(C[p])/\hat{v})+1\leq\maxdiff(M)+1$. This proves that $\maxdiff(N)\leq \maxdiff(M)+1$. Now let $v$ be a difference node of $M(C[p])$ and $M(C[p'])$ that is not a leaf of $M(C[p])$; note that $\lab(N(C[p]),v)=\lab(M(C[p]),v)\in\Delta$. If $v$ is also a difference node of $N(C[p])$ and $N(C[p'])$, then $\height(M(C[p])/v)\leq\height(M(C[p])\Psi/v)=\height(N(C[p])/v)\leq\maxdiff(N)$. Now assume that $v$ is not a difference node of $N(C[p])$ and $N(C[p'])$. Then, by Lemma~\ref{lm:diff_char}, the label of $v$ in $M(C[p'])$ is $\angl{q,p'}$ for some $q\in Q_1$ and $\lab(N(C[p]),v)=\lab(N(C[p']),v)=d_q$. This implies, again by Lemma~\ref{lm:diff_char}, that the children of $v$ are difference nodes of $N(C[p])$ and $N(C[p'])$. Let $vi$ be a child of $v$ for which $\height(M(C[p])/vi)$ is maximal. Then we have $\height(M(C[p])/v)=\height(M(C[p])/vi)+1\leq\height(N(C[p])/vi)+1\leq\maxdiff(N)+1$. This proves that $\maxdiff(M)\leq \maxdiff(N)+1$. \qed \end{proof} Note that it follows from the inequalities for $\maxdiff(\can(M))$ that if $h(M)$ is a difference bound for $M$, then $h(M)+\sumfix(M)$ is a difference bound for $\can(M)$. In fact (similar to the argument after Lemma~\ref{lm:make_initialized}), if $\diff(\can(M))$ is finite, then $\diff(M)$ is finite because $\maxdiff(M)\leq\maxdiff(\can(M))+\sumfix(M)$, hence $\maxdiff(M)\leq h(M)$ and hence $\maxdiff(\can(M))\leq \maxdiff(M)+\sumfix(M)\leq h(M)+\sumfix(M)$. Note also that the transformation in the above proof does not preserve the ultralinear property (as can be seen in the next example). \begin{example} In this example we denote by $Y$ the nonempty subsets of $\{a,b\}$, i.e., $Y=\{\{a\},\{b\},\{a,b\}\}$. Let $\Sigma=\{\sigma^{(2)},a^{(0)},b^{(0)}\}$ and $\Delta=\{\sigma_y^{(2)}\mid y\in Y\}\cup\{a^{(0)},b^{(0)}\}$. We consider an la-uniform dtla $M$ such that $M(a)=a$, $M(b)=b$ and $M(\sigma(s_1,s_2))=\sigma_y(M(s_1),M(s_2))$ for $s_1,s_2\in\T_\Sigma$, where $y$ is the set of labels of the leaves of $\sigma(s_1,s_2)$. Its set of look-ahead states is $P=\{p_y\mid y\in Y\}$ and $\delta$ is defined in the obvious way: $\delta(a)=\{a\}$, $\delta(b)=\{b\}$ and $\delta(\sigma,p_y,p_z)=p_{y\cup z}$ for $y,z\in Y$. Its set of states is $Q=\{q_y\mid y\in Y\}$ with $\rho(q_y)=p_y$, its axioms are $A(p_y)=q_y(x_0)$ for every $y\in Y$, and its set $R$ consists of the rules $q_{\{a\}}(a)\to a$, $\;q_{\{b\}}(b)\to b$ and $$q_{y\cup z}(\sigma(x_1\ot p_y,x_2\ot p_z))\to \sigma_{y\cup z}(q_y(x_1),q_z(x_2))$$ for $y,z\in Y$. The dtla $M_1$ that is obtained from $M$ by identifying all its states into one state $q$, is equivalent to $M$; it has $\rlabs_{M_1}(q)=\Delta$, but it is not la-uniform. However, $M$ is not earliest: in fact, $\rlabs_M(q_{\{a\}})=\{\sigma_{\{a\}},a\}$ and $\rlabs_M(q_{\{b\}})=\{\sigma_{\{b\}},b\}$, but $\rlabs_M(q_{\{a,b\}})=\{\sigma_{\{a,b\}}\}$. Let $N$ be the dtla obtained from $M$ by applying the transformation in the proof of Theorem~\ref{th:uniform_earliest} once. Then $Q_1=\{q_{\{a,b\}}\}$. We will write the states $\angl{q_{\{a,b\}},1}$ and $\angl{q_{\{a,b\}},2}$ as $q_{1\{a,b\}}$ and $q_{2\{a,b\}}$, respectively. So, $N$ has states $q_{\{a\}}$, $q_{\{b\}}$, $q_{1\{a,b\}}$ and $q_{2\{a,b\}}$. Its axioms are $A_N(p_y)=q_y(x_0)$ for $y=\{a\}$ or $y=\{b\}$ (just as in $M$), and $A_N(p_y)=\sigma_y(q_{1y}(x_0),q_{2y}(x_0))$ for $y=\{a,b\}$. For $y=\{a\}$ or $y=\{b\}$, its set $R_N$ of rules contains the rule $$q_y(\sigma(x_1\ot p_y,x_2\ot p_y))\to \sigma_y(q_y(x_1),q_y(x_2))$$ plus the rules $q_{\{a\}}(a)\to a$ and $\;q_{\{b\}}(b)\to b$ (just as in $M$). Moreover, for $y=\{a,b\}$, $R_N$ contains the rules \[ \begin{array}{rll} q_{1y}(\sigma(x_1\ot p_y,x_2\ot p_z)) & \to & \sigma_y(q_{1y}(x_1),q_{2y}(x_1)) \\ q_{2y}(\sigma(x_1\ot p_z,x_2\ot p_y)) & \to & \sigma_y(q_{1y}(x_2),q_{2y}(x_2)) \end{array} \] for all $z\in Y$, plus the rules \[ \begin{array}{rll} q_{1y}(\sigma(x_1\ot p_w,x_2\ot p_z)) & \to & q_w(x_1) \\ q_{2y}(\sigma(x_1\ot p_z,x_2\ot p_w)) & \to & q_w(x_2) \end{array} \] for all $w,z\in Y$ with $w\neq y$ and $w\cup z=y$. For $N$ we have $\rlabs_N(q_{\{a\}})=\{\sigma_{\{a\}},a\}$ and $\rlabs_N(q_{\{b\}})=\{\sigma_{\{b\}},b\}$ as for $M$, and we have $\rlabs_N(q_{1\{a,b\}})=\rlabs_N(q_{2\{a,b\}})=\Delta$. Hence $N$ is earliest. Obviously, $N$ is also canonical, and so $N=\can(M)$. \qed \end{example} \section{Auxiliary Bounds} \label{sec:upper} If $M$ is an initialized la-uniform dtla, then so is $M^\circ$, cf. Lemma~\ref{lm:prop_la-uniform}(3). In this section and the next, the lemmas of the previous section will be applied to $M^\circ$ instead of $M$. \medskip In view of Theorem~\ref{th:alg} and Corollary~\ref{co:goal}, we wish to compute a difference bound for $M$, i.e., an upper bound for $\maxdiff(M)$ when $\diff(M)$ is finite. Thus, we are looking for an upper bound on the height of all difference trees of $M$, i.e., all trees $M(C[p])/v$ where $C$ is a $\Sigma$-context, $p\in P$, and $v$ is a difference node of $M(C[p])$ and $M(C[p'])$ where $p'\in P$. Let $u$ and $u'$ be the respective origin nodes of $v$, i.e., $u=\orignode_{C[p]}(v)$ and $u'=\orignode_{C[p']}(v)$. We first compute an upper bound for the case where $u$ is \emph{not} a proper ancestor of $u'$. The idea is that if the height of $M(C[p])/v$ is larger than that upper bound, then we can pump the subtree at one of $u$'s children (without changing the look-ahead state at that child), turning $C$ into $\bar{C}$, in such a way that $M(\bar{C}[p])/v$ becomes arbitrarily large. Since the pumping does not change the labels of the ancestors of $u$ and $u'$, nor the look-ahead states at the children of those ancestors, $v$ is still a difference node of $M(\bar{C}[p])$ and $M(\bar{C}[p'])$ by Lemmas~\ref{lm:diff_char} and~\ref{lm:orig_preserve}. To express an upper bound for the height of $M(C[p])/v$, we use an auxiliary bound defined as follows. \begin{definition}\label{df:output_bound} A number $h_{\rm o}(M)\in \Nat$ is an \emph{output bound} for $M$ if it has the following two properties, for every $q\in Q$ and $p,p',p_1,p'_1\in P$: \begin{enumerate} \item[(1)] if the set $\{q_M(s)\mid s\in\T_\Sigma,\,\delta(s)=p_1\}$ is finite, then $\height(t)\leq h_{\rm o}(M)$ for every tree~$t$ in this set, and \item[(2)] if the set $\{q_M(C[p])\mid C\in\C_\Sigma,\,\delta(C[p])=p_1,\,\delta(C[p'])=p'_1\}$ is finite, then $\height(t)\leq h_{\rm o}(M)$ for every tree $t$ in this set. \end{enumerate} \end{definition} Note that since $M$ is la-uniform, it suffices to consider the case where $p_1=\rho(q)$. The minimal output bound $h_{\rm o}(M)$ can be computed from $M$, because every set mentioned in Definition~\ref{df:output_bound} is the image of a regular tree language by a dtla translation (as will be shown in the proof of Theorem~\ref{th:output_bound}), and because it is decidable whether or not such an image is finite, and if so, the elements of that image can be computed.\footnote{See~\cite[Theorem~4.5]{DBLP:journals/iandc/DrewesE98}. Note that every dtla translation can be realized by a macro tree transducer.} Also, for completeness sake, we will prove by a pumping argument that $\maxrhs(M)\cdot|Q|\cdot(|P|+2)$ is an output bound for $M$, in Lemma~\ref{lm:range_bound} and Theorem~\ref{th:output_bound}. We now show that the upper bound discussed above is $\maxrhs(M)+h_{\rm o}(M)$. For later use (in the proof of Lemma~\ref{lm:new_prop_ancestor}) we prove a slightly more general result; the case discussed above is obtained by taking $\bar{v}=v$. Recall that $\maxrhs(M)$ is the maximal height of the right-hand sides of rules of $M$. \begin{lemma}\label{lm:new_not_ancestor} Let $h_{\rm o}(M)$ be an output bound for $M$. Let $C\in\C_\Sigma$ and $p,p'\in P$, let $v$ be a difference node of $M(C[p])$ and $M(C[p'])$, and let $\bar{v}$ be a descendant of $v$ in $M(C[p])$ such that $\orignode_{C[p]}(\bar{v})$ is not a proper ancestor of $\orignode_{C[p']}(v)$. If $\diff(M)$ is finite, then $$\height(M(C[p])/\bar{v})\leq \maxrhs(M)+h_{\rm o}(M).$$ \end{lemma} \begin{proof} Assume that $\height(M(C[p])/\bar{v})>\maxrhs(M)+h_{\rm o}(M)$. Let $u=\orignode_{C[p]}(v)$ and $u'=\orignode_{C[p']}(v)$, and let $(q,\bar{u},z)=\orig_{C[p]}(\bar{v})$. By Lemma~\ref{lm:orig_anc}(1), $u$ is an ancestor of $\bar{u}$. By Lemma~\ref{lm:orig_prop1}(2), $M(C[p])/\bar{v}=\zeta/z[\bar{q}(x_i)\leftarrow \bar{q}_M(C[p]/\bar{u}i)\mid\bar{q}\in Q,\,i\in\Nat_+]$ where $\zeta=\rhs(q,C[p],\bar{u})$. Since $\height(M(C[p])/\bar{v})>\maxrhs(M)+h_{\rm o}(M)$, there exist $y\in\Nat_+^*$, $q'\in Q$, and $j\in\Nat$ such that $\zeta/zy=q'(x_j)$ and $\height(q'_M(C[p]/\bar{u}j))> h_{\rm o}(M)$. Note that $M(C[p])/\bar{v}y=q'_M(C[p]/\bar{u}j)$. Note also that, by the definition of $\lk$, $(q',\bar{u}j,\#)\in\lk_{C[p]}(\bar{v}y)$. The situation is shown in Figure~\ref{fi:notanc} (but note that it is also possible that $u$ is an ancestor of $u'$, as in Figure~\ref{fi:anc}). \begin{figure} \centerline{ \input 1.pdf_t } \caption{\quad $\bar{u}=\orignode_{C[p]}(\bar{v})$ is not a proper ancestor of $u'=\orignode_{C[p']}(v)$ \label{fi:notanc}} \end{figure}% \noindent There are now two cases: $p$ does or does not occur in $C[p]/\bar{u}j$. Let $p_j=\delta(C[p]/\bar{u}j)$. \emph{Case 1}: $p$ does not occur in $C[p]/\bar{u}j$, i.e., $C[p]/\bar{u}j=C/\bar{u}j\in\T_\Sigma$.\footnote{In Figure~\ref{fi:notanc}, the vertical thick line represents the node of $C$ with label $\bot$. So, it shows Case~1.} By Definition~\ref{df:output_bound}(1) of $h_{\rm o}(M)$, the set $\{q'_M(s)\mid s\in\T_\Sigma,\,\delta(s)=p_j\}$ is infinite. Hence there exist trees $s_n\in\T_\Sigma$ such that $\delta(s_n)=p_j$ and $\height(q'_M(s_n))>n$ for every $n\in\Nat$. Let $C_n=C[\bar{u}j\leftarrow s_n]$.\footnote{Recall that this is the context $C$ in which the subtree at $\bar{u}j$ is replaced by $s_n$.} Since $\delta(s_n)=\delta(C[p]/\bar{u}j)=\delta(C[p']/\bar{u}j)$ and $\bar{u}$ is not a proper ancestor of $u$ or $u'$, it follows from Lemma~\ref{lm:orig_preserve} that $\lab(M(C_n[p]),\hat{v})=\lab(M(C[p]),\hat{v})$ for every ancestor $\hat{v}$ of $v$, and similarly for $p'$. Hence, by Lemma~\ref{lm:diff_char}, $v$ is a difference node of $M(C_n[p])$ and $M(C_n[p'])$, and so $M(C_n[p])/v$ is in $\diff(M)$. Also, since $(q',\bar{u}j,\#)\in\lk_{C[p]}(\bar{v}y)$, we obtain from Lemma~\ref{lm:link_preserve}(1) that $(q',\bar{u}j,\#)\in\lk_{C_n[p]}(\bar{v}y)$. Consequently, by Lemma~\ref{lm:orig_prop1}(1), $M(C_n[p])/\bar{v}y=q'_M(C_n[p]/\bar{u}j)=q'_M(s_n)$, which implies that $\height(M(C_n[p])/v)\geq\height(M(C_n[p])/\bar{v}y) >n$. Hence $\diff(M)$ is infinite. \emph{Case 2}: $p$ occurs in $C[p]/\bar{u}j$, i.e., $C[p]/\bar{u}j=D[p]$ where $D=C/\bar{u}j\in\C_\Sigma$. Let $p_j'=\delta(C[p']/\bar{u}j)$. By Definition~\ref{df:output_bound}(2) of $h_{\rm o}(M)$, the set $\{q'_M(D[p])\mid D\in\C_\Sigma, \,\delta(D[p])=p_j, \,\delta(D[p'])=p_j'\}$ is infinite. Hence there exist contexts $D_n$ such that $\delta(D_n[p])=p_j$, $\delta(D_n[p'])=p_j'$, and $\height(q'_M(D_n[p]))>n$. Let $C_n=C[\bar{u}j\leftarrow D_n]$. It can now be shown in the same way as in Case~1 that $M(C_n[p])/v$ is in $\diff(M)$ and $\height(M(C_n[p])/v)>n$, and hence $\diff(M)$ is infinite. Note that the condition $\delta(D_n[p'])=p_j'$ is needed to ensure that Lemma~\ref{lm:orig_preserve} is applicable to $u'$. \qed \end{proof} We now consider the case where $u$ \emph{is} a proper ancestor of $u'$, and again wish to obtain an upper bound for the height of $M(C[p])/v$. It follows from Lemma~\ref{lm:new_not_ancestor} that $\height(M(C[p])/\bar{v})\leq \maxrhs(M)+h_{\rm o}(M)$ for every descendant $\bar{v}$ of $v$ of which the origin node $\bar{u}$ is not a proper ancestor of $u'$. But what about a descendant $vw$ of $v$ of which the origin node $x$ \emph{is} a proper ancestor of $u'$?\footnote{See Figure~\ref{fi:anc} for $w=w_1$.} Then, by the previous observation, it suffices to obtain an upper bound on $|w|$. Now we observe that, roughly speaking, when $M$ arrives at node $x$ of $C[p]$ it generates node $vw$ of $M(C[p])$, which is a descendant of $v$, but when $M$ arrives at node $x$ of $C[p']$ it generates a node $\hat{v}$ of $M(C[p'])$ that is an ancestor of $v$. Hence, it suffices to have an upper bound on $|vw|-|\hat{v}|$, which measures how much the translation of $M(C[p])$ is ahead of the translation of $M(C[p'])$ when $M$ arrives at $x$. Note that by Lemma~\ref{lm:diff_char}, $\lab(C[p],\tilde{v})=\lab(C[p'],\tilde{v})$ for every proper ancestor $\tilde{v}$ of $\hat{v}$. Thus, to express an upper bound for the height of $M(C[p])/v$, we use an auxiliary bound defined as follows. \begin{definition}\label{df:ancestral_bound} A number $h_{\rm a}(M)\in\Nat$ is an \emph{ancestral bound} for $M$ if either $\diff(M)$ is infinite or the following holds for every $C\in\C_\Sigma$, $p,p'\in P$, $x\in V(C)$, $y\in V(M(C[p]))$, $y'\in V(M(C[p']))$, and $q,q'\in Q$: \medskip \noindent if \begin{enumerate} \item[(1)] $(q,x,\#)\in\lk_{C[p]}(y)$ and $\;(q',x,\#)\in\lk_{C[p']}(y')$, \item[(2)] $y'$ is an ancestor of $y$, and \item[(3)] $\lab(C[p],\hat{y}')=\lab(C[p'],\hat{y}')$ for every proper ancestor $\hat{y}'$ of $y'$, \end{enumerate} then $|y|-|y'|\leq h_{\rm a}(M)$. \end{definition} We will show in the next section (Lemma~\ref{lm:no_split}) that every dtla $M$ has an ancestral bound $h_{\rm a}(M)$. Unfortunately, we do not know whether an ancestral bound can be computed from $M$. We will also show in the next section (Theorem~\ref{th:ancestral_bound}) that it can be computed in the restricted case where $M$ is ultralinear and b-erasing. \begin{lemma}\label{lm:new_prop_ancestor} Let $h_{\rm o}(M)$ be an output bound and $h_{\rm a}(M)$ an ancestral bound for $M$. Let $C\in\C_\Sigma$ and $p,p'\in P$, and let $v$ be a difference node of $M(C[p])$ and $M(C[p'])$ such that $\orignode_{C[p]}(v)$ is a proper ancestor of $\orignode_{C[p']}(v)$. If $\diff(M)$ is finite, then $$\height(M(C[p])/v)\leq 2\cdot\maxrhs(M)+h_{\rm o}(M)+h_{\rm a}(M)+1.$$ \end{lemma} \begin{proof} Let $\orignode_{C[p]}(v)=u$ and $\orignode_{C[p']}(v)=u'$. Consider an arbitrary leaf $w$ of $M(C[p])/v$. We have to show that $|w|\leq 2\cdot\maxrhs(M)+h_{\rm o}(M)+h_{\rm a}(M)+1$. The proof is illustrated in Figure~\ref{fi:anc}. \begin{figure} \centerline{ \input 2.pdf_t } \caption{\quad $u=\orignode_{C[p]}(v)$ is a proper ancestor of $u'=\orignode_{C[p']}(v)$ \label{fi:anc}} \end{figure}% Let $w=w_1w_2$ where $w_1$ is the longest ancestor of $w$ such that $\orignode_{C[p]}(vw_1)$ is a proper ancestor of $u'$. If $w_2=jw_2'$ with $j\in\Nat_+$, then $\orignode_{C[p]}(vw_1j)$ is not a proper ancestor of $u'$, and so $|w_2|=|w_2'|+1\leq \maxrhs(M)+h_{\rm o}(M)+1$ by Lemma~\ref{lm:new_not_ancestor} (with $\bar{v}=vw_1j$). It now suffices to prove that $|w_1|\leq\maxrhs(M)+h_{\rm a}(M)$. Let $\orig_{C[p]}(vw_1)=(q,x,z)$, and note that $x$ is a descendant of $u$ by Lemma~\ref{lm:orig_anc}(1) (and a proper ancestor of $u'$ by definition of $w_1$). If $x=u$, then $|w_1|\leq\maxrhs(M)$ by Lemma~\ref{lm:orig_anc}(2). Otherwise, by Lemma~\ref{lm:orig_prop2} (applied to $vw_1$) and Lemma~\ref{lm:link_anc}(1), $w_1=w_1'z$ such that $(q,x,\#)\in\lk_{C[p]}(vw_1')$. Since $|z|\leq\maxrhs(M)$, it now remains to show that $|w_1'|\leq h_{\rm a}(M)$. By Lemma~\ref{lm:anc_preserve}, since $x$ is an ancestor of $u'$, there is an ancestor $\hat{v}$ of $v$ such that $(q',x,\#)\in\lk_{C[p']}(\hat{v})$ for some $q'\in Q$. We now apply Definition~\ref{df:ancestral_bound} of $h_{\rm a}(M)$ to the nodes $y=vw_1'$ of $M(C[p])$ and $y'=\hat{v}$ of $M(C[p'])$, and obtain that $|w_1'|\leq |vw_1'|-|\hat{v}|\leq h_{\rm a}(M)$. Note that condition~(3) in Definition~\ref{df:ancestral_bound} is satisfied by Lemma~\ref{lm:diff_char}, because $\hat{v}$ is an ancestor of the difference node $v$ of $M(C[p])$ and $M(C[p'])$. \qed \end{proof} \begin{theorem}\label{th:upper_bound} If $h_{\rm o}(M)$ is an output bound and $h_{\rm a}(M)$ an ancestral bound for $M$, then $h(M)=2\cdot\maxrhs(M)+h_{\rm o}(M)+h_{\rm a}(M)+1$ is a difference bound for $M$. \end{theorem} \begin{proof} Immediate from Lemmas~\ref{lm:new_not_ancestor} (for $\bar{v}=v$) and~\ref{lm:new_prop_ancestor}. \qed \end{proof} We end this section by proving that every dtla $M$ has a computable output bound. The reader who believes that this follows from \cite[Theorem~4.5]{DBLP:journals/iandc/DrewesE98} can skip the rest of this section. However, the proof of the next lemma can also serve as an introduction to the pumping technique used in the next section for computing an ancestral bound for $M$. \begin{lemma}\label{lm:range_bound} Let $p\in P$ such that the set $\{M(s)\mid s\in\sem{p}\}$ is finite. Then $\height(M(s))\leq \maxrhs(M)\cdot|Q|$ for every $s\in\sem{p}$. \end{lemma} \begin{proof} The (obvious) idea for the proof is that we consider an output path from the root to a leaf $v$ in $M(s)$ that is longer than $\maxrhs(M)\cdot|Q|$. Then we find two ancestors $\bar{u}$ and $\tilde{u}$ of the origin node $u$ of $v$ where the computation of $M$ on $s$ is in a cycle (i.e., arrives in the same state at $\bar{u}$ and $\tilde{u}$) and produces a nonempty part of the output path. And then we pump the part of $s$ between $\bar{u}$ and $\tilde{u}$, thus pumping the path to $v$ in $M(s)$ and obtaining arbitrarily long output paths. For a formal proof it is convenient to build a graph $G^{\rm o}_M$ of which each path corresponds to a pair $(u,v)$ and a state $q$ such that there is a sentential form $\xi$ with $q_{0,p}\Rightarrow^*_s \xi$ and $\xi/v=q(u)$, i.e., such that $(q,u,\#)\in\lk_s(v)$, see Lemma~\ref{lm:linkdef}(1). The pumping of $s$ then corresponds to the repetition of a cycle in $G^{\rm o}_M$. We construct a directed edge-labeled \emph{dependency graph} $G^{\rm o}_M$ with node set $Q$ and an edge with label $(j,z)$ from $q$ to $q'$ if there is a rule $q(a(x_1\ot p_{i,1},\ldots,x_k\ot p_{i,k}))\to \zeta$ in $R$ such that $\;z\in V(\zeta)$ and $\zeta/z=q'(x_j)$. Note that $|z|\leq \maxrhs(M)$. Each path $e_1\cdots e_n$ in $G^{\rm o}_M$ (where each $e_i$ is an edge) has a label in $\Nat_+^*\times\Nat_+^*$, obtained by the component-wise concatenation of the labels of $e_1,\dots,e_n$. \smallskip {\bf Claim.} There is a path $\pi$ from $q_{0,p}$ to $q$ with label $(u,v)$ in $G^{\rm o}_M$ if and only if there is a tree $s\in\sem{p}$ such that $(q,u,\#)\in\lk_s(v)$. \smallskip Proof of Claim. ($\Rightarrow$) We proceed by induction on the length of $\pi$. If $\pi$ is empty, then $q=q_{0,p}$ and $(u,v)=(\varepsilon,\varepsilon)$, and so $(q,u,\#)\in\lk_s(v)$ for every $s\in\sem{p}$ (by the definition of $\lk$). Now let $\pi'=\pi e$ where $\pi$ is a path from $q_{0,p}$ to $q$ with label $(u,v)$, and $e$ is an edge from $q$ to $q'$ with label $(j,z)$. Thus, $\pi'$ is a path from $q_{0,p}$ to $q'$ with label $(uj,vz)$. By the definition of $G^{\rm o}_M$, the edge $e$ is obtained from a rule $q(a(x_1\ot p_1,\ldots,x_k\ot p_k))\to \zeta$ such that $z\in V(\zeta)$ and $\zeta/z=q'(x_j)$. By induction, $(q,u,\#)\in\lk_s(v)$. Choose $s_m\in\sem{p_m}$ for $m\in[k]$, and let $s'=s[u\leftarrow a(s_1,\dots,s_k)]$. By Lemma~\ref{lm:linkdef}(4), we have $\delta(s/u)=\rho(q)$ and so $\delta(s/u)=\delta(a(s_1,\dots,s_k))$ because $M$ is la-uniform. Hence $\delta(s)=\delta(s')$ and every proper ancestor of $u$ is similar in $s$ and $s'$. Consequently $(q,u,\#)\in\lk_{s'}(v)$ by Lemma~\ref{lm:link_preserve}(1). Since $\rhs(q,s',u)=\zeta$, we obtain that $(q',uj,\#)\in\lk_{s'}(vz)$ from the definition of $\lk$. ($\Leftarrow$) We proceed by induction on the length of $u$. If $u=\varepsilon$, then $q=q_{0,p}$ and $v=\varepsilon$ by the definition of $\lk$, and so the empty path satisfies the conditions. Now let $u'=uj$ with $u\in\Nat_+^*$ and $j\in\Nat_+$, and let $s\in\sem{p}$ such that $(q',u',\#)\in\lk_s(v')$. By the definition of $\lk$ there is a rule $q(a(x_1\ot p_1,\ldots,x_k\ot p_k))\to \zeta$ such that $a=\lab(s,u)$ and $p_m=\delta(s,um)$ for every $m\in[k]$; moreover, there exist $v\in\Nat_+^*$ and $z\in V(\zeta)$ such that $v'=vz$, $(q,u,\#)\in\lk_s(v)$ and $\zeta/z=q'(x_j)$. By induction, there is a path $\pi$ from $q_{0,p}$ to $q$ with label $(u,v)$. By the definition of $G^{\rm o}_M$ there is an edge $e$ from $q$ to $q'$ with label $(j,z)$. Hence $\pi e$ is a path from $q_{0,p}$ to $q'$ with label $(uj,vz)=(u',v')$. This ends the proof of the Claim. \smallskip Suppose that $\pi_0$ is a path from $q_{0,p}$ to $q$ with label $(u_0,v_0)$, and that $\pi$ is a cycle from $q$ to itself with label $(u,v)$ such that $v\neq \varepsilon$. Consider the ``pumped'' path $\pi_0\pi^k$ from $q_{0,p}$ to $q$, for any $k\geq 1$. By the Claim, there exists a tree $s_k\in\sem{p}$ such that $(q,u_0u^k,\#)\in\lk_{s_k}(v_0v^k)$. Since $v_0v^k\in V(M(s_k))$ and $|v_0v^k|\geq k$, we obtain that $\height(M(s_k))\geq k$. That contradicts the fact that $\{M(s)\mid s\in\sem{p}\}$ is finite. Now consider a tree $s\in\sem{p}$ and a leaf $v$ of $M(s)$. Let $\orig_s(v)=(q,u,z)$. Then, by Lemma~\ref{lm:orig_prop2}, there exists an ancestor $\hat{v}$ of $v$ such that $(q,u,\#)\in\lk_s(\hat{v})$ and $v=\hat{v}z$. By the Claim, there is a path $\pi$ from $q_{0,p}$ to $q$ with label $(u,\hat{v})$. As shown above, if $\pi$ contains a cycle, then its output label is empty. Hence there is a path $\pi_1$ without cycles from $q_{0,p}$ to $q$ with label $(u_1,\hat{v})$. Then $\pi_1$ has at most $|Q|-1$ edges. Since the second component of the label of every edge has length at most $\maxrhs(M)$, we obtain that $|\hat{v}|\leq \maxrhs(M)\cdot(|Q|-1)$. Hence $|v|=|\hat{v}|+|z|\leq |\hat{v}|+\maxrhs(M)\leq \maxrhs(M)\cdot|Q|$. \qed \end{proof} \begin{theorem}\label{th:output_bound} The number $\maxrhs(M)\cdot|Q|\cdot(|P|+2)$ is an output bound for $M$. \end{theorem} \begin{proof} For a set $\{q_M(s)\mid s\in\T_\Sigma,\,\delta(s)=p_1\}$ as in Definition~\ref{df:output_bound}(1), with $\rho(q)=p_1$, we construct the dtla $M_1$ from $M$ by changing its $p_1$-axiom into $q(x_0)$. Application of Lemma~\ref{lm:range_bound} to $M_1$ and $p_1$ gives the upper bound $\maxrhs(M)\cdot|Q|$ on the height of the trees in the given set. For a set $\{q_M(C[p])\mid C\in\C_\Sigma,\,\delta(C[p])=p_1,\,\delta(C[p'])=p'_1\}$ as in Definition~\ref{df:output_bound}(2), with $\rho(q)=p_1$, we first change $M$ into $M_1$ as before. It remains to consider the set $\{M_1(C[p])\mid C\in\C_\Sigma,\,\delta(C[p'])=p'_1\}$. In order to apply Lemma~\ref{lm:range_bound}, we modify the dtla $M_1^\circ$. We use a simple product construction to obtain from $M_1^\circ$ an equivalent dtla $M_2$ that ``recognizes'' input trees of the form $C[p]$ and computes $\delta(C[p'])$ by additional look-ahead. The dtla $M_2$ has the same input alphabet $\Sigma\cup P$ and output alphabet $\Delta\cup(Q\times P)$ as $M^\circ$ and $M_1^\circ$. It has the same set $Q$ of states as $M$ and $M_1$, and the set of look-ahead states $P\times(P\cup\{1,0\})$. For its transition function $\bar{\delta}$, the equality $\bar{\delta}(s)=(\bar{p},\bar{p}')$ with $s\in\T_\Sigma(P)$ means the following: $\delta(s)=\bar{p}$, if $\bar{p}'=1$ then $s\in\T_\Sigma$, if $\bar{p}'=0$ then $s$ contains more than one occurrence of $p$ or at least one occurrence of an element of $P-\{p\}$, and if $\bar{p}'\in P$ then there exists $C\in\C_\Sigma$ such that $s=C[p]$ and $\delta(C[p'])=\bar{p}'$. We leave the easy construction of $\bar{\delta}$ to the reader. The rules of $M_2$ are defined by: $\rhs_{M_2}(\bar{q},a,(\bar{p}_1,\bar{p}_1'),\dots,(\bar{p}_k,\bar{p}_k'))= \rhs_{M_1^\circ}(\bar{q},a,\bar{p}_1,\dots,\bar{p}_k)$, and the $(\bar{p},\bar{p}')$-axiom of $M_2$ is the $\bar{p}$-axiom of $M_1$. Thus, $M_2$ ignores the additional look-ahead and hence is equivalent to $M_1^\circ$. Note that $M_2$ is initialized, but need not be la-uniform. We finally turn $M_2$ into an equivalent initialized la-uniform dtla $M_3$ by an obvious variant of the construction in the proof of Lemma~\ref{lm:make_uniform}, such that $M_3$ has state set $Q_3=Q\times(P\cup\{1,0\})$ and la-map $\rho_3$ with $\rho_3(\angl{\bar{q},\bar{p}'})=\rho(\bar{q})$. We also leave this construction to the reader. Note that $\maxrhs(M_3)=\maxrhs(M)$. Application of Lemma~\ref{lm:range_bound} to $M_3$ and $(p_1,p_1')$ gives the upper bound $\maxrhs(M_3)\cdot|Q_3|=\maxrhs(M)\cdot|Q|\cdot(|P|+2)$ on the height of the trees in the given set. \qed \end{proof}
\section{Introduction} A knotted circle $K$ in $S^3$ is called \emph{smoothly} (resp. \emph{topologically}) \emph{slice} if it bounds a smoothly (locally flat) embedded disk in $B^4$. Two knots $K_1$ and $K_2$ are called \emph{smoothly} (\emph{topologically}) \emph{concordant} if $K_1\# -K_2$ is smoothly (topologically) slice, where $-K$ is the mirror of $K$ with reversed orientation. Modulo smooth concordance, the set of knots forms an abelian group, \emph{the} (\emph{smooth}) \emph{knot concordance group}, $\mathcal C$. Note that every smoothly slice knot is topologically slice, but the converse is not true. We let $\mathcal C_{TS}$ be the subgroup of $\mathcal C$ generated by topologically slice knots. The Whitehead double (positively-clasped untwisted) of a knot $K$, $D(K)$, is a satellite knot defined by the pattern in Figure \ref{fig:Whiteheaddouble}. Whitehead doubles are interesting classes of knots in the study of the knot concordance. Any class of the Whitehead double of a knot is contained in $\mathcal C_{TS}$ since it has the same Alexander-Conway polynomial as the unknot and hence is topologically slice by a result of Freedman \cite{Freedman}. However, many of them are not smoothly slice, which show remarkable distinction between the smooth and topological categories in dimension four. It is thus important to understand their concordance properties as portrayed in the knot concordance group. It is also interesting to ask about the effect of $D$ on $\mathcal{C}$, and there is a long-standing conjecture. \begin{conjecture}\cite[Problem 1.38]{Kirby} $D(K)$ is smoothly slice if and only if $K$ is smoothly slice.\label{conj:Kirby}\end{conjecture} \begin{figure}[t!] \centering \includegraphics[width=\textwidth]{whdouble.eps} \caption{The pattern of the positively-clasped untwisted Whitehead double and the Whitehead double of the $(2,\,5)$ torus knot. The $-5$ extra full twists arise from untwisting the writhe of the projection of the $(2,\,5)$ torus knot.} \label{fig:Whiteheaddouble} \end{figure} Note that it is still unknown, as far as the author knows, if the conjecture is true even for some simple knots such as left-hand trefoil or figure-eight knot. One could study Whitehead doubles in $\mathcal{C}$ using homomorphisms from $\mathcal{C}$ to $\mathbb{Z}$. The knot signature $\sigma$ gives one such homomorphism. Unfortunately the signature, indeed any invariant of topological concordance group, is not effective homomorphism for Whitehead double knots, since it vanishes for these knots \cite{Freedman}. Heegaard Floer theory provides manifestly smooth concordance invariants, and some of which give homomorphisms from $\mathcal{C}$ to $\mathbb{Z}$. One is the $\tau$-invariant, defined using the knot Floer homology of Ozsv\'{a}th-Szab\'{o} and Rasmussen \cite{Ozsvath_Szabo_tau, Ozsvath_Szabo_knot, Rasmussen}. Manolescu-Owens discovered another concordance invariant $\delta$, twice the Heegaard Floer correction term ($d$-invariant) of the double cover of $S^3$ branched over a knot \cite{Manolescu_Owens}. More recently, Peters studied another concordance invariant $d(S^3_1(K))$ given by the correction term of 1-surgery on $K\subset S^3$ \cite{Peters}. In contrast to the other two invariants, $dS^3_1$ does not induce a homomorphism to $\mathbb{Z}$. See the survey paper \cite{Jabuka_Survey} of Jabuka for some applications of Heegaard Floer theory to the concordance group. Rasmussen's $s$-invariant coming from Khovanov homology is also a powerful concordance homomorphism \cite{Rasmussen_sinvariants}. It is known that $-\sigma/2=\tau=\delta=-s/2$ for alternating knots, but they differ in general: see \cite{Hedden_Ording}, \cite{Manolescu_Owens} and \cite{Livingston}. Even though there are many concordance invariants developed, most of them are inefficient for distinguishing Whitehead doubles in $\mathcal{C}$. The invariants $|\tau|$, $-dS^3_1/2$ and $|s/2|$ are known to be bounded above by the slice genus (four-ball genus) of the knot: the minimal genus of smoothly embedded surface in the 4-ball bounded by $K\subset\partial(B^4)$. Since the slice genus of $D(K)$ is at most one for any knot $K$, so are $|\tau|$, $-dS^3_1/2$ and $|s/2|$ of $D(K)$. Moreover, $\tau(D(K))$ is determined by $\tau(K)$ followed by the Theorem of Hedden below. \begin{theorem}\cite[Theorem 1.5]{Hedden} \begin{equation*} \tau(D(K)) = \begin{cases} 0, & \mbox{for}\hspace{0.3cm} \tau(K)\leq 0\\1, & \mbox{for}\hspace{0.3cm} \tau(K)> 0\end{cases} \end{equation*} \label{thm:Hedden} \end{theorem} In particular, $\tau(D^n(K))$ is identically either 0 or 1 for any $n\geq 1$ and is determined by $\tau(K)$, where $D^n(K)$ denote the $n$th iterated positively clapsed untwisted Whitehead double of $K$. Therefore, it is interesting to ask if it is possible to distinguish the $D^n(K)$'s in $\mathcal C$. Using $\delta$-invariants, which are not constrained by the slice genus, we show: \begin{theorem} For each $m\geq 2$, $D(T_{2,\,2m+1})$ and $D^2(T_{2,2m+1})$ are not smoothly concordant. In fact, they generate a $\mathbb{Z}^2$ summand of $\mathcal{C}_{TS}$. \label{thm:easy} \end{theorem} In \cite{Manolescu_Owens} it was computed that $\delta(D(T_{2,\,2m+1}))=-4m$. See also Section \ref{subsec:dinvariant}. Here, we show that $\delta(D^2(T_{2,\,2m+1}))=-4$ for any $m\geq 1$. A tool for our results is a computation of the infinity version of the knot Floer chain complex of $D(T_{2,\,2m+1})$: \begin{theorem} For any $m\geq 1$, the chain complex $CFK^\infty(D(T_{2,\,2m+1}))$ is $\mathbb{Z}\oplus\mathbb{Z}$ filtered chain homotopy equivalent to the chain complex $CFK^\infty(T_{2,\,3})\oplus A$, where $A$ is an acyclic complex. \label{thm:main} \end{theorem} More generally, we have the following: \begin{theorem} Suppose $K$ is a knot in $S^3$. If $|\delta(D(K))|>8$, then $D(K)$ and $D^n(K)$ are not smoothly concordant for each $n\geq 2$. If, in addition, $\tau(K)>0$, they generate a $\mathbb{Z}^2$ summand of $\mathcal{C}_{TS}$. \label{cor} \end{theorem} \begin{corollary} Suppose $K$ is an alternating knot in $S^3$. If $\tau(K)>2$, then $D(K)$ and $D^n(K)$ are not smoothly concordant for each $n\geq 2$. In fact, they generate a $\mathbb{Z}^2$ summand of $\mathcal{C}_{TS}$.\label{cor4} \end{corollary} Additionally, for some classes of knots including $(p,\,q)$ torus knots, we give an algorithmic formula for testing it in terms of its Alexander-Conway polynomial along with an implementation of a computer program. Recently, Cochran-Harvey-Horn suggested a bipolar filtration of $\mathcal C$ and the induced filtration of $\mathcal C_{TS}$ \cite{Cochran_Harvey_Horn}, $$\{0\}\subset\cdots\subset\mathcal T_{n+1}\subset\mathcal T_n\subset\cdots\subset\mathcal T_0\subset\mathcal T=\mathcal C_{TS}.$$ Since $\tau$ of $D(K)$ and $D^2(K)$ are nonzero for knots $K$ in Theorem \ref{thm:easy} and \ref{cor}, both of them are contained in $\mathcal T/\mathcal T_0$ by \cite[Corollary 4.9]{Cochran_Harvey_Horn}. Therefore, their filtration cannot see the difference between $D(K)$ and $D^2(K)$. On the other hand, using those knots, we get the following corollary relating to the filtration. Let $\mathcal{C}_{\Delta}$ be the subgroup of $\mathcal C_{TS}$ generated by knots with trivial Alexander-Conway polynomial. \begin{corollary} There is a $\mathbb Z^2$ summand of $\mathcal{C}_{\Delta}/\mathcal{C}_{\Delta}\cap\mathcal T_0$. \label{cor_filter} \end{corollary} \begin{remark} Recently, in \cite{Hom} Hom showed there is $\mathbb{Z}^\infty$ summand of $\mathcal C_{TS}$, but her technique cannot see the difference between the iterated Whitehead doubles in $\mathcal C$ either. \end{remark} \subsection*{Acknowledgements} The author would like to thank his PhD supervisors, Matt Hedden for suggesting this problem as well as priceless guidance and Ron Fintushel for encouraging and supporting him. He also thanks Jae Choon Cha and Chuck Livingston for helpful communication and Tim Cochran for pointing out a wrong statement of the earlier version of this paper and suggesting Corollary above. \section{Preliminaries} \label{sec:preliminary} In this section we briefly recall the Heegaard Floer and knot Floer theory, the staircase complexes and some invariants induced by them. \subsection{Heegaard Floer homology and knot Floer invariants} For simplicity, we work with coefficients in $\mathbb{F}$, the field of two elements. For a rational homology three-sphere $Y$ equipped with a spin$^c$ structure $\mathfrak{t}$, one can associate to it a relatively $\mathbb{Z}$ graded and filtered chain complex $CF^\infty(Y,\,\mathfrak{t})$, a finitely and freely generated $\mathbb{F}[U,\,U^{-1}]$-module. In particular the filtration is given by the negative power of $U$, and $U$-multiplication lowers the homological grading by 2. The filtered chain homotopy type of $CF^\infty(Y,\,\mathfrak{t})$ is known to be an invariant of $(Y,\,\mathfrak{t})$. For more detailed and general exposition of the theory we refer to \cite{Ozsvath_Szabo_3, Ozsvath_Szabo_4}. We set ${CF}^-(Y,\,\mathfrak{t}):=CF^\infty(Y,\,t)_{\{i<0\}}$, the subcomplex consisting of the elements in $CF^\infty(Y,\,t)$ whose filtration level $i$ is less than 0, and also define the quotient complexes $CF^+(Y,\,\mathfrak{t}){:}{=}CF^\infty(Y,\,\mathfrak{t})_{\{i\geq0\}}$ and $\widehat{CF}(Y,\,\mathfrak{t}){:=}CF^\infty(Y,\,t)_{\{i=0\}}$. The homology group of $CF^\infty(Y,\,\mathfrak{t})$ is denoted by $HF^\infty(Y,\,\mathfrak{t})$, and $HF^-$, $HF^+$ and $\widehat{HF}$ denote the homology of the other chain complexes. The various versions of Heegaard Floer homologies naturally fit into a long exact sequence: \begin{equation} \cdots\rightarrow HF^-(Y,\,\mathfrak{t})\rightarrow HF^\infty(Y,\,\mathfrak{t})\rightarrow HF^+(Y,\,\mathfrak{t})\rightarrow\cdots \label{eq:les} \end{equation} For $S^3$, it is known that $HF^\infty(S^3)\cong \mathbb{F}[U,\,U^{-1}]$ and $\widehat{HF}(S^3)\cong \mathbb{F}$ \cite{Ozsvath_Szabo_4}. We usually drop the spin$^c$ structure in the notation if there is a unique one. A knot $K$ in an integer homology three-sphere $Y$ has an associated $\mathbb{Z}\oplus\mathbb{Z}$ filtered chain complex $CFK^\infty(Y,K)$ which reduces to $CF^\infty(Y)$ after forgetting the second $\mathbb{Z}$ filtration. The $U$-multiplication decreases both of the filtration levels by $1$. The filtered chain homotopy type of $CFK^\infty(Y,\,K)$ is an invariant of the knot and we refer it as the knot Floer invariant of $(Y,\,K)$ \cite{Ozsvath_Szabo_knot, Rasmussen}. We denote by $CFK^\infty(Y,\,K)_{\{(i,\,j)\}}$ the subgroup at $(i, j)$-filtration level in $CFK^\infty(Y,\,K)$ and define $\widehat{CFK}(Y,\,K){:=}CFK^\infty(Y,\,K)_{\{i=0\}}$. It is an easy fact that $H_*(CFK^\infty(Y,\,K))\cong HF^\infty(Y)$ and $H_*(\widehat{CFK}(Y,\,K))\cong \widehat{HF}(Y)$. As a consequence, for a knot $K$ in $S^3$, we obtain an induced sequence of maps: $$\iota^m_{K}:H_*(\widehat{CFK}(S^3,\,K)_{\{j\leq m\}})\rightarrow \widehat{HF}(S^3)\cong\mathbb{F}$$ An invariant $\tau$ for a knot $K$ in $S^3$ is defined by $$\tau(K){:=}\min\{m\in\mathbb{Z}{|}\,\iota^m_{K} \mbox{ is non-trivial}\}.$$ In \cite{Ozsvath_Szabo_tau, Rasmussen} it was shown that $\tau$ is a concordance invariant. For a knot $K$ in $S^3$ we abbreviate the notations by $CFK^\infty(K):=CFK^\infty(S^3,\,K)$ and $\widehat{CFK}(K):=\widehat{CFK}(S^3,\,K)$. It is useful to visualize a knot Floer complex as a collection of dots and arrows lying in a grid in the plane. In a diagram, a dot in $(i,\,j)$-coordinate box represents an $\mathbb{F}$-generator in $CFK^\infty(K)_{\{(i,\,j)\}}$, and an arrow represents the non-trivial map, $\mathbb{F}\rightarrow\mathbb{F}$. The differential is then the sum of the arrows, as a map of vector spaces. See Figure \ref{fig:cfk} for examples. \subsection{Staircase complex} \label{sec:staircase} For a given $(n-1)$-tuple of positive integers $\mathbf{v}$, a \emph{staircase complex of length $n$}, St$(\mathbf{v})$, is defined as a finitely generated $\mathbb{Z}\oplus\mathbb{Z}$-filtered chain complex over $\mathbb{F}$ with $n$ generators, where the numbers in $\mathbf{v}$ are the length of arrows, which alternate horizontal and vertical starting at the top left generator and moving to the bottom right generator in alternating right and downward steps in a grid diagram. We also locate the top left dot on the vertical axis $(i=0)$ and the bottom right on the horizontal axis $(j=0)$ on the diagram. See \cite{Borodzik_Livingston} for more detail. For instance, complexes generated by $\mbox{St}(1,\,1,\,1,\,1)$ and $\mbox{St}(1,\,2,\,2,\,1)$ are shown in Figure \ref{fig:cfk}. The knot Floer invariant is a categorification of Alexander-Conway polynomial $\Delta_K(t)$, in the following sense: $$\Delta_K(t)=\sum_{k\in\mathbb{Z}}\chi(\widehat{CFK}(K)_{\{j=k\}} )t^k,$$ where $\Delta_K(t)$ is the symmetrized Alexander-Conway polynomial of $K$ and $\chi$ is the Euler characteristic. Conversely, for some classes of knots such as alternating knots and $L$-space knots (including torus knots), the knot Floer complex of a knot is determined by its Alexander-Conway polynomial \cite{Ozsvath_Szabo_alternating, Ozsvath_Szabo_lensspacesurgery}. For an $L$-space knot, $K$, the Alexander-Conway polynomial has the form $\Delta_K(t)=\sum_{k=0}^{2m}(-1)^kt^{n_k}$, and $CFK^\infty(K)$ is generated by a stair complex, $$CFK^\infty(K)\cong\mbox{St}(n_{i+1}-n_{i})\otimes\mathbb{F}[U,\,U^{-1}],$$ where $i$ runs from $0$ to $2m-1$ and $U$-multiplication is naturally extended: i.e. if $x$ is a generator in $(i,\,j)$-filtration level, then $U^nx$ sits in $(i-n,\,j-n)$-filtration level, and $\partial(U^nx)=U^n\partial(x)$. Denote $\mbox{St}(K):=\mbox{St}(n_{i+1}-n_i)$. For example, since the Alexander-Conway polynomial of $T_{2,\,2m+1}$ is $\sum_{i=-m}^{m}(-1)^it^i$, $CFK^\infty(T_{2,\,2m+1})$ is generated by $\mbox{St}(1,\cdots,\,1)$ of length $2m+1$. The knot Floer complex of $T_{3,\,4}$ can be given by $\mbox{St}(1,\,2,\,2,\,1)\otimes\mathbb{F}[U,\,U^{-1}]$ from the fact that $\Delta_{T_{3,\,4}}(t)=t^{-3}-t^{-2}+1-t^2+t^3$. See Figure \ref{fig:cfk}. Accordingly, it is easily obtained that $\tau(T_{p,\,q})=(p-1)(q-1)/2$ from its staircase complex, see also \cite[Corollary 1.7.]{Ozsvath_Szabo_tau}. In particular, $\tau(T_{2,\,2m+1})$ is equal to $m$. \begin{figure}[t!] \psscalebox{0.6}{ \begin{pspicture}(0,0)(19,9) \psgrid[subgriddiv=1,griddots=10,gridlabels=0pt](0,0)(9,9) \psgrid[subgriddiv=1,griddots=10,gridlabels=0pt](10,0)(19,9) \psline{}(4,0)(4,9) \psline{}(5,0)(5,9) \psline{}(0,4)(9,4) \psline{}(0,5)(9,5) \psline{}(14,0)(14,9) \psline{}(15,0)(15,9) \psline{}(10,4)(19,4) \psline{}(10,5)(19,5) \multirput(0.5,2.5)(1,1){7}{\psdot} \multirput(0.5,1.5)(1,1){8}{\psdot} \multirput(0.5,0.5)(1,1){9}{\psdot} \multirput(1.5,0.5)(1,1){8}{\psdot} \multirput(2.5,0.5)(1,1){7}{\psdot} \multips(0.5,1.4)(1,1){8}{\psline{->}(0,0)(0,-0.8)} \multips(2.5,1.4)(1,1){7}{\psline{->}(0,0)(0,-0.8)} \multips(1.4,0.5)(1,1){8}{\psline{->}(0,0)(-0.8,0)} \multips(1.4,2.5)(1,1){7}{\psline{->}(0,0)(-0.8,0)} \psline{-}(0.4,1.5)(0,1.5) \psline{-}(1.5,0.4)(1.5,0) \psline{->}(8.5,9)(8.5,8.6) \psline{->}(9,8.5)(8.6,8.5) \multirput(10.5,3.5)(1,1){6}{\psdot} \multirput(10.5,2.5)(1,1){7}{\psdot} \multirput(10.5,0.5)(1,1){9}{\psdot} \multirput(12.5,0.5)(1,1){7}{\psdot} \multirput(13.5,0.5)(1,1){6}{\psdot} \multips(10.6,3.5)(1,1){6}{\psline{<-}(0,0)(0.8,0)} \multips(10.5,2.4)(1,1){7}{\psline{->}(0,0)(0,-1.8)} \multips(10.6,0.5)(1,1){7}{\psline{<-}(0,0)(1.8,0)} \multips(13.5,1.4)(1,1){6}{\psline{->}(0,0)(0,-0.8)} \psline{-}(10,2.5)(10.4,2.5) \psline{-}(12.5,0.4)(12.5,0) \psline{->}(17.5,9)(17.5,7.6) \psline{->}(19,7.5)(17.6,7.5) \psline{->}(18.5,9)(18.5,8.6) \psline{->}(19,8.5)(18.6,8.5) \end{pspicture}} \caption{Diagrams of $CFK^\infty(T_{2,\,5})$ and $CFK^\infty(T_{3,\,4})$ generated by staircase complexes $\mbox{St}(1,\,1,\,1,\,1)$ and $\mbox{St}(1,\,2,\,2,\,1)$ respectively.}\label{fig:cfk} \end{figure} \subsection{The absolute grading, correction term and concordance invariants} \label{sec:invariants} There is a well-defined $\mathbb{Q}$-valued map from generators of Heegaard Floer homology groups called \emph{the absolute grading} with the following properties \cite{Ozsvath_Szabo_fourmanifold}: \begin{itemize} \item The absolute grading respects the homological grading of $CF^\infty(Y)$. \item The absolute grading of a generator in $\widehat{HF}(S^3)\cong\mathbb{F}$ is 0. \item If $(W,\,\mathfrak{s})$ is a cobordism from $(Y_1,\,\mathfrak{t}_1)$ to $(Y_2,\,\mathfrak{t}_2)$, then $${gr}(F^+_{W,\,\mathfrak{s}}(\xi))-{gr}(\xi)=\frac{c_1^2(\mathfrak{s})-2\chi(W)-3\sigma(W)}{4}$$ for $\xi\in HF^+(Y_1,\,\mathfrak{t_1})$. Here, $c_1$ is the first Chern class, $\chi$ is the Euler characteristic, $\sigma$ is the signature of the intersection form of $W$, and $F^+_{W,\,\mathfrak{s}}{:}\,HF^+(Y_1,\,\mathfrak{t}_1)\rightarrow HF^+(Y_2,\mathfrak{t}_2)$ is the map induced by the cobordism $W$. See \cite{Ozsvath_Szabo_fourmanifold} for a discussion about the maps induced by a four-dimensional cobordism. \end{itemize} We usually write down the absolute grading using a subscript with parenthesis, for example, $\widehat{HF}(S^3)\cong\mathbb{F}_{(0)}$. With the help of the absolute grading, \emph{the correction term} (\emph{$d$-invariant}) of $(Y,\,\mathfrak{t})$ is defined \cite{Ozsvath_Szabo_d}: $$d(Y,\,\mathfrak{t}){:}{=}\min\{{gr}(\xi){|}\,\xi\in HF^\infty(Y,\,\mathfrak{t}) \mbox{ and } \pi(\xi) \mbox{ is nontrivial}\},$$ where $\pi$ is the map from $HF^\infty(Y,\,\mathfrak{t})$ to $HF^+(Y,\,\mathfrak{t})$ in (\ref{eq:les}). In \cite{Manolescu_Owens}, Manolescu-Owens showed that the $d$-invariant of the double cover of $S^3$ branched over a knot $K$ is a concordance invariant for $K$: $$\delta(K){:}=2d(\Sigma(K),\,\mathfrak{t}_0),$$ where $\Sigma(K)$ is the branched double cover of $S^3$ along $K$, and $\mathfrak{t}_0$ is the Spin$^c$ structure of $\Sigma(K)$ such that $c_1(\mathfrak{t}_0){=}0{\in} H^2(\Sigma(K);\,\mathbb{Z})$. They showed that $\delta=\tau$ for alternating knots and knots with up to nine crossings. They also showed $\delta$ of Whitehead double of an alternating knot $K$ determined by $\tau(K)$: \begin{theorem}\cite[Theorem 1.5.]{Manolescu_Owens} If $K$ is alternating, then $$\delta(D(K))=-4\max\{\tau(K),0\}.$$ \label{thm:Manolescu_Owens} \end{theorem}Therefore, it easily follows from the theorem and the computation of $\tau$ in the previous section that $\delta(D(T_{2,\,2m+1})){=}-4m$. See also Section \ref{subsec:dinvariant}. The $d$-invariant of the 1-surgery of $S^3$ along the knot $K$, $d(S^3_1(K))$, is also a concordance invariant. In \cite{Peters}, Peters gives an algorithm of computing $d(S^3_1(K))$, provided knowledge of $CFK^\infty(K)$. \section{Infinity version of knot Floer complex of $D(T_{2,\,2m+1})$} \label{sec:computation} Recently, Hedden-Kim-Livingston showed that $CFK^\infty(D(T_{2,\,3}))$ is chain homotopy equivalent to $CFK^\infty(T_{2,\,3})\oplus A$ for some acyclic complex $A$, \cite[Proposition 6.1.]{Hedden_Kim_Livingston}. Also see \cite[section 9.1]{Cochran_Harvey_Horn}. Here, we will prove that the result can be generalized to the torus knots $T_{2,\,2m+1}$ for $m\geq 1$, and furthermore $CFK^\infty(D(T_{2,\,2m+1}))$ will be completely determined. Before proving the theorem, recall the following useful lemma regarding how a basis change in a filtered chain complex over $\mathbb{F}$ affects the diagram of a knot Floer chain complex. \begin{lemma}\cite[Lemma A.1.]{Hedden_Kim_Livingston} Let $C_*$ be a knot Floer complex with a 2-dimensional arrow diagram $D$ given by an $\mathbb{F}$-basis. Suppose that $x$ and $y$ are two basis elements of the same grading such that each of the $i$ and $j$ filtrations of $x$ is not greater than that of $y$. Then the $\mathbb{Z}\oplus\mathbb{Z}$ filtered basis change given by $y'=y+x$ gives rise to a diagram $D'$ of $C_*$ which differs from $D$ only at $y$ and $x$ as follows: \begin{itemize} \item Every arrow from some $z$ to $y$ in $D$ adds an arrow from $z$ to $x$ in $D'$ \item Every arrow from $x$ to some $w$ in $D$ adds an arrow from $y'$ to $w$ in $D'$ \end{itemize} \end{lemma} We use the above lemma for the purpose of removing certain boundary arrows in chain complexes over $\mathbb{F}$. For example, the proposition below will be useful for proving Theorem \ref{thm:main}. \begin{proposition} Suppose $C$ is one of the $\mathbb{Z}\oplus\mathbb{Z}$ filtered chain complexes over $\mathbb{F}$ given by the diagrams in Figure \ref{fig:uvw} with any possible combination of dotted arrows. Then all dotted arrows can be removed by a basis change. \label{cor2} \end{proposition} \begin{proof} First, consider the complex (I). Suppose that $$ \begin{array}{l} \partial a=b+c+Ax+By,\\ \partial b=d+Cz,\\ \mbox{and } \partial c=d+Dz \end{array} $$ for some $A$, $B$, $C$ and $D$ in $\mathbb{F}$. Since $\partial^2=0$, \begin{equation} \begin{array}{lcl} 0=\partial^2a&=&\partial(b+c+Ax+By)\\ &=&(A+B+C+D)z.\end{array} \label{eq:1}\end{equation} Therefore, the coefficients have to satisfy the equation that $A+B+C+D{=}0$. Now, we consider every possible coefficient of $A$, $B$, $C$ and $D$ in $\mathbb{F}$ satisfying the equation and show that each case can be transformed to have $A{=}B{=}C{=}D{=}0$, as desired, after proper change of basis. \begin{itemize} \item If $A{=}B{=}1$ and $C{=}D{=}0$, change the basis by $b'{=}b+x$, $c'{=}c+y$ and $d'{=}d+z$. \item If $A{=}C{=}1$ and $B{=}D{=}0$, change the basis by $b'{=}b+x$. \item If $A{=}D{=}1$ and $B{=}C{=}0$, change the basis by $b'{=}b+x$ and $d'{=}d+z$. \item If $B{=}C{=}1$ and $A{=}D{=}0$, change the basis by $c'{=}c+y$ and $d'{=}d+z$. \item If $B{=}D{=}1$ and $A{=}C{=}0$, change the basis by $c'{=}c+y$. \item If $C{=}D{=}1$ and $A{=}B{=}0$, change the basis by $d'{=}d+z$. \item If $A{=}B{=}C{=}D{=}1$, change the basis by $b'{=}b+x$ and $c'{=}c+y$. \end{itemize} Similar argument is applied to remove any combination of possible dotted arrows in the complexes (II) and (III). \end{proof} \begin{figure}[t!] \centering \psscalebox{0.55}{ \begin{pspicture}*(0,15)(24,26) \psset{nodesep=1pt, xunit=20pt, yunit=20pt} \rput(1,33){\Large(I)} \cnodeput(1,23){v1}{$z$} \cnodeput(1,26){u}{$x$} \cnodeput(4,23){w}{$y$} \cnodeput(4,26){v3}{$w$} \ncline{->}{v3}{u} \ncline{->}{v3}{w} \ncline{->}{u}{v1} \ncline{->}{w}{v1} \cnodeput(6,29){v21}{$d$} \cnodeput(6,32){u2}{$b$} \cnodeput(9,29){w2}{$c$} \cnodeput(9,32){v23}{$a$} \ncline{->}{v23}{u2} \ncline{->}{v23}{w2} \ncline{->}{u2}{v21} \ncline{->}{w2}{v21} \ncarc{->,linestyle=dashed}{v23}{u} \ncput*{$A$} \ncarc{->, linestyle=dashed}{v23}{w} \ncput*{$B$} \ncarc{->, linestyle=dashed}{u2}{v1} \ncput*{$C$} \ncarc{->, linestyle=dashed}{w2}{v1} \ncput*{$D$} \rput(12,33){\Large(II)} \dotnode(12,23){y1} \dotnode(12,26){x} \dotnode(15,23){z} \dotnode(15,26){y2} \ncline{->}{y2}{x} \ncline{->}{y2}{z} \ncline{->}{x}{y1} \ncline{->}{z}{y1} \dotnode(17,29){v11} \dotnode(17,32){u1} \dotnode(20,29){w1} \dotnode(20,32){v12} \ncline{->}{v12}{u1} \ncline{->}{u1}{v11} \ncline{->}{w1}{v11} \ncline{->}{v12}{w1} \ncarc{->,linestyle=dashed}{u1}{x} \ncarc{->, linestyle=dashed}{u1}{z} \ncarc{->, linestyle=dashed}{v12}{y2} \ncarc{->, linestyle=dashed}{w1}{x} \ncarc{->, linestyle=dashed}{w1}{z} \rput(23,33){\Large(III)} \dotnode(23,26){x} \dotnode(26,23){z} \dotnode(26,26){y} \ncline{->}{y}{x} \ncline{->}{y}{z} \dotnode(28,29){v21} \dotnode(28,32){u2} \dotnode(31,29){w2} \dotnode(31,32){v23} \ncline{->}{v23}{u2} \ncline{->}{v23}{w2} \ncline{->}{u2}{v21} \ncline{->}{w2}{v21} \ncarc{->,linestyle=dashed}{u2}{x} \ncarc{->, linestyle=dashed}{u2}{z} \ncarc{->, linestyle=dashed}{w2}{x} \ncarc{->, linestyle=dashed}{w2}{z} \ncarc{->, linestyle=dashed}{v23}{y} \end{pspicture}} \caption{Any possible combination of dotted arrows can be removed by a basis change.} \label{fig:uvw} \end{figure} \begin{proof}[Proof of Theorem 4] Let $D$ be $D(T_{2,\,2m+1})$ for $m\geq 1$. Theorem 1.2 of \cite{Hedden} together with the computation of $\widehat{CFK}(T_{2,\,2m+1})$ shows that $$\widehat{HFK}_*(D,\,j)= \begin{cases} \mathbb{F}^{2m}_{(0)}\oplus\mathbb{F}^2_{(-1)}\oplus\mathbb{F}^2_{(-3)}\oplus\cdots\oplus\mathbb{F}^2_{(-2m+1),}&j=1\\ \mathbb{F}^{4m-1}_{(-1)}\oplus\mathbb{F}^4_{(-2)}\oplus\mathbb{F}^4_{(-4)}\oplus\cdots\oplus\mathbb{F}^4_{(-2m),}&j=0\\ \mathbb{F}^{2m}_{(-2)}\oplus\mathbb{F}^2_{(-3)}\oplus\mathbb{F}^2_{(-5)}\oplus\cdots\oplus\mathbb{F}^2_{(-2m-1),}&j=-1\\ 0&otherwise. \end{cases}$$ We assign an $\mathbb{F}$-basis to each summand in the direct decomposition as below: $$\widehat{HFK}_*(D,\,j)= \begin{cases} \langle x^0_1,\,\cdots,\,x^0_{2m}\rangle\oplus\langle u^{-1}_{1,\,1},\,u^{-1}_{1,\,2}\rangle\oplus\cdots\oplus\langle u^{-2m+1}_{m,\,1},\,u^{-2m+1}_{m,\,2}\rangle&j=1\\ \langle y^{-1}_1,\,\cdots,\,y^{-1}_{4m-1}\rangle\oplus\langle v^{-2}_{1,\,1},\cdots,\,v^{-2}_{1,\,4}\rangle\oplus\cdots\oplus\langle v^{-2m}_{m,\,1},\cdots,\,v^{-2m}_{m,\,4}\rangle&j=0\\ \langle z^{-2}_1,\,\cdots,\,z^{-2}_{2m}\rangle\oplus\langle w^{-3}_{1,\,1},\,w^{-3}_{1,\,2}\rangle\oplus\cdots\oplus\langle w^{-2m-1}_{m,\,1},\,w^{-2m-1}_{m,\,2}\rangle&j=-1\\ 0&otherwise, \end{cases}\label{eq:generators}$$ where the superscript of a generator represents its absolute grading. Since $\widehat{HFK}_*(D)$ is homotopic equivalent to the $\widehat{CFK}(D)$ \cite[Lemma 4.5]{Rasmussen}, we assume that $CFK^\infty(D)_{\{(0,\,j)\}}=\widehat{HFK}_*(D,\,j)$ and $CFK^\infty(D)_{\{(i,\,j)\}}\cong U^{-i}CFK^\infty(D)_{\{(0,\,j)\}}=\widehat{HFK}_*(D,\,j-i)$. Now, we investigate all differentials in $CFK^\infty(D)$ by using the facts, $\partial^2=0$, $H_*(\widehat{CFK}(D))\cong\widehat{HF}(S^3)\cong \mathbb F_{(0)}$ and $H_*(CFK^\infty(D))\cong HF^\infty(S^3)\cong \mathbb{F}[U,\,U^{-1}]$. First, note that there are no components of the boundary maps between generators of the same $(i,\,j)$-filtration since they would reduce $\widehat{HFK}_*(D)$. Thus, we can decompose the boundary maps $\partial$ to the vertical, horizontal and diagonal components, $\partial=\partial_V+\partial_H+\partial_D$. Also, we remark that it is enough to determine boundary maps of $\mathbb{F}[U,\,U^{-1}]$-generators in $CFK^\infty$ because the boundary map is $U$-equivariant. Exactly the same argument as \cite[Proposition 6.1]{Hedden_Kim_Livingston} can be used to determine $\partial_V$ and $\partial_H$ i.e. by the fact that $H_*(\widehat{CFK}(D))=\mathbb{F}_{(0)}$ and using grading consideration, after changing basis, we can assume that $$ \begin{array}{ll}\partial_V(x^d_k)=y^{d-1}_{k-1} &\mbox{for } k=2,\cdots, 2m\\ \partial_V(y^{d-1}_{2m+l-1})=z^{d-2}_l &\mbox{for } l=1,\cdots,2m.\\ \partial_V(u^{d-1}_{p,\,i})=v^{d-2}_{p,\,i}\mbox{ and }\partial_V(v^{d-2}_{p,\,i+2})=w^{d-3}_{p,\,i}&\mbox{for } p=1,\cdots,m \mbox{ and } i=1,\,2, \end{array} $$ for $d\in 2\mathbb{Z}$ and $\partial_V$'s of other elements are trivial. Analogously, since $H_*(CFK^\infty(D)_{\{j=0\}})$ is isomorphic to $\widehat{HF}(S^3)\cong\mathbb{F}_{(0)}$ and $\partial^2=0$, the horizontal boundary components can be assumed as following: $$ \begin{array}{ll}\partial_H(z^d_k)=y^{d-1}_{k-1} &\mbox{for } k=2,\cdots, 2m\\ \partial_H(y^{d-1}_{2m+l-1})=x^{d-2}_l &\mbox{for } l=1,\cdots,2m.\\ \partial_H(w^{d-1}_{p,\,i})=v^{d-2}_{p,\,i}\mbox{ and }\partial_H(v^{d-2}_{p,\,i+2})=u^{d-3}_{p,\,i}&\mbox{for } p=1,\cdots,m \mbox{ and } i=1,\,2, \end{array} \label{eq:v} $$ and $\partial_H$'s of other elements are trivial. We drop $\mathbb{F}[U,\,U^{-1}]$ coefficients of generators since they are canonically determined by the grading consideration. See Figure \ref{fig:5} for a diagram. In fact we will show that we can assume there are no $\partial_D$ components for any elements as shown in Figure \ref{fig:5}. \input{fig_complex} We can split $CFK^\infty(D)$ as following disjoint subsets: $$\begin{array}{l} A^d_{p,\,i}:=\{v^{d}_{p,\,i+2},\,u^{d-1}_{p,\,i},\,w^{d-1}_{p,\,i},\,v^{d-2}_{p,\,i}\}\\ B^d_q:=\{y^{d-1}_{2m+q},\,x^{d-2}_{q+1},\,z^{d-2}_{q+1},\,y^{d-3}_q\}\\ \mbox{and }C^d:=\{y^{d-1}_{2m},\,x^{d-2}_1,\,z^{d-2}_1\}, \end{array}$$ for $1\leq p\leq m$, $1\leq q\leq 2m-1$, $i=1,\,2$ and $d\in 2\mathbb{Z}$. Note that any arrows between subsets must be diagonal. Disregarding the diagonal arrows between subsets, each complex of $A$'s and $B$'s has four generators with square-shaped grid diagram, and each complex of $C$'s has three generators which looks like $St(1,1)$. Therefore, if we remove all arrows between subsets i.e. $CFK^\infty(D)$ is a direct sum of $A$'s, $B$'s and $C$'s, the theorem follows. Define a subset of $A$ as $A'^d_{p,\,i}:=\{v^{d}_{p,\,i+m},\,u^{d-1}_{p,\,i},\,w^{d-1}_{p,\,i}\}$. Due to grading constraints on the filtered complex, we observe the following: \begin{itemize} \item $\partial_D$ of any generator in $A'^d_{p,\,i}$ has components only in $A^d_{k,\,j}$, $B^d_{q}$ and $C^d$ for $k<p$, $j=1,\,2$, and $q=1,\cdots,\,2m-1$ (i.e. diagonal arrows between $A$'s going from higher to smaller first index.) \item $\partial_D$ of the generators in $B$'s and $C$'s are zero. \end{itemize} These observations allow us to apply Corollary \ref{cor2} inductively to remove all diagonal arrows in the complex. We start to remove any diagonal arrows from $A'^d_{m,\,1}$. First, we remove all diagonal arrows going from $A'^d_{m,\,1}$ to $A^d_{m-1,\,1}$, using basis-change of the case (I) of Corollary \ref{cor2}. Differently from the corollary, there can be other components in $\partial_D$ of elements in $A'^d_{m,\,1}$, not in $A^d_{m-1,\,1}$. However, considering the grading constraints again, one can easily check that the other components cannot induce $z$ component of $\partial^2a$ in the Equation (\ref{eq:1}), hence the equation that $A+B+C+D=0$ in the proof of the corollary still holds and we remove diagonal arrows using a basis-change in the corollary. After applying the basis-change, two types of newer diagonal arrows will be added due to arrows coming to $A'^d_{m,\,1}$ and arrows going from $A^d_{m-1,\,1}$. First note that there are no diagonal arrows coming to $A^d_{m,\,1}$ by the observations above ($m$ is the greatest index for $A$.) Secondly, a diagonal arrow from $A^d_{m-1,\,1}$ to some generator adds an arrow going from $A^d_{m,\,1}$ to the generator after basis-change, but note that these arrows are going to the subsets $A^d_{p,\,i}$ with $p<m-1$ and $i=1,\,2$, which we will remove later. Now, we similarly change the basis for removing diagonal arrows from $A'^d_{m,\,1}$ to $A^d_{m-1,\,2}$, $A^d_{m-2,\,1}$, $A^d_{m-2,\,2}\cdots,\,A^d_{1,\,1},$ and $A^d_{1,\,2}$ in sequence. Then, case (II) and (III) of the corollary will be applied to remove arrows from $A'^d_{m,\,1}$ to $B^d_q$'s and $C^d$. The induction ends with removing any $\partial_D$ from $A'^d_{m,\,1}$, since there are no diagonal arrows from $A^d_{1,\,i}$'s, $B^d_q$'s and $C^d$. Then, we remove $\partial_D$ of $A'^d_{m,\,2}$, $A'^d_{m-1d,\,1}$, $A'^d_{m-1,\,2},\cdots,\,A'^d_{1,\,1},$ and $A'^d_{1,\,2}$ likewise. After removing the diagonal arrows from $A'_{p,\,i}$ for all $p=1,\cdots,\,m$ and $i=1,\,2$, the only remaining non-trivial $\partial_D$ are ones of $v_{p,\,1}$ and $v_{p,\,2}$. It is easy to see that $\partial_D$'s of $v_{p,\,1}$ and $v_{p,\,2}$ also vanish: $0{=}\partial^2(u_{p,\,i}){=}\partial(v_{p,\,i})$. Thus, we may assume that $\partial_D$'s of $CFK^\infty$ are all zero. \end{proof} \section{$\delta$-invariant of $D^2(T_{2,\,2m+1})$ and proof of Theorem \ref{thm:easy}} \label{sec:proof} First, we present a lemma that relates the $\delta$-invariant of a Whitehead double to $dS^3_1$. \begin{lemma} \label{lemma:delta} For any knot $K$, $\delta(D(K))=2dS^3_1(K\#K^r)$. \end{lemma} \begin{proof} Let $K_p$ denote the 3-manifold obtained by $p$-surgery of $S^3$ along a knot $K$. Manolescu-Owens showed that $d(K_{-1/2})=d(K_{-1})$ for any knot in the proof of \cite[Proposition 6.2]{Manolescu_Owens}; in fact both of them equal $2h_0(K)$, where $h_0(K)$ is an invariant defined by Rasmussen in \cite{Rasmussen}. Thus, using the behaviour of $d$-invariants under orientation reversal \cite[Propostion 4.2]{Ozsvath_Szabo_d}, we have $$\begin{array}{lcl}d(K_{1/2})&=&-d(\overline{K}_{-1/2})\\&=&-d(\overline{K}_{-1})\\&=&d(K_1),\end{array}$$ where $\overline{K}$ is the mirror image of $K$. Recall that the double cover of $S^3$ branched over $D(K)$, $\Sigma(D(K))$, can be obtained by $1/2$-surgery along $K\#K^r$ in $S^3$, where $K^r$ is the knot $K$ with its orientation reversed, see \cite[Proposition 6.1]{Manolescu_Owens}. From the definition of $\delta$-invariant, \begin{equation*}\begin{array}{lcl}\delta(K)&:=&2d(\Sigma(K))\\&=&2d((K\#K^r)_{1/2})\\&=&2d((K\#K^r)_1).\end{array}\end{equation*} \end{proof} \begin{proof}[Proof of Theorem \ref{thm:easy}] By applying the lemma above to $D^2(T)$, we have $$\delta(D^2(T))=2d((D(T)\#D(T)^r)_1),$$ where $T=T_{2,\,2m+1}$. To compute $d(K_1)$, it suffices to understand $CFK^\infty(K)$ \cite{Peters}. Let us recall the algorithm. Pick $\xi$, a generator of $H_*(CFK^\infty(K)_{\{i=0\}})\cong\mathbb{F}$. Note that any generator become zero in $CFK^\infty_{\{i\geq 0 \mbox{ or } j\geq 0\}}$ by the multiplication by high enough power of $U$. Then \begin{equation}\label{eq:d} d(K_1)=-2\min\{n\geq 0: [U^{n+1}\xi]=0\in H_*(CFK^\infty(K)_{\{i\geq 0 \mbox{ or } j\geq 0\}})\}. \end{equation} Observe that $d(K_1)$ is derived from a direct summand of $CKF^\infty(K)$ containing a generator of $H_*(CFK^\infty(K)_{\{i=0\}})\cong\mathbb{F}$. In particular, if $CFK^\infty(K)$ and $CFK^\infty(K')$ are differ only by an acyclic complex, then $d(K_1)=d(K'_1)$. Now, let us understand $CFK^\infty(D(T)\#D(T)^r)$. Since the knot Floer complex is unchanged under the orientation reversal \cite[Proposition 3.9.]{Ozsvath_Szabo_knot} and by the connected sum formula for knot Floer complexes in \cite[Theorem 7.1.]{Ozsvath_Szabo_knot}, $$CFK^\infty(D(T)\#D(T)^r)\cong CFK^\infty(D(T))\otimes CFK^\infty(D(T)).$$ Thus, $d((D(T)\#D(T)^r)_1)$ equals to $d((T_{2,\,3}\#T_{2,\,3})_1)$ by Theorem \ref{thm:main}. It is computed that $d((T_{2,3}\#T_{2,3})_1)=-2$ as an example of the computer program, {\texttt dCalc} in \cite{Peters}, (it can also be computed by Proposition \ref{prop}) so that $\delta(D^2(T))=-4$. On the other hand, $\delta(D(T))=-4m$ as stated in Section \ref{sec:invariants}. The first part of Theorem \ref{thm:easy} follows. Recall that $\delta\equiv\sigma/2$ mod 4 \cite[(2.1)]{Manolescu_Owens} and $\sigma=0$ for any knot in $\mathcal{C}_{TS}$. Consider the homomorphism $\psi=(\tau,\,\delta/4):\mathcal{C}_{TS}\rightarrow\mathbb{Z}\oplus\mathbb{Z}$. Since $\psi(D(T))=(1,\,-m)$ and $\psi(D^2(T))=(1,\,-1)$, $\psi$ is surjective if $m\geq2$. Therefore, $\mathcal{C}_{TS}$ has a $\mathbb{Z}^2$ summand generated by $D(T)$ and $D^2(T)$. \end{proof} \begin{proof}[Proof of Corollary \ref{cor_filter}] By \cite[Corollary 4.9, Corollary 6.11]{Cochran_Harvey_Horn} both $\tau$ and $\delta$ vanish for the knots in $\mathcal{C}_{\Delta}\cap\mathcal T_0$. Now, consider the induced homomorphism $(\tau,\,\delta/4):\,\mathcal{C}_{\Delta}/\mathcal{C}_{\Delta}\cap\mathcal T_0\rightarrow\mathbb Z\oplus\mathbb Z$, and the subjectivity of it can be shown by the knots, $D(T)$ and $D^2(T)$. \end{proof} \section{Generalization of the result} \label{sec:generalization} In this section we discuss how to generalize the result for $T_{2, 2m+1}$ to other knots. First, we use a genus-bound property of the concordance invariant $dS^3_1$ to show that, provided that $|\delta(D(K))|>8$, $D(K)$ and $D^n(K)$ are not smoothly concordant for each $n\geq 2$. Secondly, we present formulas to compute $dS^3_1(K)$ and $\delta(D(K))=2dS^3_1(K\#K)$ for a given staircase complex of $K$ introduced in Section \ref{sec:staircase}. \subsection{Genus bound and proof of Theorem \ref{cor}} \begin{proof}[Proof of Theorem \ref{cor}]It is shown in \cite[Theorem 1.5.]{Peters} that $-dS^3_1(K)/2$ is a lower bound for the slice genus, $g^*(K)$, of $K$, and note that the slice genus of $D(K)$ is at most 1 for any knot $K$. Hence, for $n\geq2$, $$-\delta(D^n(K))=-2dS^3_1(D^{n-1}(K)\#D^{n-1}(K)^r)\leq 4g^*(D^{n-1}(K)\#D^{n-1}(K)^r)\leq 8.$$ Therefore, if $\delta(D(K))<-8$, equivalently $dS^3_1(K\#K)<-4$, $D(K)$ and $D^n(K)$ are not smoothly concordant. (According to \cite[Theorem 1.5]{Manolescu_Owens}, $\delta(D(K))$ is nonpositive for any knot $K$.) If $\tau(K)>0$, both $\tau(D(K))$ and $\tau(D^n(K))$ are 1 by Theorem \ref{thm:Hedden}. Now one can prove the second part of the theorem by considering the surjective homomorphism $(\tau,\,\delta/4):\,\mathcal{C}_{TS}\rightarrow\mathbb{Z}\oplus\mathbb{Z}$. \end{proof} \begin{proof}[Proof of Corollary \ref{cor4}] This is obtained by Theorem \ref{cor} together with Theorem \ref{thm:Manolescu_Owens}. \end{proof} \begin{remark} Since $\delta(D(T_{2,\,2m+1}))=-4m$, Theorem \ref{thm:easy} is a special case of Theorem \ref{cor} for $m\geq 3$, but we have shown it for the case $m=2$ as well by computing $\delta(D^2(T_{2,\,2m+1}))=-4$. \end{remark} \subsection{$dS^3_1(K)$ and $\delta(D(K))$ of a Staircase Complex} \label{subsec:dinvariant} If a knot admits a knot Floer complex generated by a staircase complex (equivalently, $L$-space knots), then its $d$-invariant can be easily obtained. For the $d$-invariants of higher surgery coefficients, we refer to \cite[Section 4.2]{Borodzik_Livingston}. \begin{proposition} Suppose the knot Floer complex of $K$ can be given by a staircase complex $\mbox{St}(K)$, then $$d(S^3_1(K))=-2\min_{(i,\,j)\in \mathrm{Vert}(\mathrm{St}(K))}\max\{i,\,j\}$$ $$\delta(D(K))/2=d(S^3_1(K\#K))=-2\min_{(i,\,j),\,(k,\,l)\in \mathrm{Vert}(\mathrm{St}(K))}\max\{i+k,\,j+l\},$$ where $\mathrm{Vert}(\mathrm{St}(K))$ is the set of the coordinates of the generators of $\mathrm{St}(K)$. \label{prop} \end{proposition} \begin{proof} Suppose that $CFK^\infty(K)$ is generated by $\mbox{St}(K)$, then the top left element in $\mbox{St}(K)$ represents the generator of $H_*({CFK}^\infty(K)_{\{i=0\}})\cong\mathbb{F}$: say $\xi$. The chain complex $\mbox{St}(K)$ has the form $0\rightarrow\mathbb{F}^k_{(+1)}\rightarrow \mathbb{F}_{(0)}^{k+1}\rightarrow 0$. Observe that any non-trivial generator $\eta$ of $\mathbb{F}_{(0)}^{k+1}$ is homologous to $\xi$. For $\eta$ with $(i,\,j)$-coordinates, $U^{\max\{i,\,j\}+1}\eta$ lies in the subcomplex $CFK^\infty(K)_{\{i<0 \mbox{ and } j<0\}}$, whereas $U^{\max\{i,\,j\}}\eta$ does not. Note also that since $\mathrm{St}(K)$ is a $\mathbb Z\oplus\mathbb Z$-filtered complex, $\min_{(i,\,j)\in \mathrm{Vert}(\mathrm{St}(K))}\max\{i,\,j\}$ is realized by the elements in $\mathbb{F}_{(0)}^{k+1}$, not ones in $\mathbb{F}^k_{(+1)}$. Hence, the first formula follows from the Equation (\ref{eq:d}). Although $CFK^\infty(K\#K)$ is not generated by a staircase complex, it can be constructed from the tensor complex, $\mbox{St}(K)\otimes \mbox{St}(K)$ by the connected-sum formula \cite[Theorem 7.1]{Ozsvath_Szabo_knot}. The coordinates of the generators in $\mbox{St}(K)\otimes \mbox{St}(K)$ are given by the sums of a pair of coordinates of the generators of $\mbox{St}(K)$. The complex $\mbox{St}(K)\otimes \mbox{St}(K)$ has the form $0\rightarrow \mathbb{F}^{k^2}_{(+2)}\rightarrow\mathbb{F}^{2k(k+1)}_{(+1)}\rightarrow\mathbb{F}^{(k+1)^2}_{(0)}\rightarrow 0$, and the generators with $(0)$-grading are homologous to the generator of $H_*({CFK}^\infty(K\#K)_{\{i=0\}})\cong\mathbb{F}$. See Figure \ref{fig:tensor} for the tensor complex of two copies of $\mbox{St}(1,\,2,\,2,\,1)$. Therefore, we get the second formula similarly. \end{proof} \input{fig_tensoring} For example, since $\mbox{St}(T_{2,\,2m+1})=(1,\cdots,\,1)$ of length $2m+1$, one can compute that $\delta(D(T_{2,\,2m+1}))=-4m$ again. In the case of $(3,\,4)$ torus knot, $\mbox{St}(T_{3,\,4})=(1,\,2,\,2,\,1)$, and so $$\mbox{Vert}(T_{3,\,4})=\{(0,\,3),\,(1,\,3),\,(1,\,1),\,(3,\,1),\,(3,\,0)\}.$$ Thus $\delta(D(T_{3,\,4}))=-8$, and hence we cannot figure out if $D(T_{3,\,4})$ and $D^2(T_{3,\,4})$ are concordant, using Theorem \ref{cor}. Note that there are many knots such that $\delta(D(K))=-8$: for example, any knot $K$ whose $CFK^\infty$ is generated by $\mbox{St}(1,\,n,\,n,\,1)$. However, one can similarly apply the arguments in Section \ref{sec:computation} to show $CFK^\infty(D(T_{3,\,4}))\cong CFK^\infty(T_{2,\,3})\oplus A$ for some acyclic complex $A$. Therefore, $\delta(D^2(T_{2,\,3}))=-4$, and we conclude that $D(T_{3,\,4})$ and $D^2(T_{3,\,4})$ are not concordant. For the right-handed trefoil knot, as far as the author knows, all concordance invariants of $D(T_{2,\,3})$ and $D^2(T_{2,\,3})$ are same, so it is still mysterious if $D(T_{2,\,3})$ and $D^2(T_{2,\,3})$ are smoothly concordant. \begin{conjecture} $D(T_{2,\,3})$ and $D^2(T_{2,\,3})$ are not smoothly concordant. \end{conjecture} \begin{remark} This conjecture is possibly approached using gauge-theoretic invariants. See \cite{Hedden_Kirk}. \end{remark} \subsection*{Implementation} We wrote a C++ program computing $dS^3_1(K)$ and $\delta(D(K))$ for a $(p,\,q)$ torus knot or a staircase complex of $K$. You may download the source file in the author's webpage. In order to algorithmically obtain the staircase complex (equivalently Alexander-Conway polynomial) of a $(p,\,q)$ torus knot in the program, we used the subsemigroup of $\mathbb{N}$ generated by $p$ and $q$, see \cite{Borodzik_Nemethi}. \newpage \input{bib} \end{document}
\section{Acknowledgements} We thank CERN for the very successful operation of the LHC, as well as the support staff from our institutions without whom ATLAS could not be operated efficiently. We acknowledge the support of ANPCyT, Argentina; YerPhI, Armenia; ARC, Australia; BMWF and FWF, Austria; ANAS, Azerbaijan; SSTC, Belarus; CNPq and FAPESP, Brazil; NSERC, NRC and CFI, Canada; CERN; CONICYT, Chile; CAS, MOST and NSFC, China; COLCIENCIAS, Colombia; MSMT CR, MPO CR and VSC CR, Czech Republic; DNRF, DNSRC and Lundbeck Foundation, Denmark; EPLANET, ERC and NSRF, European Union; IN2P3-CNRS, CEA-DSM/IRFU, France; GNSF, Georgia; BMBF, DFG, HGF, MPG and AvH Foundation, Germany; GSRT and NSRF, Greece; ISF, MINERVA, GIF, DIP and Benoziyo Center, Israel; INFN, Italy; MEXT and JSPS, Japan; CNRST, Morocco; FOM and NWO, Netherlands; BRF and RCN, Norway; MNiSW, Poland; GRICES and FCT, Portugal; MNE/IFA, Romania; MES of Russia and ROSATOM, Russian Federation; JINR; MSTD, Serbia; MSSR, Slovakia; ARRS and MIZ\v{S}, Slovenia; DST/NRF, South Africa; MINECO, Spain; SRC and Wallenberg Foundation, Sweden; SER, SNSF and Cantons of Bern and Geneva, Switzerland; NSC, Taiwan; TAEK, Turkey; STFC, the Royal Society and Leverhulme Trust, United Kingdom; DOE and NSF, United States of America. The crucial computing support from all WLCG partners is acknowledged gratefully, in particular from CERN and the ATLAS Tier-1 facilities at TRIUMF (Canada), NDGF (Denmark, Norway, Sweden), CC-IN2P3 (France), KIT/GridKA (Germany), INFN-CNAF (Italy), NL-T1 (Netherlands), PIC (Spain), ASGC (Taiwan), RAL (UK) and BNL (USA) and in the Tier-2 facilities worldwide. \end{document}
\section{Unsolved problems from earlier issues} \begin{issue}Is $\binom{\Omega}{\Gamma}=\binom{\Omega}{\Tau}$?\end{issue \begin{issue}Is $\ufin(\Op,\Omega)=\sfin(\Gamma,\Omega)$?And if not, does $\ufin(\Op,\Gamma)$ imply $\sfin(\Gamma,\Omega)$?\end{issue \stepcounter{issue}\begin{issue}Does $\sone(\Omega,\Tau)$ imply $\ufin(\Gamma,\Gamma)$?\end{issue} \begin{issue}Is $\fp=\fp^*$? (See the definition of $\fp^*$ in that issue.)\end{issue} \begin{issue}Does there exist (in ZFC) an uncountable set satisfying $\sfin(\cB,\cB)$?\end{issue} \stepcounter{issue} \begin{issue}Does $X \nin \NON(\cM)$ and $Y\nin\mathsf{D}$ imply that $X\cup Y\nin \COF(\cM)$?\end{issue} \begin{issue}[CH]Is $\split(\Lambda,\Lambda)$ preserved under finite unions?\end{issue} \begin{issue}Is $\cov(\cM)=\fo$? (See the definition of $\fo$ in that issue.)\end{issue} \stepcounter{issue} \begin{issue}Could there be a Baire metric space $M$ of weight $\aleph_1$ and a partition $\mathcal{U}$ of $M$ into $\aleph_1$ meager sets where for each ${\mathcal U}'\subset\mathcal U$, $\bigcup {\mathcal U}'$ has the Baire property in $M$?\end{issue} \stepcounter{issue} \begin{issue}Does there exist (in ZFC) a set of reals $X$ of cardinality $\fd$ such that all finite powers of $X$ have Menger's property $\sfin(\Op,\Op)$?\end{issue} \begin{issue}Can a Borel non-$\sigma$-compact group be generated by a Hurewicz subspace?\end{issue} \begin{issue}[MA]Is there $X\sbst\bbR$ of cardinality continuum, satisfying $\sone(\BO,\BG)$?\end{issue} \begin{issue}[CH]Is there a totally imperfect $X$ satisfying $\ufin(\Op,\Gamma)$ that can be mapped continuously onto $\Cantor$?\end{issue} \begin{issue}[CH]Is there a Hurewicz $X$ such that $X^2$ is Menger but not Hurewicz?\end{issue} \begin{issue}Does the Pytkeev property of $C_p(X)$ imply that $X$ has Menger's property?\end{issue} \begin{issue}Does every hereditarily Hurewicz space satisfy $\sone(\BG,\BG)$?\end{issue} \begin{issue}[CH]Is there a Rothberger-bounded $G\le\Bgp$ such that $G^2$ is not Menger-bounded?\end{issue} \begin{issue}Let $\cW$ be the van der Waerden ideal. Are $\cW$-ultrafilters closed under products?\end{issue} \begin{issue}Is the $\delta$-property equivalent to the $\gamma$-property $\binom{\Omega}{\Gamma}$?\end{issue} \stepcounter{issue}\stepcounter{issue} \general\end{document}} \newcommand{\fbx}[1]{\fbox{$#1$}}\newcommand{\nop}{$\times$}\newcommand{\fbn}{\!\!\fbox{\!\nop\!}\!\!} \newcommand{\yup}{\checkmark}\newcommand{\fby}{\!\!\fbox{\!\yup\!}\!\!}\newcommand{\mbq}{\mb{?}} \newcommand{\mb}[1]{{\mbox{\textbf{#1}}}}\newcommand{\smb}[1]{{\!\!\mb{#1}\!\!}}\newcommand{\x}{\times} \newcommand{\<}{\left <}\renewcommand{\>}{\right >}\newcommand{\Cantor}{{\{0,1\}^\N}}\newcommand{\oo}{\infty} \newcommand{\NR}{{\bbR^\N}}\newcommand{\Iff}{\Leftrightarrow}\newcommand{\mypar}[1]{\par\medskip\noindent\textbf{#1.}} \newcommand{\roth}{{[\N]^{\alephes}}}\newcommand{\sr}[2]{{\txt{$#1$\\$#2$}}} \newcommand{\gpbl}{{\mbox{\textit{\tiny gp}}}}\newcommand{\fx}{\mathfrak{x}}\newcommand{\fb}{\mathfrak{b}} \newcommand{\fg}{\mathfrak{g}}\newcommand{\fc}{\mathfrak{c}}\newcommand{\fd}{\mathfrak{d}} \newcommand{\fp}{\mathfrak{p}}\newcommand{\fs}{\mathfrak{s}}\newcommand{\ADD}{{\mathsf {ADD}}} \newcommand{\COV}{{\mathsf {COV}}}\newcommand{\NON}{{\mathsf {NON}}}\newcommand{\COF}{{\mathsf {COF}}} \newcommand{\sseq}[1]{\{#1 : n\in\N\}}\newcommand{\Impl}{\Rightarrow} \newcommand{\upannouncement}[1]{[\S\ref{#1} above]}\newcommand{\dnannouncement}[1]{[\S\ref{#1} below]} \newcommand{\E}{\exists}\newcommand{\cI}{\mathcal{I}}\newcommand{\cN}{\mathcal{N}}\newcommand{\cP}{\mathcal{P}} \newcommand{\cA}{\mathcal{A}}\newcommand{\cM}{\mathcal{M}}\newcommand{\Null}{\mathcal{N}} \newcommand{\op}{\operatorname}\newcommand{\cov}{\mathsf{cov}}\newcommand{\add}{\mathsf{add}} \newcommand{\cof}{\mathsf{cof}}\newcommand{\cf}{\mathsf{cf}}\newcommand{\non}{\mathsf{non}}\newcommand{\spst}{\supseteq} \newcommand{\CH}{the Continuum Hypothesis}\newcommand{\bbR}{\mathbb{R}}\newcommand{\Q}{\mathbb{Q}} \newcommand{\EdNote}[1]{\par\medskip\noindent\textbf{#1.}}\newcommand{\fo}{\mathfrak{od}} \newcommand{\cl}[1]{\overline{#1}}\newcommand{\impl}{\rightarrow}\newcommand{\arrays}{{{\{0,1\}}^{\N\x\N}}} \newcommand{\w}{\omega}\newcommand{\ft}{\mathfrak{t}}\newcommand{\h}{\mathfrak{h}}\newcommand{\Cite}[2]{{\cite[#1]{#2}}} \renewcommand{\split}{\mathsf{Split}}\newcommand{\bq}{\begin{quote}}\newcommand{\eq}{\end{quote}} \newcommand{\cK}{\mathcal{K}}\newcommand{\cB}{\mathcal{B}}\newcommand{\BG}{\cB_\Gamma} \newcommand{\BL}{\cB_\Lambda}\newcommand{\BT}{\cB_\Tau}\newcommand{\BTstar}{\cB_{\Tau^*}}\newcommand{\BO}{\cB_\Omega} \newcommand{\CG}{C_\Gamma}\newcommand{\CL}{C_\Lambda}\newcommand{\CT}{C_\Tau}\newcommand{\CTstar}{C_{\Tau^*}} \newcommand{\CO}{C_\Omega}\newcommand{\sone}{\mathsf{S}_1}\newcommand{\sfin}{\mathsf{S}_\mathrm{fin}} \newcommand{\Sc}{\mathsf{S}_c}\newcommand{\ufin}{\mathsf{U}_\mathrm{fin}}\newcommand{\gone}{\mathsf{G}_1} \newcommand{\gfin}{\mathsf{G}_\mathrm{fin}}\newcommand{\seq}[1]{\{#1\}_{n\in\N}}\newcommand{\Union}{\bigcup} \newcommand{\nin}{\not\in}\newcommand{\cF}{\mathcal{F}}\newcommand{\cG}{\mathcal{G}}\newcommand{\cU}{\mathcal{U}} \newcommand{\cV}{\mathcal{V}}\newcommand{\cW}{\mathcal{W}}\newcommand{\fU}{\mathfrak{U}}\newcommand{\fu}{\mathfrak{u}} \newcommand{\fV}{\mathfrak{V}}\newcommand{\fW}{\mathfrak{W}}\newcommand{\psin}{pseudo-intersection} \newcommand{\NN}{{\N^\N}}\newcommand{\N}{\mathbb{N}}\newcommand{\bbN}{\mathbb{N}}\newcommand{\Z}{\mathbb{Z}} \newcommand{\as}{\subseteq^*}\newcommand{\sm}{\setminus}\newcommand{\sbst}{\subseteq} \newcommand{\by}[2]{\par\hfill\emph{#1}, #2}\newcommand{\nby}[1]{\par\hfill\emph{#1}}\newcommand{\Tau}{\mathrm{T}} \newcommand{\CE}{\textsc{CE}} \newtheorem{thm}{Theorem}[section]\newcommand{\bthm}{\begin{thm}} \newcommand{\ethm}{\end{thm}} \newtheorem{prop}[thm]{Proposition}\newcommand{\bprp}{\begin{prop}} \newcommand{\eprp}{\end{prop}} \newtheorem{fact}[thm]{Fact}\newcommand{\bfct}{\begin{fact}} \newcommand{\efct}{\end{fact}} \newtheorem{prob}[thm]{Problem}\newcommand{\bprb}{\begin{prob}} \newcommand{\eprb}{\end{prob}} \newtheorem{lem}[thm]{Lemma}\newcommand{\blem}{\begin{lem}} \newcommand{\elem}{\end{lem}} \newtheorem{claim}[thm]{Claim}\newcommand{\bclm}{\begin{claim}} \newcommand{\eclm}{\end{claim}} \newtheorem{cor}[thm]{Corollary}\newcommand{\bcor}{\begin{cor}} \newcommand{\ecor}{\end{cor}} \newtheorem{conj}[thm]{Conjecture}\newcommand{\bcnj}{\begin{conj}} \newcommand{\ecnj}{\end{conj}} \theoremstyle{definition}\newtheorem{defn}[thm]{Definition}\newcommand{\bdfn}{\begin{defn}} \newcommand{\edfn}{\end{defn}} \theoremstyle{remark}\newtheorem{rem}[thm]{Remark}\newcommand{\brem}{\begin{rem}} \newcommand{\erem}{\end{rem}} \newtheorem{cnv}[thm]{Convention}\newcommand{\bcnv}{\begin{cnv}} \newcommand{\ecnv}{\end{cnv}} \newtheorem{exam}[thm]{Example}\newcommand{\bexm}{\begin{exam}} \newcommand{\eexm}{\end{exam}} \newtheorem{issue}{Issue}\newcommand{\bpf}{\begin{proof}} \newcommand{\epf}{\end{proof}} \newcommand{\be}{\begin{enumerate}}\newcommand{\ee}{\end{enumerate}}\newcommand{\bi}{\begin{itemize}} \newcommand{\ei}{\end{itemize}}\newcommand{\itm}{\item} \newcommand{\general}{\small\vfill\par\noindent\hrulefill\par \noindent\textbf{Previous issues.} The previous issues of this bulletin are available online at\\ \url{http://front.math.ucdavis.edu/search?\&t=\%22SPM+Bulletin\%22} \\[0.1cm] \textbf{Contributions.} Announcements, discussions, and open problems should be emailed to \texttt{<EMAIL>}\\[0.1cm] \textbf{Subscription.} To receive this bulletin (free) to your e-mailbox, e-mail us. } \newcommand{\link}[1]{\par\hfill{\texttt{#1}}} \newcommand{\fillin}{{\Huge To be completed}} \newcommand{\arXivl}[4]{\subsection{#2}{#4}\par\hfill{\arx{#1}}\par\hfill\emph{#3}} \newcommand{\arXiv}[3]{\subsection{#2}\mbox{}\par\hfill{\arx{#1}}\par\hfill\emph{#3}} \newcommand{\DOIpaper}[5]{\subsection{#2}{#4}\par\hfill{\texttt{http://dx.doi.org/#1}}\par\hfill\emph{#3}} \newcommand{\AMSPaper}[5]{\subsection{#3}{#5}\par\hfill{\texttt{#1}}\par\hfill\emph{#4}\par\hfill{#2}} \newcommand{\nAMSPaper}[4]{\subsection{#2}{#4}\par\hfill{\texttt{#1}}\par\hfill\emph{#3}} \newcommand{\AMS}[3]{\subsection{#1}\mbox{}\par\hfill{\texttt{#3}}\par\hfill\emph{#2}} \newcommand{\SPMBul}{\textbf{$\mathcal{SPM}$ Bulletin}} \newcommand{\BulEnd}{\par\bigskip\noindent Boaz Tsaban\\ \emph{E-mail}: tsaban@math.biu.ac.il\\ \emph{URL}: http://www.cs.biu.ac.il/\~{}tsaban} \newcommand{\arx}[1]{\url{http://arxiv.org/abs/#1}} \newcommand{\probissue}{\emph{Problem of the issue}} \title[$\mathcal{SPM}$ Bulletin \textbf{\issuenumber} (\issuemonth{} \issueyear)] $\mathcal{SPM}$ Bulletin\\[0.5cm] Issue number \issuenumber: \issuemonth{} \issueyear{} \CE{}} \begin{document} \maketitle \forget \section{Editor's note} \medskip With best regards, \by{Boaz Tsaban}{<EMAIL>} \hfill \texttt{http://www.cs.biu.ac.il/\~{}tsaban} \forgotten \section{Proceedings of the Fourth Workshop on Coverings, Selections and Games in Topology} Volume 160, Issue 18, of Topology and its Applications (1 December 2013, Pages 2233--2566), is dedicated to the proceedings of the Fourth Workshop on Coverings, Selections and Games in Topology, on the occasion of Ljubi\v{s}a D.R. Ko\v{c}inac's 65$^\mathrm{th}$ birthday. This issue is now available online at \centerline{\url{http://www.sciencedirect.com/science/journal/01668641/160/18}} Contents: \be \item \emph{Preface}, Page 2233, G. Di Maio, B. Tsaban. \item \emph{The mathematics of Ljubi\v{s}a D.R. Ko\v{c}inac}, Pages 2234--2242, B. Tsaban. \item \emph{Selectively c.c.c.\ spaces}, Pages 2243-2250, L.F. Aurichi. \item \emph{Weak covering properties and selection principles}, Pages 2251--2271, L. Bab\-inkostova, B.A. Pansera, M. Scheepers. \item \emph{Detecting topological groups which are (locally) homeomorphic to LF-spaces}, Pages 2272--2284, T. Banakh, K. Mine, D. Repov\v{s}, K. Sakai, T. Yagasaki. \item \emph{Gap topologies in metric spaces}, Pages 2285-2308, G. Beer, C. Costantini, S. Levi. \item \emph{On two selection principles and the corresponding games}, Pages 2309--2313, A. Bella. \item \emph{Scale function vs topological entropy}, Pages 2314--2334, F. Berlai, D. Dikranjan, A. Giordano Bruno. \item \emph{On diagonal resolvable spheres}, Pages 2335--2339, P.V.M. Blagojevi\'c, A.D. Blagojevi\'c, L.D.R. Ko\v{c}inac. \item \emph{The Schwarz genus of the Stiefel manifold}, Pages 2340--2350, P.V.M. Blagojevi\'c, R.N. Karasev. \item \emph{Dowker-type example and Arhangelskiiʼs image-property}, Pages 2351--2355, M. Bonanzinga, M. Matveev. \item \emph{Modifications of sequence selection principles}, Pages 2356--2370, L. Bukovsk\'y, J. \v{S}upina. \item \emph{On the cardinality of the $\theta$-closed hull of sets}, Pages 2371--2378, F. Cammaroto, A. Catalioto, B.A. Pansera, B. Tsaban. \item \emph{Variations on selective separability in non-regular spaces}, Pages 2379--2385, A. Caserta, G. Di Maio. \item \emph{Some generalizations of Backʼs Theorem}, Pages 2386--2395, A. Caterino, R. Ceppitelli, L. Hol\'a. \item \emph{On Haar meager sets}, Pages 2396--2400, U.B. Darji. \item \emph{Some further results on image-γ and image-image-covers}, Pages 2401--2410, P. Das, D. Chandra. \item \emph{Indestructibility of compact spaces}, Pages 2411--2426, R.R. Dias, F.D. Tall. \item \emph{On characterized subgroups of compact abelian groups}, Pages 2427--2442, D. Dikranjan, S.S. Gabriyelyan. \item \emph{Productively Lindelöf and indestructibly Lindelöf spaces}, Pages 2443--2453, H. Duanmu, F.D. Tall, L. Zdomskyy. \item \emph{Universal frames}, Pages 2454--2464, T. Dube, S. Iliadis, J. van Mill, I. Naidoo. \item \emph{Continuous weak selections for products}, Pages 2465--2472, S. Garc\'{\i}a--Ferreira, K. Miyazaki, T. Nogura. \item \emph{Selections and metrisability of manifolds}, Pages 2473--2481, D. Gauld. \item \emph{On base dimension-like functions of the type Ind}, Pages 2482--2494, D.N. Georgiou, S.D. Iliadis, A.C. Megaritis. \item \emph{Selection properties of uniform and related structures}, Pages 2495--2504, L.D.R. Ko\v{c}inac, H.P.A. K\"unzi. \item \emph{A characterization of the Menger property by means of ultrafilter convergence}, Pages 2505--2513, P. Lipparini. \item \emph{Transfinite extension of dimension function image}, Pages 2514--2522, N.N. Martynchuk. \item \emph{Homogeneity and h-homogeneity}, Pages 2523--2530, S.V. Medvedev. \item \emph{The weak Hurewicz property of Pixley–Roy hyperspaces}, Pages 2531--2537, M. Sakai. \item \emph{Comparing weak versions of separability}, Pages 2538--2566, D.T. Soukup, L. Soukup, S. Spadaro. \ee We use this opportunity to thank the Chief Editors and the referees for their tremendous help. \nby{Giuseppe Di Maio, Boaz Tsaban} \hfill{Guest Editors, Topology and its Applications} \section{Long announcements} \arXivl{1304.6628} {Completeness and related properties of the graph topology on function spaces} {Lubica Hol\'a, L\'aszl\'o Zsilinszky} {The graph topology $\tau_{\Gamma}$ is the topology on the space $C(X)$ of all continuous functions defined on a Tychonoff space $X$ inherited from the Vietoris topology on $X\times \mathbb R$ after identifying continuous functions with their graphs. It is shown that all completeness properties between complete metrizability and hereditary Baireness coincide for the graph topology if and only if $X$ is countably compact; however, the graph topology is $\alpha$-favorable in the strong Choquet game, regardless of $X$. Analogous results are obtained for the fine topology on $C(X)$. Pseudocompleteness, along with properties related to 1st and 2nd countability of $(C(X),\tau_{\Gamma})$ are also investigated.} \arXivl{1306.5463} {Topological games and Alster spaces} {Leandro F. Aurichi and Rodrigo R. Dias} {In this paper we study connections between topological games such as Rothberger, Menger and compact-open, and relate these games to properties involving covers by $G_{\delta}$ subsets. The results include: (1) If Two has a winning strategy in the Menger game on a regular space $X$, then $X$ is an Alster space. (2) If Two has a winning strategy in the Rothberger game on a topological space $X$, then the $G_\delta$-topology on $X$ is Lindelof. (3) The Menger game and the compact-open game are (consistently) not dual.} \arXivl{1306.5421} {Strongly Summable Ultrafilters, Union Ultrafilters, and the Trivial Sums Property} {David J. Fern\'andez Bret\'on} {We answer two questions of Hindman, Stepr\=ans and Strauss, namely we prove that every strongly summable ultrafilter on an abelian group is sparse and has the trivial sums property. Moreover we show that in most cases the sparseness of the given ultrafilter is a consequence of its being isomorphic to a union ultrafilter. However, this does not happen in all cases: we also construct (assuming $\cov(\cM)=\fc$), on the Boolean group, a strongly summable ultrafilter that is not additively isomorphic to any union ultrafilter.} \arXivl{1307.7928}{Topological games and productively countably tight spaces} {Leandro F. Aurichi and Angelo Bella} {The two main results of this work are the following: if a space $X$ is such that player II has a winning strategy in the game $\gone(\Omega_x, \Omega_x)$ for every $x \in X$, then $X$ is productively countably tight. On the other hand, if a space is productively countably tight, then $\sone(\Omega_x, \Omega_x)$ holds for every $x \in X$. With these results, several other results follow, using some characterizations made by Uspenskii and Scheepers.} \arXivl{1310.0409} {On a game theoretic cardinality bound} {Leandro F. Aurichi and Angelo Bella} {The main purpose of the paper is the proof of a cardinal inequality for a space with points $G_\delta$, obtained with the help of a long version of the Menger game. This result improves a similar one of Scheepers and Tall.} \arXivl{1311.1468} {Pytkeev $\aleph_0$-spaces} {Taras Banakh} {A regular topological space $X$ is defined to be a Pytkeev $\aleph_0$-space if it has countable Pytkeev network. A family $\mathcal P$ of subsets of a topological space $X$ is called a Pytkeev network in $X$ if for a subset $A\subset X$, a point $x\in{\bar A}$ and a neighborhood $O_x\subset X$ there is a set $P\in\mathcal P$ such that $x\in P\subset O_x$ and moreover $P\cap A$ is infinite if $x$ is an accumulation point of $A$. The class of Pytkeev $\aleph_0$-spaces contains all metrizable separable spaces and is (properly) contained in the class of $\aleph_0$-spaces. This class is closed under many operations over topological spaces: taking a subspace, countable Tychonoff product, small countable box-product, countable direct limit, the hyperspace. For an $\aleph_0$-space $X$ and a Pytkeev $\aleph_0$-space $Y$ the function space $C_k(X,Y)$ endowed with the compact-open topology is a Pytkeev $\aleph_0$-space. A topological space is second countable if and only if it is Pytkeev $\aleph_0$-space with countable fan tightness.} \arXivl{1310.8622} {Selective covering properties of product spaces, II: $\gamma$ spaces} {Arnold W. Miller, Boaz Tsaban, Lyubomyr Zdomskyy} {We study productive properties of $\gamma$ spaces, and their relation to other, classic and modern, selective covering properties. Among other things, we prove the following results: \be \item Solving a problem of F. Jordan, we show that for every unbounded tower set $X\sub\R$ of cardinality $\aleph_1$, the space $\Cp(X)$ is productively Fr\'echet--Urysohn. In particular, the set $X$ is productively $\gamma$. \item Solving problems of Scheepers and Weiss, and proving a conjecture of Bab\-inkostova--Scheepers, we prove that, assuming \CH{}, there are $\gamma$ spaces whose product is not even Menger. \item Solving a problem of Scheepers--Tall, we show that the properties $\gamma$ and Gerlits--Nagy (*) are preserved by Cohen forcing. Moreover, every Hurewicz space that Remains Hurewicz in a Cohen extension must be Rothberger (and thus (*)). \ee We apply our results to solve a large number of additional problems, and use Arhangel'\-ski\u{\i} duality to obtain results concerning local properties of function spaces and countable topological groups.} \arXivl{1311.2011} {Productively countably tight spaces of the form $C_k(X)$} {Leandro Fiorini Aurichi, Renan Maneli Mezabarba} {Some results in $C_k$-theory are obtained with the use of bornologies. We investigate under which conditions the space of the continuous real functions with the compact-open topology is a productively countably tight space, which yields some applications on Alster spaces.} \section{Short announcements}\label{RA} \arXiv{1304.0472} {Partitioning bases of topological spaces} {Daniel T. Soukup, Lajos Soukup} \arXiv{1304.0486} {Compactness of $\omega^\lambda$} {Paolo Lipparini} \arXiv{1304.2042} {Non-meager free sets for meager relations on Polish spaces} {Taras Banakh and Lyubomyr Zdomskyy} \AMS{On strong $P$-points} {Andreas Blass; Michael Hrusak; Jonathan Verner} {http://www.ams.org/journal-getitem?pii=S0002-9939-2013-11518-2} \arXiv{1307.0184} {Products and countable dense homogeneity} {Andrea Medini} \arXiv{1307.1989} {Strong colorings yield kappa-bounded spaces with discretely untouchable points} {Istvan Juhasz and Saharon Shelah} \arXiv{1310.1827}{Selective and Ramsey ultrafilters on $G$-spaces} {O.V. Petrenko, I.V. Protasov} \arXiv{1311.1677} {Seven characterizations of non-meager P-filters} {Kenneth Kunen, Andrea Medini, Lyubomyr Zdomskyy} \ed
\section[]{Introduction} \label{sec:intro} The presence of a diffuse population of intergalactic stars in galaxy clusters was first proposed by \citet{zwicky37}, and later confirmed by the same author using observations of the Coma cluster with a 48-inch schmidt telescope \citep{zwicky52}. More recent observational studies have confirmed that a substantial fraction of stars in clusters are not bound to galaxies. This diffuse component is generally referred to as Intra-Cluster Light (hereafter ICL). Both from the observational and the theoretical point of view, it is not trivial to define the ICL component. A fraction of central cluster galaxies are characterized by a faint and extended stellar halo. These galaxies are classified as \emph{cD galaxies}, where `c' refers to the fact that these galaxies are very large and stands for supergiant and `D' for diffuse (\citealt{matthews64}), to highlight the presence of a diffuse stellar envelope made of stars that are not bound to the galaxy itself. Separating these two components is not an easy task. On the observational side, some authors use an isophotal limit to cut off the light from satellite galaxies, while the distinction between the brightest cluster galaxy (hereafter BCG) and ICL is based on profile decomposition \citep[e.g.][]{zibetti}. Others \citep[e.g.][]{gonzalez05} rely on two-dimensional profile fittings to model the surface brightness profile of brightest cluster galaxies. In the framework of numerical simulations, additional information is available, and the ICL component has been defined using either a binding energy definition \citep[i.e. all stars that are not bound to identified galaxies, e.g.][]{giuseppe}, or variations of this technique that take advantage of the dynamical information provided by the simulations \citep[e.g.][]{Dolag_etal_2010}. In a recent work, \citet{rudick11} discuss different methods that have been employed both for observational and for theoretical data, and apply them to a suite of N-body simulations of galaxy clusters. They find that different methods can change the measured fraction\footnote{This is the ratio between the mass or luminosity in the ICL component and the total stellar mass or luminosity enclosed within some radius, usually $R_{200}$ or $R_{500}$. In this work, we will use $R_{200}$, defined as the radius that encloses a mean density of 200 times the critical density of the Universe at the redshift of interest.} of ICL by up to a factor of about four (from $\sim 9$ to $\sim 36$ per cent). In contrast, \cite{puchwein} apply four different methods to identify the ICL in hydrodynamical SPH simulations of cluster galaxies, and consistently find a significant ICL stellar fraction ($\sim$ 45 per cent). There is no general agreement in the literature about how the ICL fraction varies as a function of cluster mass. \citet{zibetti} find that richer clusters (the richness being determined by the number of red-sequence galaxies), and those with a more luminous BCG have brighter ICL than their counterparts. However, they find roughly constant ICL fractions as a function of halo mass, within the uncertainties and sample variance. In contrast, \citet{lin} empirically infer an increasing fraction of ICL with increasing cluster mass. To estimate the amount of ICL, they use the observed correlation between the cluster luminosity and mass and a simple merger tree model for cluster formation. Results are inconclusive also on the theoretical side, with claims of increasing ICL fractions for more massive haloes (e.g. \citealt{giuseppe04,purcell07,giuseppe,purcell08}), as well as findings of no significant increase of the ICL fraction with cluster mass (e.g. \citealt{pigi,henriques,puchwein}), at least for systems more massive than $10^{13} \, M_{\odot}/h$. Different physical mechanisms may be at play in the formation of the ICL, and their relative importance can vary during the dynamical history of the cluster. Stars can be stripped away from satellite galaxies orbiting within the cluster, by tidal forces exerted either during interactions with other cluster galaxies, or by the cluster potential. This is supported by observations of arclets and similar tidal features that have been identified in the Coma, Centaurus and Hydra I clusters \citep{gregg, trentham, calcaneo, arnaboldi12}. As pointed out by several authors, in a scenario where galaxy stripping and disruption are the main mechanisms for the production of the ICL, the major contribution comes from galaxies falling onto the cluster along almost radial orbits, since tidal interactions by the cluster potential are strongest for these galaxies. Numerical simulations have also shown that large amounts of ICL can come from `pre-processing' in galaxy groups that are later accreted onto massive clusters \citep{mihos,willman,rudick06,sommer}. In addition, \citet{giuseppe} found that the formation of the ICL is tightly linked to the build-up of the BCG and of the other massive cluster galaxies, a scenario supported by other theoretical studies (e.g. \citealt{diemand,abadi,font,read}). It is important, however, to consider that results from numerical simulations might be affected by numerical problems. \citet{giuseppe} find an increasing fraction of ICL when increasing the numerical resolution of their simulations. In addition, \citet{puchwein} show that a significant fraction ($\sim 30$ per cent) of the ICL identified in their simulations forms in gas clouds that were stripped from the dark matter haloes of galaxies infalling onto the cluster. Fluid instabilities, that are not well treated within the SPH framework, might be able to disrupt these clouds suppressing this mode of ICL formation. In this paper, we use the semi-analytic model presented in \citet[][ hereafter DLB07]{dlb}, that we extend by including three different prescriptions for the formation of the ICL. We couple this model to a suite of high-resolution N-body simulations of galaxy clusters to study the formation and evolution of the ICL component, as well as its physical properties, and the influence of the updated prescriptions on model basic predictions (in particular, the galaxy stellar mass function, and the mass of the BCGs). There are some advantages in using semi-analytic models to describe the ICL formation with respect to hydrodynamical simulations: they do not suffer from numerical effects related to the fragility of poorly resolved galaxies, and allow the relative influence of different channels of ICL generation to be clearly quantified. However, the size and abundance of satellite galaxies (that influence the amount of predicted ICL) might be estimated incorrectly in these models. We will comment on these issues in the following. The layout of the paper is as follows. In Section~\ref{sec:sim} we introduce the simulations used in our study, and in Section~\ref{sec:model} we describe the prescriptions we develop to model the formation of the ICL component. In Section~\ref{sec:massfunction} we discuss how our prescriptions affect the predicted galaxy stellar mass function, and in Section~\ref{sec:haloprop} we discuss how the predicted fraction of ICL varies as a function of halo properties. In Section~\ref{sec:formation}, we analyse when the bulk of the ICL is formed, and which galaxies provide the largest contribution. We then study the correlation between the ICL and the properties of the corresponding BCGs in Section~\ref{sec:bcgprop}, and analyse the metal content of the ICL in Section~\ref{sec:metallicity}. Finally, we discuss our results and give our conclusions in Section~\ref{sec:discussion}. \section[]{N-body simulations} \label{sec:sim} In this study we use collisionless simulations of galaxy clusters, generated using the `zoom' technique \citep*[][ see also \citealt{Katz_and_White_1993}]{Tormen_etal_1997}: a target cluster is selected from a cosmological simulation and all its particles, as well as those in its immediate surroundings, are traced back to their Lagrangian region and replaced with a larger number of lower mass particles. Outside this high-resolution region, particles of increasing mass are displaced on a spherical grid. All particles are then perturbed using the same fluctuation field used in the parent cosmological simulations, but now extended to smaller scales. The method allows the computational effort to be concentrated on the cluster of interest, while maintaining a faithful representation of the large scale density and velocity. Below, we use 27 high-resolution numerical simulations of regions around galaxy clusters, carried out assuming the following cosmological parameters: $\Omega_m=0.24$ for the matter density parameter, $\Omega_{\rm bar}=0.04$ for the contribution of baryons, $H_0=72\,{\rm km\,s^{-1}Mpc^{-1}}$ for the present-day Hubble constant, $n_s=0.96$ for the primordial spectral index, and $\sigma_8=0.8$ for the normalization of the power spectrum. The latter is expressed as the r.m.s. fluctuation level at $z=0$, within a top-hat sphere of $8\,\hm$Mpc radius. For all simulations the mass of each Dark Matter particle in the high resolution region of $10^8\,\hm {\rm M}_{\odot}$, and a Plummer-equivalent softening length is fixed to $\epsilon=2.3 \hm$~kpc in physical units at $z<2$, and in comoving units at higher redshift. Simulation data have been stored at 93 output times, between $z=60$ and $z=0$. Dark matter haloes have been identified using a standard friends-of-friends (FOF) algorithm, with a linking length of 0.16 in units of the mean inter-particle separation in the high-resolution region. The algorithm {\small SUBFIND} \citep{Springel_etal_2001} has then been used to decompose each FOF group into a set of disjoint substructures, identified as locally overdense regions in the density field of the background halo. As in previous work, only substructures that retain at least 20 bound particles after a gravitational unbinding procedure are considered to be genuine substructures. Finally, merger histories have been constructed for all self-bound structures in our simulations, using the same post-processing algorithm that has been employed for the Millennium Simulation \citep{springel3}. For more details on the simulations, as well as on their post-processing, we refer the reader to \citet{myself}. For our analysis, we use a sample of 341 haloes extracted from the high resolution regions of these simulations, with mass larger than $10^{13} \hm M_{\odot}$. In Table~\ref{tab:tab0}, we give the number of haloes in different mass ranges. \begin{table} \caption{The sample of haloes used in this study, split in five subsamples according to their mass. The first column indicates the mass range, and the second column gives the number of haloes in each subsample.} \begin{center} \begin{tabular}{lllll} \hline Halo mass range & Number of haloes \\ \hline $ \geq 10^{15}\,\hm {\rm M}_{\odot}$ & 13 \\ $[$5-10$]\times10^{14}\,\hm {\rm M}_{\odot}$ & 15 \\ $[$1-5$]\times10^{14}\,\hm {\rm M}_{\odot}$ & 25 \\ $[$5-10$]\times10^{13}\,\hm {\rm M}_{\odot}$ & 29 \\ $[$1-5$]\times10^{13}\,\hm {\rm M}_{\odot}$ & 259 \\ \hline \end{tabular} \end{center} \label{tab:tab0} \end{table} \section[]{Semi-analytic models for the formation of the ICL} \label{sec:model} In this study, we use the semi-analytic model presented in DLB07, but we update it in order to include three different prescriptions for modelling the formation of the ICL component. These prescriptions are described in detail in the following. For readers who are not familiar with the terminology used within our model, we recall that we consider three different types of galaxies: \begin{itemize} \item Type 0: these are central galaxies, defined as those located at the centre of the main halo\footnote{This is the most massive subhalo of a FOF, and typically contains about 90 per cent of its total mass.} of each FOF group; \item Type 1: these are satellite galaxies associated with a distinct dark matter substructure. A Type 0 galaxy becomes Type 1 once its parent halo is accreted onto a more massive system; \item Type 2: also called \emph{orphan} galaxies, these are satellites whose parent substructures have been stripped below the resolution limit of the simulation. Our reference model assumes that, when this happens, the baryonic component is unaffected and the corresponding galaxy survives for a residual time before merging with the corresponding central galaxy. The position of an orphan galaxy is traced by following the position of the particle that was the most bound particle of the parent substructure at the last time it was identified. \end{itemize} The residual merger time assigned to each galaxy that becomes Type 2 is estimated using the following implementation of the Chandrasekhar dynamical friction formula: \begin{equation}\label{dff} \tau_{merge} = f_{fudge}\frac{1.17}{\ln \Lambda}\frac{D^2}{R^2_{vir}}\frac{M_{par}}{M_{sat}}\tau_{dyn} \end{equation} where $D$ is the distance between the satellite and the centre of its parent FOF, $R_{vir}$ is the virial radius of the parent halo, $M_{sat}$ the sum of the dark and baryonic mass of the satellite, $M_{par}$ the (dark matter) mass of the accreting halo, $\tau_{dyn} = R_{vir}/V_{vir}$ is the dynamical time of the parent halo, and $\Lambda = 1+M_{par}/M_{sat}$ is the Coulomb logarithm. All quantities entering equation \ref{dff} are evaluated at the last time the substructure hosting the satellite galaxy is identified, before falling below the mass limit for substructure identification. As in DLB07, we have assumed $f_{fudge} = 2$, which is in better agreement with recent numerical work indicating that the classical dynamical friction formulation tends to under-estimate the merging times measured from simulations \citep{bk, Jiang_etal_2008}. The reference model does not include a prescription for the formation of the ICL. Below, we describe three different models \footnote{That are ''either/or'' prescriptions.} that we have implemented to account for the ICL component. We assume that it is formed through two different channels: (i) stellar stripping from satellite galaxies and (ii) relaxation processes that take place during mergers \footnote{That we couple with each of the three stellar stripping prescriptions.} and may unbind some fraction of the stellar component of the merging galaxies. In the following, we describe in detail each of our prescriptions. \subsection[]{Disruption Model} This model is equivalent to that proposed by \cite{qiguo}, and assumes that the stellar component of satellite galaxies is affected by tidal forces only after their parent substructures have been stripped below the resolution of the simulation (i.e. the galaxies are Type 2). We assume that each satellite galaxy orbits in a singular isothermal potential, \begin{equation} \phi(R) = V_{vir}^2 \ln R, \end{equation} and assume the conservation of energy and angular momentum along the orbit to estimate its pericentric distance: \begin{equation} \left(\frac{R}{R_{peri}}\right)^2 = \frac{\ln R/R_{peri} + \frac{1}{2} (V/V_{vir})^2}{\frac{1}{2} (V_t/V_{vir})^2} . \end{equation} In the equation above, $R$ is the distance of the satellite from the halo centre, and $V$ and $V_t$ are the velocity of the satellite with respect to the halo centre and its tangential part, respectively. Following \citet{qiguo}, we compare the main halo density at pericentre with the average baryon mass density (i.e. the sum of cold gas mass and stellar mass) of the satellite within its half mass radius. Then, if the following condition is verified: \begin{equation}\label{eq2} \frac{M_{DM, halo}(R_{peri})}{R_{peri}^3} = \rho_{DM,halo} > \rho_{sat} = \frac{M_{sat}}{R_{half}^3} \, , \end{equation} we assume the satellite galaxy to be disrupted and its stars to be assigned to the ICL component of the central galaxy. In the equation above, we approximate $R_{half}$ by the mass weighted average of the half mass radius of the disk and the half mass radius of the bulge, and $M_{sat}$ is the baryonic mass (cold gas plus stellar mass). The cold \footnote{In our model, no hot component is associated with satellite galaxies.} gas mass that is associated with the disrupted satellite is added to the hot component of the central galaxy. When a central Type 0 galaxy is accreted onto a larger system and becomes a Type 1 satellite, it carries its ICL component until its parent substructure is stripped below the resolution limit of the simulation. At this point, its ICL is added to that of the new central galaxy. In a recent paper, \cite{alvaro} discuss the limits of this implementation with respect to results from controlled numerical simulations of the evolution of disk galaxies within a group environment. The model discussed above is applied only after the galaxy's dark matter subhalo has been completely disrupted, but the simulations by Villalobos et al. show that the stellar component of the satellite galaxy can be significantly affected by tidal forces when its parent subhalo is still present. In addition, this model assumes that the galaxy is completely destroyed when equation (\ref{eq2}) is satisfied, but the simulations mentioned above show that galaxies can survive for a relatively long time (depending on their initial orbit) after they start feeling the tidal forces exerted by the cluster potential. Predictions from this model are affected by numerical resolution. We carried out a convergence test by using a set of low-resolution simulations with the same initial conditions of the high-resolution set used in this paper, but with the dark matter particle mass one order of magnitude larger than the one adopted in the high resolution set and with gravitational softening increased accordingly by a factor $10^{1/3}$. We find that, on average, the ICL fraction is $\sim 30$ per cent higher in the low-resolution set ($\sim 20$ per cent in group-like haloes with mass $\sim 10^{13} M_{\odot} \hm$ and $\sim 50$ in the most massive haloes considered in our study, as shown in the right panel of Figure \ref{fig:ICLconv} in Appendix \ref{sec:numconv}). This is due to the fact that, decreasing the resolution, a larger fraction of satellite galaxies are classified as Type 2 and are subject to our stripping model. While the lack of numerical convergence does not affect the qualitative conclusions of our analysis, we point out that the amount of ICL measured in our simulations should be regarded as an upper limit. \subsection[]{Tidal Radius Model} In this prescription, we allow each satellite galaxy to lose mass in a continuous fashion, before merging or being totally destroyed. Assuming that the stellar density distribution of each satellite can be approximated by a spherically symmetric isothermal profile, we can estimate the \emph{tidal radius} by means of the equation: \begin{equation} R_{t} = \left(\frac{M_{sat}}{3 \cdot M_{DM,halo}}\right)^{1/3} \cdot D \end{equation} \citep{binney}. In the above equation, $M_{sat}$ is the satellite mass (stellar mass + cold gas mass), $M_{DM,halo}$ is the dark matter mass of the parent halo, and $D$ the satellite distance from the halo centre. In our model, a galaxy is a two-component system with a spheroidal component (the bulge), and a disk component. If $R_t$ is smaller than the bulge radius, we assume the satellite to be completely disrupted and its stellar and cold mass to be added to the ICL and hot component of the central galaxy, respectively. If $R_t$ is larger than the bulge radius but smaller than the disk radius, we assume that the mass in the shell $R_t -R_{sat}$ is stripped and added to the ICL component of the central galaxy. A proportional fraction of the cold gas in the satellite galaxy is moved to the hot component of the central galaxy. We assume an exponential profile for the disk, and $R_{sat} = 10 \cdot R_{sl}$, where $R_{sl}$ is the disk scale length ($R_{sat}$ thus contains 99.9 per cent of the disk stellar mass). After a stripping episode, the disk scale length is updated to one tenth of the tidal radius. This prescription is applied to both kinds of satellite galaxies. For Type 1 galaxies, we derive the tidal radius including the dark matter component in $M_{sat}$, and we impose that stellar stripping can take place only if the following condition is verified: \begin{equation}\label{eqn:eq_radii} R^{DM}_{half} < R^{Disk}_{half} \, , \end{equation} where $R^{DM}_{half}$ is the half-mass radius of the parent subhalo, and $R^{Disk}_{half}$ the half-mass radius of the galaxy's disk, that is $1.68\cdot R_{sl}$ for an exponential profile. When a Type 1 satellite is affected by stellar stripping, the associated ICL component is added to that of the corresponding central galaxy. As for the Disruption model, predictions from the Tidal Radius model are affected by numerical resolution. We carried out the same convergence test used for the Disruption model. Again, we find for the low resolution set a larger amount of ICL (by about 40 per cent, almost independent of halo mass, as shown in the left panel of Figure \ref{fig:ICLconv} in Appendix \ref{sec:numconv}). In this model, Type 2 galaxies are the dominant contributors to the ICL and, as for the Disruption model, the larger number of these satellites in the low-resolution set translates in a larger number of galaxies eligible for tidal stripping. \subsection[]{Continuous Stripping Model} \label{sec:modc} This model is calibrated on recent numerical simulations by \citet{alvaro}. These authors have carried out a suite of numerical simulations aimed to study the evolution of a disk galaxy within the global tidal field of a group environment (halo mass of about $10^{13}\,{\rm M}_{\odot}$). In the simulations, both the disk galaxy and the group are modelled as multi-component systems composed of dark matter and stars. The evolution of the disk galaxy is followed after it crosses the group virial radius, with initial velocity components consistent with infalling substructure from cosmological simulations (see \citealt{benson}). The simulations cover a broad parameter space and allow the galaxy-group interaction to be studied as a function of orbital eccentricity, disk inclination, and galaxy-to-group mass ratio. We refer to the original paper for a more detailed description of the simulations set-up, and of the results. Analysing the outputs of these simulations (Villalobos et al, in preparation), we have derived a fitting formula that describes the evolution of the stellar mass lost by a satellite galaxy as a function of quantities estimated at the time of accretion (i.e. at the time the galaxy crosses the virial radius of the group). Our fitting formula reads as follow: \begin{equation} M^*_{lost} = M^*_{accr}\exp\left[\left(\frac{-16}{1-\eta}\right)\left(\frac{M_{sub}}{M_{par}}\right)^{\frac{1}{2}}\left(1-\frac{t}{t_{merg}} \right) \right] \label{eq:modc} \end{equation} where $M^*_{accr}$ is the stellar mass at the time of accretion, $\eta$ is the circularity of the orbit, $M_{sub}$ and $M_{par}$ are the subhalo and parent halo dark matter masses respectively, and $t_{merg}$ is the residual merger time of the satellite galaxies. We approximate the accretion time as the last time the galaxy was a central galaxy (a Type 0), and compute the circularity using the following equation: \begin{equation}\label{eta} \eta = V_{\theta}\sqrt{\frac{2f-V^2_r -V^2_{\theta}}{2f-1}}, \end{equation} where $V_r$ and $V_{\theta}$ are the radial and tangential velocities of the accreted subhalo, and $f=1+M_{sub}/M_{par}$. For each accreted galaxy, $V_r$ and $V_{\theta}$ are extracted randomly from the distributions measured by \citet{benson} from numerical simulations. Since Eq.~\ref{eq:modc} is estimated at the time of accretion, we cannot use the merger time prescription that is adopted in the reference model, where a residual merger time is assigned only at the time the substructure is stripped below the resolution of the simulation. To estimate merger times at the time of accretion, we use the fitting formula by \cite{bk} (hereafter BK08), that has been calibrated at the time the satellite galaxy crosses the virial radius of the accreting system, and is therefore consistent with our Eq.~\ref{eq:modc}. In a few cases, it happens that the merger time is elapsed when the satellite galaxy is still a Type 1. In this case, we do not allow the galaxy to merge before it becomes a Type 2. \citet{gab2010} compared the merger times predicted using this formula with no orbital dependency and for circular orbits with those provided by Equation \ref{dff} used in the reference model, and found a relatively good agreement. We find that, once the orbital dependency is accounted for, the BK08 fitting formula predicts merger times that are on average shorter than those used in our reference model by a factor of $\sim 3$ for mass-ratios $M_{sub}/M_{par} > 0.025$. Therefore, in our continuous stripping model, mergers are on average shorter than in the disruption and tidal radius models. To implement the model described in this section for the formation of the ICL, we use Eq.~\ref{eq:modc} to compute how much stellar mass has to be removed from each satellite galaxy at each time-step. The stripped stars are then added to the ICL component of the corresponding central galaxy, and a proportional fraction of the cold gas in the satellite is moved into the hot component associated with the central galaxy. If the stripped satellite is a Type 1, and it carries an ICL component, this is removed at the first episode of stripping and added to the ICL component of the central galaxy. It is worth stressing that Eq.~\ref{eq:modc} is valid for disk galaxies that are accreted on a system with velocity dispersion typical of a galaxy group, and that we are extrapolating the validity of this equation to a wider halo mass ranges. \subsection[]{Merger channel for the formation of the ICL} \label{sec:mergers} \citet{giuseppe} argue that the bulk of the ICL is not due to tidal stripping of satellite galaxies (that in their simulations accounts for no more than 5-10 per cent of the total diffuse stellar component), but to relaxation processes taking place during the mergers that characterize the build-up of the central dominant galaxy. In our reference model, if two galaxies merge, the stellar mass of the merging satellite is added to the stellar (bulge) mass of the central galaxy. Based on the findings described above, we add a `merger channel' to the formation of the ICL by simply assuming that, when two galaxies merge, 20 per cent of the satellite stellar mass gets unbound and is added to the ICL of the corresponding central galaxy. We have verified that this simple prescriptions reproduces approximately the results of the numerical simulations by \cite{alvaro}, though in reality the fraction of stars that is unbound should depend on the orbital circularity (Villalobos et al., in preparation). We have also verified that assuming that a larger fraction of the satellite stellar mass gets unbound, obviously leads to higher ICL fractions. In particular, assuming that 50 per cent of the stellar mass of the satellite is unbound, almost doubles the ICL fractions predicted. Assuming an even higher fraction does not affect further the ICL fraction because the effect of having more stellar mass unbound is balanced by the fact that merging galaxies get significantly less massive. A similar prescription was adopted in \citet{pigi} who showed that this has important consequences on the assembly history of the most massive galaxies, and in \citet{somerville} as one possible channel for the formation of the ICL in the context of a hierarchical galaxy formation model. In order to test the influence of this channel on our results, in the following we will present results with this channel both on and off. \subsection[]{Modelling the bulge and disk sizes} Our reference model (DLB07) does not include prescriptions to model the bulge size, that we use in our stellar stripping models. To overcome this limitation, we have updated the reference model including the prescriptions for bulge and disk growth described in \cite{qiguo}. Both the gaseous and stellar components of the disk are assumed to follow an exponential profile. Assuming a flat circular velocity curve, the scale-lengths of these two components can be written as: \begin{equation} R^g_{sl}=\frac{J_{gas}/M_{gas}}{2V_{max}}, \hspace{1cm} R^*_{sl}=\frac{J_{*}/M_{*,disk}}{2V_{max}}, \end{equation} where $J_{gas}$ and $J_*$ are the angular momenta of the gas and stars, $M_{gas}$ and $M_{*,disk}$ are the gas and stellar mass of the disk, and $V_{max}$ is the maximum circular velocity of the dark halo associated with the galaxy. Following \citet{qiguo}, we assume that the change in the angular momentum of the gas disk during a timestep can be expressed as the sum of the angular momentum changes due to addition of gas by cooling, accretion from minor mergers, and gas removal through star formation. The latter causes a change in the angular momentum of the stellar disk. We refer to the original paper by Guo et al. for full details. We have verified that switching back to the simple model for disk sizes of \citet{mo98} used in our reference model, does not affect significantly the results discussed in the following\footnote{Note that the disk size enters in the calculation of the star formation rate. Therefore, a change in the disk size model could, in principle, affect significantly model results.}. Bulges grow through two different channels: mergers (both major and minor) and disk instability. Following \citet{qiguo}, we estimate the change in size due to a merger using energy conservation and the virial theorem: \begin{equation} \frac{G M_{new,b}^2}{R_{new,b}} = \frac{G M_1^2}{R_1} + \frac{G M_2^2}{R_2} + \frac{M_1 M_2}{R_1+R_2} \end{equation} The same approach is adopted in case of disk instability, simply replacing $R_1$ and $M_1$ with the size and mass of the existing bulge, and $R_2$ and $M_2$ with the size and mass of the stars that are transferred from the disk to the bulge so as to keep the disk marginally stable. $R_2$ is determined assuming that the mass is transferred from the inner part of the disk, with the newly formed bulge occupying this region. Again, we refer to the original paper by \citet{qiguo} for full details on this implementation. \section[]{The Galaxy Stellar Mass Function} \label{sec:massfunction} \begin{center} \begin{figure*} \includegraphics[scale=.47]{figures/csmf_groups.eps} \includegraphics[scale=.47]{figures/csmf_clusters.eps} \caption{Left panel: The conditional stellar mass function of satellite galaxies in haloes in the mass range $13.5 < \log M_{200} [M_{\odot} \hm] < 13.8$. Our simulations provide a total of 49 haloes in this mass range. Predictions from our reference model (DLB07) are shown by a solid black line, while predictions from our three models including disruption or/and stripping of satellite galaxies are shown by thinner lines of different style, as indicated by the legend. Symbols with error bars show observational measurements based on SDSS by \citet{liu10}. Right panel: as in the left panel, but for the 3 haloes from our simulations with $14.4 < \log M_{200} [M_{\odot} \hm] < 14.7$.} \label{fig:MF} \end{figure*} \end{center} Before focusing our discussion on the ICL component, it is interesting to analyse how the proposed prescriptions affect one basic prediction of our model, that is the galaxy stellar mass function. In Figure~\ref{fig:MF}, we show the conditional stellar mass function of satellite galaxies in the 49 haloes from our simulations that fall in the mass range $13.5 < \log M_{200} [M_{\odot} \hm] < 13.8$ in the left panel, and in the 3 haloes from our simulations with $14.4 < \log M_{200} [M_{\odot} \hm] < 14.7$ in the right panel. We show predictions from both our reference model (DLB07, shown as a solid black line), and from the three models including the treatments for the stripping and/or disruption of satellite galaxies discussed in Section~\ref{sec:model} (lines of different style, see legend). Model predictions are compared with observational measurements by \citet{liu10}. These are based on group catalogues constructed from the Sloan Digital Sky Survey (SDSS) Data Release 4, using all galaxies with extinction corrected magnitude brighter than $r=18$ and in the redshift range $0.01 \leq z \leq 0.2$. We do not show here predictions from the models with the merger channel for the formation of the ICL switched on, as these do not deviate significantly from the corresponding models with the merger channel off. For the lower halo mass range considered, the reference model fails in reproducing the observed stellar mass function, over-predicting the abundance of galaxies with stellar mass below $\sim 10^{10}\,{\rm M}_{\odot}$. This problem is somewhat alleviated when including a model for stellar stripping, but not solved. In particular, our `disruption model' (model Disr. in the figure and hereafter) does not significantly affect the abundance of the most massive satellites, while reducing the number of their lower mass counterparts (not by the amount required to bring model predictions in agreement with observational results). This is expected as this model only acts on Type 2 galaxies, that dominate the low-mass end of the galaxy mass function. Predictions from the `tidal radius model' (model Tid.) are not significantly different from those of model Disr., while the `continuous stripping model' (model Cont. Strip.) significantly under-predicts the abundance of the most massive satellite galaxies. There are two possible explanations for this behaviour: (i) the abundance of massive satellites is reduced because these are significantly affected by our stripping model or (ii) these massive satellites have disappeared because they have merged with the central galaxies of their parent haloes. As discussed in Section~\ref{sec:modc}, our model Cont. Strip. uses a different prescription for merger times with respect to that employed in the reference model and in models Disr. and Tid. We find that, in this model, merger times are on average shorter than in the other models, which is the reason for the under-prediction of massive satellites shown in Figure~\ref{fig:MF}. As we will discuss in the following, this also implies that the stellar mass of the BCGs in model Cont. Strip. are on average larger than those predicted by models Disr. and Tid. For the higher halo mass range considered, the observed number density of intermediate-to-low mass galaxies is higher, and all our models appear to be in agreement with observational measurements. The agreement remains good also at the massive end, with the exception of model Cont. Strip. that significantly under-predicts the number density of massive galaxies. As explained above, this is due to the shorter galaxy merger times used in this model. We will show below that, in our models, the merger channel does not provide the dominant contribution to the ICL formation so that this is mainly driven by stripping and/or disruption of satellites. Therefore, the excess of intermediate to low-mass galaxies for haloes in the lower mass range might invalidate our predictions. However, as we will show below, the bulk of the ICL originates from relatively massive galaxies so that this particular failure of our models does not significantly affect our results. \section[]{ICL fraction and dependency on halo properties} \label{sec:haloprop} \begin{center} \begin{figure*} \includegraphics[scale=.48]{figures/icl_halomass.eps} \hspace{10pt} \includegraphics[scale=.48]{figures/icl_bcg_halomass.eps} \caption{Left panel: ICL fraction as a function of halo mass. Lines of different style and colour show median results from different models, as indicated in the legend. Right panel: ratio between the ICL component plus BCG stellar mass and the total stellar mass within $R_{500}$. Symbols with error bars show observational estimates by \citealt{gonzalez07}. In both panels, the gray shaded region shows the 20th and 80th percentiles of the distribution for model Disr. The other models have comparable scatter.} \label{fig:ICLfrac} \end{figure*} \end{center} We now turn our analysis to the ICL component, and start by analysing the overall fraction of ICL predicted by our models, and how it depends on halo properties. The left panel of Figure~\ref{fig:ICLfrac} shows the ICL fraction as a function of halo mass. To measure the predicted fractions, we have considered all galaxies within $R_{200}$ and with stellar mass larger than $M_* = 10^{8.3} M_{\odot} $, that approximately corresponds to the resolution limit of our simulations. Lines of different style and colour show the median relations predicted by our different prescriptions, as indicated in the legend. The grey region marks the 20th and 80th percentiles of the distribution found for model Disr. (the other models exhibit a similar dispersion). The predicted fraction of ICL varies between $\sim 20$ per cent for model Disr. with the merger channel off, and $\sim 40$ per cent for model Cont. Strip. with the merger channel on. A relatively large halo-to-halo variation is measured for all models. For none of our models, we find a significant increase of the ICL fraction with increasing halo mass (see discussion in Section \ref{sec:intro}), at least over the range of $M_{200}$ shown. In the figure, we have also considered the ICL associated with Type 1 galaxies within $R_{200}$ of each halo. We stress, however, that their contribution is, on average, smaller than 7 per cent of the total ICL associated with the cluster. As expected, when including a merger channel for the formation of the ICL, its fraction increases, by about $\sim 25$ per cent in models Disr. and Tid., and by about $\sim 50$ per cent in model Cont. Strip. As mentioned in the previous Section, this difference is due to the different dynamical friction formula used in this model, which makes merger times significantly shorter than in the other two models. The amount of ICL that comes from the merger channel is not negligible in the framework of our models and, as expected, increases in the case of shorter merger times. Overall, predictions from our models agree well with fractions of ICL quoted in the literature, i.e. $10-40$ per cent going from groups to clusters (e.g. \citealt{feldmeier04,zibetti08,mcgee10,toledo11}). As discussed in Section~\ref{sec:intro}, it is not an easy task to separate the ICL from the stars that are bound to the BCG. To avoid these difficulties, and possible biases introduced by the adoption of different criteria in the models and in the observations, we consider the ratio $(M_{ICL}+M_{BCG})/M^*_{total}$. The right panel of Figure~\ref{fig:ICLfrac} shows how this fraction varies for the different models considered in this study, and compares our model predictions with observational measurements by \cite{gonzalez07}. To be consistent with these observational measurements, we consider in this case all galaxies within $R_{500}$ and brighter than $m_I = 18$ (i.e. galaxies more massive than $\sim 10^{10} M_{\odot}$). Considering the scatter in both our model predictions and in the observational data, the Figure shows that model Disr. (as well as its variation with the merger channel on) is in relatively good agreement with observational data, though it tends to predict higher ratios for the lowest halo masses considered. Model Tid. predicts a higher fraction of stars in ICL+BCG than model Disr., and the median relation lies close to the upper envelope of the observational data. Finally, model Cont. Strip. over-predicts the fraction of stars in ICL+BCG over the entire mass range considered. The merger channel does not affect the predicted trend as a function of halo mass, but this channel increases on average the fraction of stars in ICL+BCG. This is surprising: stars that are contributed to the ICL through the merger channel would contribute to the BCG mass if the channel is off so that the merger channel should not affect the sum of the ICL and BCG stellar masses. The difference found is due to slight changes in the merger history of the BCGs, due to variations in the merger times of satellite galaxies. In fact, satellite merger times become slightly longer than in a model that does not include stellar stripping because satellite galaxies gain less stellar mass through accretion as central galaxies (also the satellites that were accreted on them were stripped). Because of the shorter merger times mentioned above, model Cont. Strip. is affected more by this change. \begin{center} \begin{figure*} \includegraphics[scale=.503]{figures/icl_c200_clusters.eps} \hspace{10pt} \includegraphics[scale=.50]{figures/icl_tform_clusters_f05.eps} \caption{Left Panel: Fraction of ICL as a function of halo concentration for the 53 haloes from our simulations with $M_{200}[M_{\odot}/h] > 10^{14}$. Halo concentration is defined as $c_{200} = r_s /r_{200}$, where $r_s$ is the characteristic scale obtained by fitting the halo density profile to a NFW profile. Right Panel: Fraction of ICL as a function of halo formation time, defined as the time when the main progenitor of the halo has acquired 50 per cent of its final mass. Predictions from different models are shown by lines of different colour, as indicated in the legend. The gray shaded region shows the 20th and 80th percentile of the distribution found for model Disr.} \label{fig:ICLconc} \end{figure*} \end{center} As discussed above, the ICL fraction predicted by our models does not vary as a function of the host halo mass but exhibits a relatively large halo-to-halo scatter, particularly at the group mass scale. The natural expectation is that this scatter is largely determined by a variety of mass accretion histories at fixed halo mass. We can address this issue explicitly using our simulations. In Figure~\ref{fig:ICLconc}, we show how the ICL fraction correlates with the halo concentration in the left panel, and with the halo formation time in the right panel. We note that these two halo properties are correlated \citep[see e.g.][]{giocoli12}. As usually done in the literature, we have defined the formation time of the halo as the time when its main progenitor has acquired half of its final mass. The concentration has been computed by fitting the density profile of the simulated haloes with a NFW profile \citep*{nfw}. To remove the known correlations between halo mass and concentration/formation time \citep[][]{bullock01,Neto_etal_2007,power12}, we consider in Figure~\ref{fig:ICLconc} only the 53 haloes from our simulations with $M_{200}[M_{\odot}/h] > 10^{14}$. As in previous figures, we show the dispersion (20th and 80th percentiles of the distribution) only for model Disr. The other two models exhibit a similar scatter. The figure shows that, for the halo mass range considered, the ICL fraction increases with increasing concentration/formation time for models Disr. and Tid while remaining approximately constant for model Cont. Strip. Therefore, for models Disr. and Tid., large part of the scatter seen in Figure~\ref{fig:ICLfrac} for haloes in the mass range considered can be explained by a range of dynamical histories of haloes. Haloes that `formed' earlier (those were also more concentrated) had more time to strip stars from their satellite galaxies or accumulate ICL through accretion of smaller systems, and therefore end up with a larger fraction of ICL. For model Cont. Strip., no clear trend as a function of either halo concentration or halo formation time is found. This happens because, contrary to the other two models, the way we model stellar stripping in the Cont. Strip. model does not introduce any dependence on halo concentration. In this model, the amount of stellar mass stripped from satellite galaxies depends only on properties computed at the time of accretion (i.e. mass ratio, circularity of the orbit). In contrast, in models Disr. and Tid., the stripping efficiency is computed considering the instantaneous position of satellite galaxies. Dynamical friction rapidly drags more massive satellites (those contributing more to the ICL) towards the centre, where stripping becomes more efficient. Our results show that equation \ref{eq:modc} is not able to capture this variation, although it is by construction included in the simulations used to calibrate our model. For lower mass haloes, there is no clear correlation between ICL fraction and the two halo properties considered. This is in part due to the fact that, as we will see in the next section, the bulk of the ICL forms very late - later than the typical formation time of low-mass haloes. In this low-mass range, large part of the scatter in the ICL fraction is driven by the fact that these haloes typically contain relatively few massive galaxies, that are those contributing the bulk of the ICL (see next section). So it is the scatter in the accretion of single massive galaxies that drives the relatively large dispersion seen in Figure~\ref{fig:ICLfrac} for haloes with mass smaller than $10^{14}\,{\rm M}_{\odot}$. We have explicitly verified this by considering the 20 per cent haloes in this mass range with the highest and lowest ICL fractions. We find that for those haloes that have larger ICL components, this has been contributed by a few relatively massive satellite galaxies (stellar mass larger than $10^{10}\,{\rm M}_{\odot}$). In contrast, the ICL in haloes with lower ICL fractions was contributed by less massive satellite galaxies. \begin{center} \begin{figure} \includegraphics[scale=.43]{figures/iclfrac_sat_fixmass_rbin.eps} \caption{Fraction of ICL as a function of the number of satellites within $R_{200}$, with mass larger than the thresholds indicated in the legend. The figure refers to haloes in the mass range $10^{13.4} < M_{200} [M_{\odot} \hm] < 10^{13.6}$, and to the model Disr.} \label{fig:ICLnumsat} \end{figure} \end{center} In a recent work, \citet{purcell07} study how the fraction of ICL varies over a wide range of halo masses (from that of spiral galaxies like our Milky Way to that of massive galaxy clusters) using an analytic model for subhalo infall and evolution, and empirical constraints to assign a stellar mass to each accreted subhalo. In their model, the stellar mass associated with a subhalo is assumed to be added to the diffuse component when a certain fraction of the subhalo dark matter mass has been stripped by tidal interaction with the parent halo. In particular, their fiducial model assumes that disruption of the stellar component starts when 20 per cent of the subhalo mass remains bound. Over the range of halo masses sampled in our study, \citet{purcell07} predict a weak increase of the ICL fraction, from $\sim 20$ per cent for haloes of mass $\sim 10^{13}\,{\rm M}_{\odot}$ to about 30 per cent for massive galaxy clusters of mass $\sim 10^{15}\,{\rm M}_{\odot}$. One specific prediction of their model is that the fraction of ICL correlates strongly with the number of surviving satellite galaxies. As they explain, this correlation arises from the fact that haloes that acquired their mass more recently had relatively less time to disrupt the subhaloes they host, and therefore have less ICL. We analyse the same correlation within our models in Figure~\ref{fig:ICLnumsat}. This shows the ICL fraction as a function of the number of satellite galaxies within $R_{200}$ for haloes in the mass range $10^{13.4} < M_{200} [M_{\odot} \hm] < 10^{13.6}$, and for model Disr.. Lines of different style correspond to different mass cuts, as indicated in the legend. The figure shows that, when considering all galaxies with stellar mass larger than $10^9\,{\rm M}_{\odot}$, no significant trend is found between the ICL fraction and the corresponding number of surviving satellite galaxies (this number ranges between $\sim 20$ and $\sim 45$ for this stellar mass cut). When increasing the stellar mass cut, the number of surviving satellites decreases (as expected), and a trend appears in the sense that larger ICL fractions are measured for lower numbers of surviving satellites. Similar trends are found for model Tid., while for model Cont. Strip. no trend is found between the number of surviving satellites and the ICL fraction, for any mass cut used. \section[]{Formation of the ICL} \label{sec:formation} In this section, we take advantage of our models to analyse when the bulk of the ICL forms, which galaxies provide the largest contribution to it, what is the fraction of ICL that has been accreted from other haloes during the hierarchical growth of clusters, and what fraction is instead contributed from the merger channel. \begin{figure} \begin{center} \includegraphics[scale=.45]{figures/strip_disr.eps} \caption{Fraction of the ICL component as a function of the galaxy stellar mass that contributed to it. In the case of model Disr., the galaxy stellar mass on the x-axis refers to the mass that the satellite has before its disruption. For model Cont. Strip., it corresponds to the satellite mass before stripping takes place. For model Tid., both cases can occur.} \label{fig:ICLstarmass} \end{center} \end{figure} In Figure~\ref{fig:ICLstarmass}, we show the contribution to the ICL from galaxies with different stellar mass. For model Disr., the galaxy mass on the x-axis corresponds to that of the satellite right before its disruption, while for model Cont. Strip. it corresponds to the satellite mass before stripping takes place. For model Tid., both cases can occur. The distributions shown in Figure~\ref{fig:ICLstarmass} represent the average obtained considering all haloes in our simulated sample. When considering haloes in different mass bins, the distributions are similar but, as expected, they shift towards lower stellar masses for lower mass haloes. The figure shows that the bulk of the ICL comes from galaxies with stellar masses $\sim 10^{11}\,{\rm M}_{\odot}$ for models Disr. and Tid., and $\sim 10^{10.3}\,{\rm M}_{\odot}$ for model Cont. Strip. In particular, we find that for models Disr. and Tid., about 26 per cent of the ICL is contributed by galaxies with stellar mass in the range $10^{10.75}-10^{11.25} M_{\odot}$. About 68 per cent comes from satellites more massive than $10^{10.5} M_{\odot}$, while dwarf galaxies contribute very little. For model Cont. Strip., almost all the ICL mass ($\sim 90$ per cent of it) comes from satellites with mass in the range $10^{9}-10^{11} M_{\odot}$. The merger channel does not affect significantly the distributions shown. The result discussed above can be easily understood in terms of dynamical friction: the most massive satellites decay through dynamical friction to the inner regions of the halo on shorter time-scales than their lower mass counterparts. Tidal forces are stronger closer to the halo centre, so that the contribution to the ICL from stripping and/or disruption of massive galaxies is more significant than that from low mass satellites. The latter tend to spend larger fractions of their lifetimes at the outskirts of their parent halo, where tidal stripping is weaker. The differences between predictions from models Disr. and Tid. and those from model Cont. Strip. are due to a combination of different effects. On one side, model Cont. Strip. uses a different merger time prescription that leads to significantly shorter merger times than in models Disr. and Tid. This is particularly important for the most massive satellites that have the shortest merger times. In addition, while in models Disr. and Tid. satellite galaxies can be completely destroyed, stripping takes place in a more continuous and smooth fashion in model Cont. Strip. These two effects combine so that the largest contribution to the ICL in this model comes from `intermediate' mass satellites that orbit long enough in the cluster potential to be affected significantly by stellar stripping. Similar results have been found in other studies. In the work by \citet{purcell07} mentioned above, the ICL on the cluster mass scale is largely produced by the disruption of satellite galaxies with mass $\sim 10^{11}\,{\rm M}_{\odot}$. In a more recent work, \citet{martel12} combine N-body simulations with a subgrid treatment of galaxy formation. They find that about 60 per cent of the ICL in haloes more massive than $\sim 10^{14}\,{\rm M}_{\odot}$ is due to the disruption of galaxies with stellar mass in the range $6\times 10^{8} -3\times10^{10}\,{\rm M}_{\odot}$. Their results are close to predictions from our model Cont. Strip., with an enhanced contribution of intermediate-mass galaxies with respect to the other two models discussed in this study and results from \citet{purcell07}. However, in agreement with our results and those from \citet{purcell07}, they also find that the contribution from low-mass galaxies to the ICL is negligible, though they dominate the cluster galaxy population in number. \begin{figure} \begin{center} \includegraphics[scale=.43]{figures/icl_bcg_growth.eps} \caption{Fraction of stellar mass in ICL as a function of the look-back time, normalized to the amount of ICL at present (thick lines of different style and color). Thinner lines show the evolution in the stellar mass of the main progenitor of BCGs as a function of lookback time.} \label{fig:ICLwhen} \end{center} \end{figure} The analysis discussed above answers the question on `which' galaxies contribute (most) to the ICL component. We can now take advantage of results from our models to ask `when' the ICL is produced. We address this issue in Figure~\ref{fig:ICLwhen} that shows the ICL fraction (normalized to the total amount of ICL measured at present) as a function of cosmic time, for all different prescriptions used in this study (thick lines of different colour and style). The cosmic evolution of the ICL component is compared with the evolution of the stellar mass in the main progenitor of the corresponding BCGs, shown as thin lines. In agreement with previous studies both based on simulations \citep{willman,giuseppe} and on analytic or semi-analytic models \citep{conroy,pigi}, we find that the bulk of the ICL forms relatively late, below $z=1$. Models Disr. and Tid. predict very similar ICL growth histories, while in model Cont. Strip. the ICL formation appears to be anticipated with respect to the other two models. At redshift $\sim 1$, less than 10 per cent of the ICL was already formed in models Disr. and Tid. If the merger channel for the formation of the ICL is switched on in these models, the ICL fraction formed at the same redshift increases to $\sim 15-20$ per cent. As explained earlier, the importance of the merger channel is enhanced in model Cont. Strip. The figure also shows that the ICL grows slower than the mass in the main progenitor of the BCG down to $z\sim 1$, i.e. at a lookback time of $\sim 8$~Gyr. Below this redshift, the ICL component grows much faster than the BCG, with more than 80 per cent of the total ICL mass found at z=0 being formed during this redshift interval. \begin{figure} \begin{center} \includegraphics[scale=.60]{figures/icl_accreted.eps} \caption{Top panel: fraction of accreted ICL as a function of the BCG stellar mass. Bottom panel: fraction of ICL contributed by the merger channel as a function of the BCG stellar mass. The grey shaded region shows the 20th and 80th percentiles of the distribution measured for model Disr. The other models exhibit similar dispersions.} \label{fig:ICLaccr} \end{center} \end{figure} We now want to quantify what is the fraction of ICL that is accreted onto the cluster during its hierarchical growth. As explained in Section~\ref{sec:model}, the amount of ICL associated with a central galaxy can increase through three channels: \begin{itemize} \item[(i)] stripping of satellite galaxies orbiting in the same parent halo; \item[(ii)] mergers, if this particular channel is switched on; \item[(iii)] accretion of the ICL component that is associated with new galaxies infalling onto the cluster during its assembly history or with satellite galaxies (i.e. when a Type 1 becomes a Type 2 galaxy in model Disr. or when a Type 1 is stripped for the first time in models Tid. and Cont. Strip). \end{itemize} In the following, we define the `accreted' component as the ICL fraction that is coming through the third channel described above. The contribution from this component is shown in the top panel of Figure~\ref{fig:ICLaccr} as a function of the stellar mass of the BCG. In models Disr. and Tid., the fraction of the accreted component increases from a few per cent for the least massive BCGs in our sample to $\sim 13-25$ per cent for the most massive BCGs, in case the merger channel is off. For model Cont. Strip., the increase as a function of the BCG stellar mass is less pronounced, and the fraction of accreted ICL is always below 10 per cent even in the case the merger channel is on. The bottom panel of Figure~\ref{fig:ICLaccr} quantifies the amount of ICL that comes from the merger channel. If this channel is switched on, as we have seen in the left panel of Figure~\ref{fig:ICLfrac}, the ICL fraction increases in each model. We note that the amount of ICL that comes from this channel cannot be inferred precisely by comparing each model in Figure~\ref{fig:ICLfrac} with its counterpart including the merger channel, because it affects slightly the merger times of galaxies. The contributions to the ICL coming from mergers are shown in the bottom panel of Figure~\ref{fig:ICLaccr}, and have been stored using the three prescriptions used in this study with the merger channel on. We find that in models Disr. and Tid. the merger channel contributes to $\sim 15$ per cent of the total ICL. For model Cont. Strip., the contribution from mergers is significantly larger, ranging from $\sim 30$ per cent for the least massive BCGs in our sample, to $\sim 40$ per cent for the most massive ones. \section[]{ICL and BCG properties} \label{sec:bcgprop} We now focus on the relation between the ICL and the main properties of BCGs, such as stellar mass and luminosity, and analyse how these are affected by the inclusion of our prescriptions for the formation of the ICL. \begin{figure} \begin{center} \includegraphics[scale=.46]{figures/icl_bcg.eps} \caption{Stellar mass in the ICL component as a function of the BCGs stellar mass. Lines of different style and colour correspond to different models, as indicated in the legend. The grey shaded region shows the 20th and 80th percentiles of the distribution obtained for model Disr.} \label{fig:ICLBCGmass} \end{center} \end{figure} In Figure \ref{fig:ICLBCGmass} we show the relation between the mass in the ICL component and the stellar mass of the BCG. As expected, more massive BCGs reside in haloes that host a more conspicuous ICL component. For models Disr. and Tid., the correlation is strong for BCGs more massive than $\sim 3\times10^{11}\,{\rm M}_{\odot}$, while it is very weak for less massive central galaxies. Model Cont. Strip. predicts a weaker correlation over the mass range explored, and a significantly lower mass in the ICL component with respect to the other two models when the merger channel is off. \begin{figure} \begin{center} \includegraphics[scale=.45]{figures/lumbcg_halomass.eps} \caption{Relation between the BCG luminosity in K-band and the cluster mass ($M_{200}$). Black stars with error bars show the observational measurements by \citet{lin}, while black diamonds with error bars are observational measurements by \citet{popesso07}. The thick black solid line shows predictions from our reference model (DLB07), while thinner lines of different style show predictions from the models including our different prescriptions for the formation of the ICL. The grey shaded region shows the 20th and 80th percentiles of the distribution obtained for the reference model. Our other models have comparable scatter.} \label{fig:BCGlumhalomass} \end{center} \end{figure} Figure~\ref{fig:BCGlumhalomass} shows the luminosity of the BCG in the K-band as predicted by our models as a function of the halo mass. Model predictions are compared with observational measurements by \citet{lin} and \citet{popesso07}. It is worth recalling that our model luminosities are `total' luminosities, that are difficult to measure observationally: \citet{popesso07} use SDSS `model magnitudes', while \citet{lin} use elliptical aperture magnitudes corresponding to a surface brightness of $\mu_K = 20 \, \rm{mag/arcsec^2}$, and include an extra correction of 0.2 mag to get their `total magnitudes'. Our reference model is in very good agreement with both sets of observational data. Models Disr. and Tid. predict slightly lower luminosities than our reference model, as a consequence of the reduced accretion of stellar mass from satellites (either because part of these are destroyed - model Disr., or because they are stripped - model Tid.). Model Cont. Strip. predicts luminosities of central galaxies brighter than those measured by Popesso et al., on the cluster mass scale. This is due to the fact that, as mentioned earlier, merger times are shorter in this model, which increases the stellar mass and the luminosity of the BCGs by accretion of (massive) satellites. \begin{figure} \begin{center} \includegraphics[scale=.85]{figures/ratio_bcgmass.eps} \caption{Ratio between the stellar mass of BCGs in the reference model and the corresponding value in model Disr., as a function of the stellar mass in the reference model.} \label{fig:bcgmass} \end{center} \end{figure} Surprisingly, we find that also in models Disr. and Tid. a small fraction of the BCGs are brighter (more massive) than in the reference model. Naively, we do not expect this to be possible as the only effect of our prescriptions should be that of reducing the stellar mass of satellite galaxies by stripping or disruption. In Figure~\ref{fig:bcgmass}, we show the ratio between the BCG stellar mass in the reference model and the corresponding stellar mass in model Disr., as a function of the former quantity. The Figure shows that the majority of the BCGs are more massive in the reference model, but about 23 per cent of the BCGs in our sample are actually {\it more} massive when we switch on our prescription for the disruption of satellite galaxies. If we additionally switch on the merger channel for the formation of the ICL, the fraction of BCGs that are more massive in model Disr. than in the reference model reduces to about 10 per cent. Model Tid. behaves in a similar way (the corresponding fractions are 28 and 9 per cent, respectively). Model Cont. Strip., as also evident from Figure~\ref{fig:BCGlumhalomass} behaves differently. In this model, about half (42 per cent) of the BCGs are more massive than in the reference model, with this fraction reducing to about 26 per cent when the merger channel is switched on. While in model Disr. (and Tid.) the effect seems to be limited to the less massive BCGs (those with stellar mass lower than $\sim 10^{11.5} M_{\odot}$), in model Cont. Strip. this happens for BCGs of any mass. \begin{figure} \begin{center} \includegraphics[scale=.55]{figures/halo1401.eps} \caption{Top panel: Mass in the hot component available at each redshift, for a BCG that is found to be more massive in model Disr. (and less massive in model Cont. Strip.) than in the reference model. Bottom panel: hot gas metallicity at each redshift. Results are shown for different models, as indicated in the legend.} \label{fig:BCGevol} \end{center} \end{figure} In order to understand this finding, we have analysed the evolution of the stellar mass, mass gained through mergers, and cooling rate for a number of the BCGs that are more massive in model Disr. than in the reference model. We have found that in all cases, the reason for the increased mass of the BCGs can be traced back to a more efficient cooling rate. In order to illustrate this, we show in Figure~\ref{fig:BCGevol} a representative example. The top panel shows the amount of hot gas available for cooling onto the main progenitor of the BCG, up to $z\sim 3$, while the bottom panel shows the corresponding metallicity. Results are shown for the reference model, and for our models Disr. and Cont. Strip. (model Tid. behaves similarly to the reference model). For this particular BCG, the evolution of both the hot gas content and that of its metallicity in model Disr. follow very closely the evolution in the reference model. At $z\sim 1.5$, the metallicity of the hot gas in model Disr. becomes higher than in the reference model, and this causes a significant increase in the cooling rate. In turn, the more efficient cooling determines an increase in the star formation, and therefore of the final BCG stellar mass. The behaviour is different in model Cont. Strip. where the metallicity of the hot gas actually falls below the corresponding value in the reference model. As we have explained earlier, for this particular model we have used a different implementation of the merger times which introduces a net decrease in the merger time of satellite galaxies. This more rapid merger rates causes about half of the BCGs to be more massive in model Cont. Strip. than in the reference model. The difference in the hot gas metallicity of the BCG between model Disr. and Cont. Strip. is due to a different metal content of the gas accreted from satellite galaxies. We recall that when a satellite galaxy is destroyed (or stripped), its stellar mass goes to the ICL while its gaseous content, including the corresponding metals, go to the hot gas component associated with the central galaxy. In model Cont. Strip., stripping is a continuous process that starts as soon as a galaxy becomes a satellite. In addition, in this model the formation of the ICL component starts earlier than in the other models, as shown in Figure~\ref{fig:ICLwhen}. In this case, the gas that is removed from satellites tends to dilute the metallicity of the hot gas component. In model Disr. (as well as in model Tid.), the formation of the ICL starts a bit later, and satellite galaxies are more massive than in model Cont. Strip. (and therefore also more metal rich - see Figure~\ref{fig:ICLstarmass}), so that their disruption tends on average to increase the metal content of the hot gas component. \section[]{ICL metallicity} \label{sec:metallicity} From the observational viewpoint, little is known about the stellar populations of the ICL component. Using I-band HST data, \citet{durrell02} compared the brightness of Virgo ICL red giant branch (RGB) stars to that of RGB stars in a metal-poor dwarf galaxies, and estimated an age for the ICL population older than $\sim 2$~Gyr, and a relatively high metallicity ($-0.8\lesssim [{\rm Fe/H}] \lesssim -0.2$). \citet{williams07} use HST observations of a single intra-cluster field in the Virgo Cluster and find that the field is dominated by low-metallicity stars ($[{\rm M/H}]\lesssim -1.$) with ages older than $\sim 10$~Gyr. However, they find that the field contains stars of the full range of metallicities probed ($-2.3 \leq [{\rm M/H}] \leq 0.0$), with the metal-poor stars exhibiting more spatial structure than metal-rich stars, suggesting that the intra-cluster population is not well mixed. Using long-slit spectra and measuring the equivalent width of Lick indices, \citet{Coccato_etal_2011} find that most of the stars in the dynamically hot halo of NGN3311 (the BCG in the Hydra I cluster) are old and metal-poor ($[{\rm Z/H}]\sim -0.35$). In this section, we present predictions of our models concerning in particular the metallicity of the ICL component. We recall that our model adopts an instantaneous recycling approximation for chemical enrichment. In particular, we assume that a constant yield of heavy elements is produced per solar mass of gas converted into stars, and that all metals are instantaneously returned to the cold phase. Metals are then exchanged between the different phases proportionally to the mass flows. When a satellite galaxy is stripped of some fraction of its stars (or destroyed), a proportional fraction (or all) of the metals are also moved from the satellite stars into the ICL. \begin{center} \begin{figure*} \includegraphics[scale=.48]{figures/icl_metallicity.eps} \hspace{10pt} \includegraphics[scale=.48]{figures/metstar_icl.eps} \caption{Left panel: mean metallicity of the ICL component as a function of halo mass. Predictions from different models are shown using lines of different style and colour. The shaded gray region shows the 20th and 80th percentiles of the distribution obtained for model Disr. Right panel: average distribution of the metallicities of the stars in the ICL component for all haloes in our sample.} \label{fig:ICLmet} \end{figure*} \end{center} The left panel of Figure~\ref{fig:ICLmet} shows the median metallicity of the ICL component, as a function of halo mass, for our different prescriptions. The grey shaded region shows the 20th and 80th percentiles of the distribution found for model Disr. (the other models exhibit similar dispersions). In our models, the ICL metallicity does not vary significantly as a function of the halo mass. Assuming $Z_{\odot}=0.02$, the average metallicity of the ICL in models Disr. and Tid. is $ \sim 0.63 Z_{\odot}$, while for model Cont. Strip. the ICL metallicity is significantly lower ($\sim 0.32-0.50 Z_{\odot}$). This is a consequence of the fact that the galaxies contributing to the ICL are on average less massive (and more metal-poor) than those that contribute to the ICL in models Disr. and Tid. (see Figure~\ref{fig:ICLstarmass}). In the right panel of Figure~\ref{fig:ICLmet}, we show the average metallicity distribution of stars in the ICL component for all haloes in our sample. As a consequence of the results shown in Figure~\ref{fig:ICLstarmass}, model Cont. Strip. predicts a distribution shifted towards lower metallicities, with a peak at $\sim 0.4 Z_{\odot}$. Models Disr. and Tid. predict distributions that are less broad and peaked at higher metallicities. Our model results are therefore qualitatively consistent with observational measurements by \citet{williams07}, with most of the stars in the ICL having sub-solar metallicity but covering a relatively wide range. \begin{center} \begin{figure} \includegraphics[scale=.45]{figures/bcg_metallicity.eps} \caption{BCG stellar metallicity as a function of halo mass. Lines of different style and colour correspond to the different prescriptions used in our study, as indicated in the legend. The thick solid black line shows predictions from our reference model.} \label{fig:BCGmet} \end{figure} \end{center} Figure~\ref{fig:BCGmet} shows the BCG metallicity predicted by the different models used in our study, as a function of the halo mass. The figure shows that all models predict very similar metallicities for the BCGs, of about $ 0.74 \, Z_{\odot}$. Predictions are close to those of the reference model (thick solid black line). On average, stellar stripping slightly increases the BCG metallicity, particularly in model Tid. This happens because the most massive galaxies receive less low-metallicity stars from satellite galaxies whose masses (and metal contents) are reduced because of stellar stripping. Our results confirm findings by \citet{gabste} who show that the model mass-metallicity relation is offset low with respect to the observational measurements at the massive end. In particular, the observed BCG metallicities (similar to those of the most massive galaxies) are expected to be at least 0.2-0.3~dex larger \citep{vonderlinden07,loubser09}. Figure~\ref{fig:BCGmet} shows that the inclusion of a model for the formation of the ICL component does not significantly improve this disagreement. More in general, we find that our modelling of the formation of the ICL component does not significantly affect the predicted mass-metallicity relation. This is in apparent contrast with findings by \citet{henriques} who claim that the introduction of satellite disruption is sufficient to bring the stellar metallicities of the most massive galaxies in agreement with the observational data. We note that \citet{henriques} use the same reference model adopted in our study and employ a Monte Carlo Markov Chain parameter estimation technique to constrain the model with the K-band luminosity function, the B-V colours, and the black hole-bulge mass relation. The `best fit' model found by \citet{henriques} includes a model for tidal stripping of the satellite galaxies, but also adopts different parameters with respect to the reference model, in particular for the supernovae feedback and gas recycling process (see their Table 2). We therefore argue that, as discussed also in \citet{gabste}, that stellar stripping cannot provide alone the solution to the problem highlighted above, and that modifications of the star formation and feedback processes are required. \section[]{Discussion and conclusions} \label{sec:discussion} In this work, we build upon the semi-analytic model presented in \citet[][ DLB07]{dlb} to describe the generation of intra-cluster light (ICL). We include different implementations for modelling the formation of the diffuse ICL. In particular, we consider: (i) a model that assumes the stellar component of satellite galaxies can be affected only after their parent dark matter substructures are stripped below the resolution limit of the simulation (Disruption model); (ii) a model that accounts for stellar stripping also from satellites sitting in distinct dark matter subhaloes, and based on a simple estimate of the tidal radius (Tidal Radius model); and (iii) a model based on a fitting formula derived from a suite of numerical simulations aimed to study the evolution of a disk galaxy within the global tidal field of a group environment \citep[][ Continuous Stripping model]{alvaro}. In addition, we have also considered the relaxation processes acting during galaxy-galaxy mergers by simply assuming that 20 per cent of the stellar mass of the merging satellite gets unbound and ends-up in the ICL component associated with the remnant galaxy. In our implementations, the bulk of the ICL is produced through tidal stripping and disruption of the satellite galaxies, with the merger channel contributing only for a minor fraction. The reference model we have used is known to over-predict the abundance of galaxies with mass below $\sim 10^{10}\,{\rm M}_{\odot}$ \citep{Fontanot_etal_2009,qiguo}. The inclusion of a model for stellar stripping of satellite galaxies alleviates this problem, but does not solve it. In a recent work, \citet{budzynski12} use a catalogue of groups and clusters from SDSS DR7 in the redshift range $0.15 \leq z \leq 0.4$ and compare the galaxy number density profiles with predictions from the same reference model used in our study. They show that the model follows very well the observational measurements but in the very central regions (within $\sim 0.2 R_{500}$), where the predicted profile is steeper than observational measurements. The inclusion of stellar stripping would improve the agreement with data in this region where the tidal field is stronger and galaxies are more likely to be stripped. However, the same comparison with the data by \citet{lin2} would lead to an opposite conclusion. This suggests that the uncertainty on the density profiles in the inner region is probably too large to put strong constraints on stripping models. As we discuss below, we find that the dominant contribution to the ICL formation comes from stripping and/or disruption of massive satellites so that the excess of intermediate to low-mass galaxies does not affect significantly our results. In our Cont. Strip. model, we use a different prescription for merger times with respect to that employed in the reference model and in the other two models considered (see Section \ref{sec:modc}). As a consequence, merger times are on average shorter than in the other models which result in an under-prediction of massive satellites. We note that recent work by \citet{Villalobos_etal_2013} has pointed out that the dynamical friction formula used in our Continuous Stripping model (\citet{bk}) under-estimate merger times at higher redshift. Implementing their proposed modification would make merger times longer in this model, making results more similar to the other two models. We have explicitly tested (by adding a fudge factor that increases again the merger times so as to bring the predicted mass function in agreement with observations) that this does not affect significantly the results presented in this work. A number of recent studies have focused on the formation of the ICL, using both hydrodynamical numerical simulations \citep[e.g.][]{giuseppe,puchwein,rudick06}, and analytic models based on subhalo infall and evolution \citep[e.g.][]{purcell07,watson}. Less work on the subject has been carried out using semi-analytic models of galaxy formation, but for basic predictions in terms of how the fraction of ICL depends on the parent halo mass \citep{pigi,somerville,qiguo}. Our models predict an ICL fraction that varies between $\sim 20$ and $\sim 40$ per cent (depending on the particular implementation adopted), with no significant trend as a function of the parent halo mass. Results are in qualitative agreement with observational data, in particular on the cluster mass scale. We note, however, that the ICL fractions predicted by our models depend on the resolution of the simulations: for a set of simulations that use a particle mass one order of magnitude larger than that adopted in the high resolution runs used in our study, the predicted ICL fractions increase by ~30-40 per cent. We stress that both the data and the model predictions exhibit a relatively large halo-to-halo scatter. On the cluster mass scale, we find that the scatter is largely due to a variety of mass accretion histories at fixed halo mass, as argued by \citet{purcell07}: objects that formed earlier (that were also more concentrated) had more time to strip stars from their satellite galaxies or accumulate ICL through accretion of smaller systems. On group scale, where the predicted scatter is very large, we do not find any clear correlation between ICL fraction and halo concentration or formation time. We show that, on these scales, large part of the scatter is driven by individual accretion events of massive satellites. The (albeit weak) correlation between the ICL fraction and concentration on the cluster mass scale can be tested observationally, e.g. by measuring the ICL fraction and concentration for system lying in a relatively narrow halo mass range. Our models predict that the ICL forms very late, below redshift $z\sim 1$, in agreement with previous analysis based on hydrodynamical numerical simulations \citep[e.g.][]{giuseppe}. About 5 to 25 per cent of the diffuse light has been accreted during the hierarchical growth of dark matter haloes, i.e. it is associated with new galaxies falling onto the haloes during their assembly history. In addition, we find that the bulk of the ICL is produced by the most massive satellite galaxies, $M \sim 10^{10-11}\,{\rm M}_{\odot}$, in agreement with recent findings based on N-body simulations \citep{martel12} and analytic models \citep{purcell07}. Low-mass galaxies ($M_* < 10^{9} M_{\odot}$) contribute very little to the ICL in terms of mass, although they dominate in terms of number. This is a natural consequence of dynamical friction triggering the generation of the ICL: the most massive satellites approach the inner cluster regions faster than their less massive counterparts. Close to the cluster centre, tidal forces are stronger, increasing the stripping efficiency. In contrast, small satellites spend most of their time in the outer regions where tidal stripping is weaker. Since most of the ICL is produced by tidal stripping of massive satellites, this component is found to have a metallicity that is similar to that of these galaxies. Our model predictions are in qualitative agreement with observations, with most of the stars in the diffuse component having on average sub-solar metallicities (but covering a relatively large range). We also find that the mean metallicity of the ICL is approximately constant as a function of halo mass, and exhibit a relatively small halo-to-halo scatter. In contrast, \citet{purcell08} predict a weak increase of the ICL metallicity with increasing halo mass, over the same halo mass range considered in our study. Finally, we show that the inclusion of a model for tidal stripping of satellite galaxies does not significantly affect the predicted mass-metallicity relation, and only slightly increases the metallicity of the most massive galaxies. For all models, these galaxies have stellar metallicities significantly lower than observed (see also \citealt{gabste}). Future and more detailed observations focused e.g. on age and metallicity of the ICL component will help constraining our models and understanding the physical mechanisms driving the formation of this important stellar component. \section*{Acknowledgements} EC, GDL and AV acknowledge financial support from the European Research Council under the European Community's Seventh Framework Programme (FP7/2007-2013)/ERC grant agreement n. 202781. This work has been supported by the PRIN-INAF 2009 Grant ``Towards an Italian Network for Computational Cosmology'', the PRIN-MIUR 2009 grant ''Tracing the Growth of Structures in the Universe'' and the PD51 INFN grant. Simulations have been carried out at the CINECA National Supercomputing Centre, with CPU time allocated through an ISCRA project and an agreement between CINECA and University of Trieste. We acknowledge partial support by the European Commissions FP7 Marie Curie Initial Training Network CosmoComp (PITN-GA-2009-238356). We thank the referee, Stefano Zibetti, Magda Arnaboldi and Douglas Watson for useful comments that helped us improving our manuscript. \bsp \label{lastpage} \bibliographystyle{mn2e}
\section{Introduction} \label{intro} Coupling light and matter is an essential part in many quantum information protocols \cite{RevModPhysGisin2002}. For many of these protocols to be implemented successfully this coupling should be as high as possible. The scheme used most frequently for achieving high coupling efficiencies is to place the matter system into a high-quality resonator. Here we will rather focus on light-matter coupling in free space and the measurement and characterization of the coupling efficiency. When investigating this efficiency typically three distinct effects are measured: Firstly, efficient coupling increases the probability of a photon being absorbed by a matter system. This provides an opportunity to use matter as a quantum memory in which one can store the state of a photon. Here, a measure for the coupling efficiency could be the probability with which one can store a single photon in a matter system \cite{NatPhysPiro2011}. As for all other types of experiments described below, spatially mode matching the light field to the transition is essential. In addition, depending on the inner structure of the matter system involved, it is necessary to create an optimal temporal shape of the incident field \cite{OptCommQuabis2000,PhyScrLeuchs2012,PhysRevLettAljunid2013}. Thus the absorption probability is affected by two different effects that have to be distinguished by additional measurements. A second effect occurring in light-matter interaction is the phase shift that a light field acquires when interacting dispersively with a medium. Here, the phase a field accumulates in comparison to a non-interacting field provides a good measure for the strength of the interaction \cite{JEurOptSocRapPublicSondermann2013,PhysRevLettPotoschnig2011,PRLAljunid2009,PRAHetet2013}, given the light is scattered coherently. This, however, is only the case if no upper-level population is induced and hence there is no incoherent scattering. This is only the case if the driving field is zero. To account for the amount of incoherently scattered light the upper-level population has to be determined. Additionally, investigating the phase of a field requires some way of stabilizing the phase of the non-interacting field. A further effect that is often investigated in the context of high coupling efficiencies is the extinction of a light field traveling past a matter system. This effect is related to the previous one, because both effects originate from the interference between the impinging light and the light scattered by the emitter \cite{PRLZumofen2008,PRLAljunid2009,PhysRevLettPotoschnig2011}. In contrast to the phase shift, which is maximized by illumination from full solid angle \cite{JEurOptSocRapPublicSondermann2013,JourModOptLeuchs2012}, optimal results are obtained when focusing light from half the solid angle \cite{PRLZumofen2008}. The depth of the dip in the transmission through the system provides a good measure for the amount of light that was interacting \cite{PhysRevAKochan1994,NatPhysWrigge2008,NJPTey2009,PRLSlodicka2010}. Like in the case of the phase shift the effect is reduced by incoherently scattered light \cite{NatPhysWrigge2008}. Here, we will establish saturation measurements as a tool for characterizing the coupling efficiency in free space in an unambiguous way, utilizing the very effect that is detrimental in the types of measurements discussed above. The next section discusses the advantages of saturation measurements in more detail and reviews the relation between coupling efficiency and the necessary power to reach a given upper-level population. The experimental set-up is described in Sec. \ref{setup}, whereas the experimental results are presented in Sec. \ref{results} and discussed in Sec. \ref{discussion}. \section{Saturation Measurements} \label{saturation} In what follows we present an approach that provides a measure for the spatial overlap of the light field with the driven transition while neglecting temporal effects. A two-level-system \mbox{(TLS)} responds to the power of the driving field in a non-linear way. The amount of light scattered by a TLS is directly proportional to its upper-level population $\rho$. Solving the Bloch equations one finds that for strong driving fields the upper-level population in the steady state solution asymptotically reaches $\rho=1/2$ where the TLS scatters at a rate of $\Gamma/2$, where $\Gamma$ is the spontaneous emission rate of the TLS. Thus, one can directly relate the upper-level population to the amount of scattered photons. For example, at an upper-level population of $\rho=1/4$ the TLS scatters at a rate of $\Gamma/4$. This value is commonly associated with a saturation parameter $S=1$ in the literature by $\rho=\frac{S}{2(S+1)}$. Following Ref. \cite{PRAvanEnk2004} the rate at which the ion scatters photons equals \begin{eqnarray} \textrm{R}_\textrm{sc} = \frac{\Gamma}{2}\left(1-\frac{1+\delta^2}{1+\delta^2+8\left|\beta\right|^2/\Gamma}\right) \label{eq:scat} \end{eqnarray} with $|\beta|^2$ the rate of dipolar photons, $\delta=2\Delta/\Gamma$ with $\Delta$ the detuning from resonance. However, a real beam is in general comprised of dipolar and non-dipolar parts. Hence the rate of interacting photons is lowered by the ratio of dipolar parts. This ratio can be expressed by the coupling efficiency which characterizes the focusing geometry including the overlap of the incident radiation pattern with the dipole transition of the TLS. In Refs. \cite{JourModOptLeuchs2012,EurPhysJGolla2012} the coupling efficiency has been defined as $G=\Omega\eta^2$, where $\Omega$ denotes the solid angle fraction of the focusing geometry weighted with the dipolar emission pattern. The overlap with the dipole is accounted for by $\eta$ which may also account for distortions of the phase front of the incident beam. In addition, there might be some other potentially unknown sources of imperfection influencing the coupling efficiency like e.g. residual motion \cite{OptCommTeo2011}. We will include these by a loss factor $1-L$ resulting in $G=\Omega\eta^2\left(1-L\right)$. Thus, a scattering rate of $\textrm{R}_\textrm{sc}=\Gamma/4$ and an upper-level population of $\rho=1/4$ is reached at an impinging photon rate of \begin{eqnarray} |\tilde{\beta}|^2=\frac{\Gamma}{8}\left(1+\delta^2\right)/G \label{eq:ratein} \end{eqnarray} In the optimal case all incident photons interact with the TLS and $G=1$. From this the minimal power to reach an upper-level population of $\rho=1/4$ can be calculated to be $P_{\rho=1/4}=\hbar \omega_0 \Gamma /8$ on resonance \cite{PhysRevAKochan1994}, with the atomic transition frequency $\omega_0$. By varying the incident power and evaluating \mbox{Eq. (\ref{eq:scat})} one can deduce the necessary power $P^{\textrm{exp}}_{\rho=1/4}$ at which this upper-level population is reached. Comparing the minimal power and incident power measured in the experiment results in the coupling efficiency \begin{eqnarray} G=P_{\rho=1/4}/P^{\textrm{exp}}_{\rho=1/4} \label{eq:G_spa} \end{eqnarray} This way of measuring the coupling efficiency has several advantages: First of all it is independent of the losses in the detection system since they can be factored out by normalizing the rate of detected photons when strongly driven to its asymptotic value at full saturation. Furthermore, compared to measurements that are restricted to the weak excitation regime which is often noise-dominated, this is a fast method if the overall detection efficiency is sufficiently high. Finally it is not subject to detrimental effects based on the saturation of the upper-level population since these are the very effects that are utilized in this method. \section{Experimental Set-up} \label{setup} Fig. \ref{fig:setup} shows a schematic of the experimental set-up. The basic arrangement has been described earlier in Ref. \cite{PhysRevAMaiwald2012}. At the heart of the experiment lies an ion trap similar to that used in Ref. \cite{NatPhysMaiwald2009} and a parabolic mirror of focal length $\textit{f} = 2.1~\textrm{mm}$. For the saturation measurements performed here this mirror serves two purposes: It collects the light scattered by the ion and thus is the first element of the detection system. Additionally it serves as a mode converter that transforms an incoming first order Laguerre-Gaussian mode with radial polarization into a linear dipole wave converging at the focus\cite{LaserPhysLindlein2007,ApplPhysBSondermann2007}. This light mode is generated by sending a linearly polarized Gaussian beam onto a segmented half wave plate \mbox{(SHWP)}. The resulting beam is spatially filtered by a 30-$\mu \textrm{m}$ pinhole such that only the two lowest order modes are transmitted \cite{EurPhysJGolla2012}. The size of the Laguerre-Gaussian beam is adjusted such that it yields the maximal overlap with a linear dipole field. Afterwards part of the beam is transmitted through a non-polarizing beamsplitter to the mirror. In order to control the position of the ion relative to the focus of the parabolic mirror the trap is mounted on a linear \textit{x-y-z}-piezo stage. The correct position is found by imaging the light scattered by the ion onto an electron-multiplying charge-coupled device \mbox{(EMCCD)}. Afterwards the ion is fine positioned by maximizing the amount of light scattered by the ion when driven by the Laguerre-Gaussian mode. In order to distinguish the light driving the ion from the scattered light it is necessary to filter the impinging laser light out of the detection path. In our set-up this is done by an aperture \mbox{(HSA)} with a diameter of $D=2\times\textit{f}$. Hence we illuminate the ion only from the inner part of the mirror which corresponds to half of the solid angle. After a first reflection on the parabolic mirror the inner half of the beam is focused onto the ion and then collimated again after a second reflection in the outer half of the mirror. In this way the excitation beam is blocked by the same aperture used to cut it in the first place. Due to the high collection efficiency provided by the mirror the ion can be simultaneously monitored by the \mbox{EMCCD} and by an avalanche photodiode \mbox{(APD)}. \begin{figure} [Htb] \centering \resizebox{0.95\columnwidth}{!}{ \includegraphics{setup.pdf} } \caption{Schematic of the set-up. The Laguerre-Gaussian beam is created by sending a Gaussian beam onto a segmented half wave plate \mbox{(SHWP)} and subsequent filtering by a pinhole \mbox{(PH)}. It is then transmitted to the parabolic mirror \mbox{(PM)} through a 50/50-beamsplitter \mbox{(BS)}. In order to be able to distinguish the light scattered by the ion from the driving laser field the beam is cut by an aperture \mbox{(HSA)} cutting the beam to half of the solid angle. Solid lines indicate the extreme most parts of the beam going towards the ion. The dashed lines show the path of the light after interaction with the ion. The light scattered by the ion is simultaneously monitored by an EMCCD and an APD.} \label{fig:setup} \end{figure} To minimize losses originating from residual motion the correlation between the ion's fluorescence and the radio frequency applied to the trap is monitored \cite{JApplPhysBerkeland1998}. Since the Laguerre-Gaussian beam excites the ion from almost all directions it is possible, by using different apertures, to measure the correlation signal from three linearly independent directions. The excess micromotion is compensated by minimizing the modulation of the correlation signal for these directions. In order to measure the couping efficiency $G$ the saturation of the $^{2}\textrm{S}_{1/2}\leftrightarrow$$^{2}\textrm{P}_{1/2}$ cooling transition at a wavelength of $369.5~\textrm{nm}$ of $^{174}\textrm{Yb}^+$ with a natural linewidth of $\Gamma/2\pi=19.6~\textrm{MHz}$ is investigated. Since the ion decays to the $^2D_{3/2}$ level with a probability of $0.5~\%$ \cite{PRAOlmschenk2007} we repump it with light at a wavelength of $935~\textrm{nm}$ to the $^3\textrm{D}[3/2]_{1/2}$ level from where it decays back into the cooling cycle \cite{PRABell1991}. The power of the Laguerre-Gaussian beam is varied by an acousto-optic modulator \mbox{(AOM)}. The scattered light is detected for $100~\textrm{ms}$. During this time the repumping laser is applied to avoid optical pumping to the meta-stable $^{2}\textrm{D}_{3/2}$ state. The $^{2}\textrm{D}_{3/2}\leftrightarrow$$^{3}\textrm{D}[3/2]_{1/2}$ transition is strongly driven to ensure that the ion can be treated as a close approximation of a pure TLS. Afterwards the repumping beam is switched off and the $369.5~\textrm{nm}$ transition is strongly driven for $10~\mu\textrm{s}$ to pump the ion into the $^{2}\textrm{D}_{3/2}$ state. Then the background light is measured for $100~\textrm{ms}$ and subtracted. In order to provide cooling during the experimental sequence the Laguerre-Gaussian beam is detuned by half a linewidth from resonance. \section{Results} \label{results} \begin{figure} [Htb] \centering \resizebox{0.95\columnwidth}{!}{ \includegraphics{saturationMeas.pdf} } \caption{Saturation curve of a single $^{174}\textrm{Yb}^+$ ion illuminated by a Laguerre-Gaussian beam from half solid angle. The data points are background corrected and were taken at half linewidth detuning from resonance. The dashed line shows a fit to the data, yielding a necessary power of $1081~\textrm{pW}$ incident onto the parabolic mirror to reach an upper-level population of $\rho=1/4$. Each data point was measured for $100~\textrm{ms}$.} \label{fig:saturation} \end{figure} Figure \ref{fig:saturation} shows the results of a typical measurement under the condition of optimal alignment. Fitting Eq. (\ref{eq:scat}) to the data yields a required power of $P^{\textrm{exp}}_{\rho=1/4}=1081~\textrm{pW}$ impinging on the mirror to reach an upper-level population of $\rho=1/4$. Factoring out the reflectivity of the mirror of $64~\%$ for a radially polarized Laguerre-Gaussian beam leads to a power of $692\pm20~\textrm{pW}$ impinging on the ion. For the $^{2}\textrm{S}_{1/2}\leftrightarrow$$^{2}\textrm{P}_{1/2}$ transition of $^{174}\textrm{Yb}^+$ Eq. (\ref{eq:ratein}) yields a minimal necessary power of $P_{\rho=1/4}=16.6~\textrm{pW}$ at a detuning of $\Delta=\Gamma/2$ to reach an upper-level population of $\rho=1/4$. However since in our experiment we are only aiming at driving the $\pi$-transition and not the $\sigma^\pm$-transitions this power has to be increased by a factor of 3 to $49.7~\textrm{pW}$ according to the Clebsch-Gordan coefficients of a $\textrm{J}=1/2$ to $\textrm{J}=1/2$ transition. Thus our system has a coupling efficiency of $G=7.2\pm0.2\%$. One has to keep in mind, that we are illuminating the ion only from one half of the solid angle, thus the maximally achievable coupling strength is $50~\%$ in this configuration. However, one has to account for known deficiencies. Interferometric measurements on the parabolic mirror \cite{ApplOptLeuchs2008} predict a Strehl ratio of $87~\%$. In combination with the measured overlap of the field incident on the parabolic mirror \cite{EurPhysJGolla2012} this yields an overlap of $\eta=0.91$. Accounting for the hole in the vertex of the parabolic mirror as well as the constraint to half solid angle leads to $\Omega=0.49$. This suggests an expected coupling efficiency $G=40.5~\%$. In order to investigate the origin of the discrepancy from the measured value, a scan of the focal intensity distribution is performed. The focus of the Laguerre-Gaussian beam inside the parabolic mirror is characterized utilizing the fact that the trap is mounted on a \textit{x-y-z}-piezo stage. Thus it is possible to move the ion through the focus and measure the necessary power to reach an upper-level population of $\rho=1/4$ at every point. The inverse of this power provides a quantity proportional to the local intensity of the electric field. Since moving the ion to different locations relative to the mirror's focus changes the detection efficiency of the set-up, the saturation measurements are a reliable tool as their outcome is independent of this efficiency. Figure \ref{fig:focus} shows a scan through the focus perpendicular to the optical axis of the parabolic mirror. The scan is taken at a resolution of $50~\textrm{nm}/\textrm{pixel}$ and takes approximately seven minutes. This resolution is chosen to match the extent of the wave function of the ion assuming Doppler-limited cooling and considering the measured trap frequencies of $560~\textrm{kHz}$ in the radial directions. Analyzing the data yields a full width at half maximum \mbox{(FWHM)} of $530~\textrm{nm}$ in X-direction and $610~\textrm{nm}$ in Y-direction. Scanning the focus along the optical axis yields a FWHM of $660~\textrm{nm}$ in Z-direction. From interferometric measurements performed on the parabolic mirror we calculate the expected shape of the intensity distribution in the focus when illuminated by a Laguerre-Gaussian beam from half solid angle. These simulations predict a FWHM of $140~\textrm{nm}$ perpendicular to the optical axis and $415~\textrm{nm}$ along the optical axis. Additional measurements of the phasefront and polarization of the Laguerre-Gaussian beam were performed. These measurements do not indicate that the focus should broaden significantly. Since the measured focal distribution is clearly larger than that expected from the simulations there seems to be a blurring effect. This blurring might originate in turning grooves that stem from the manufacturing process of the mirror. \begin{figure} [Htb] \centering \resizebox{1.0\columnwidth}{!}{ \includegraphics{focusScan.pdf} } \caption{Intensity distribution of the focus of the Laguerre-Gaussian beam after reflection off the parabolic mirror. The scan is taken perpendicular to the optical axis. The resolution is $50~\textrm{nm}/\textrm{pixel}$. Each pixel value is obtained by a single saturation measurement. The plots on the left and bottom side show a cut through the maximum of the focus. The solid black lines show a simulation of the intensity distribution accounting for the known aberrations of the parabolic mirror. The FWHM of the measured intensity distribution amounts to $530~\textrm{nm}$ in X-direction and $610~\textrm{nm}$ in Y-direction.} \label{fig:focus} \end{figure} \section{Discussion} \label{discussion} By spatially tailoring the incident light mode to the emission pattern of the linear dipole transition we are able to reach coupling strengths of approximately 7~\% while exciting the ion from half solid angle. Scanning the ion through the focus of the parabolic mirror shows that the extent of the focus is greater than what is to be expected from simulations by a factor of 4.1 in the transversal directions and by a factor of 1.6 in longitudinal direction. Since these factors are on the same order of magnitude as the deviation of the measured and the expected coupling efficiency, we conjecture that the blurring of the focus is responsible for this mismatch. Nevertheless, this coupling efficiency is, to our knowledge, among the best reported for single atoms in free space. In other experiments, the coupling efficiency as defined here is seldom specified. An exception is Ref. \cite{PhysRevLettAljunid2013} in which a coupling efficiency of $3~\%$ is reported. For most of the other works on light-matter coupling in free space one has to estimate the coupling efficiency from the data provided. For Ref. \cite{NatPhysWrigge2008} based on an extinction of $22~\%$ we infer a coupling efficiency with a magnitude comparable to the one reported here. One should, however, recall that our results were obtained without correcting for the aberrations of the parabolic mirror and a non-optimal surface quality. This suggests that the use of a mirror with better surface quality as well as aberration correction will boost the coupling efficiency to values on the order of $90~\%$ when illuminating the ion from full solid angle, as it has been envisioned in Ref. \cite{EurPhysJGolla2012} based on the quality of the incident optical mode. This would provide a powerful tool for many quantum information protocols as well as promising steps towards new quantum mechanical technologies such as e.g. a quantum transistor \cite{NatureHwang2009}. \section{Acknowledgments} \label{acknowledgments} The authors thank S. Heugel for fruitful discussions. M. S. acknowledges financial support from the Deutsche Forschungsgemeinschaft (DFG). G. L. wants to thank the German Federal Ministry of Education and Research (BMBF) for financial support in the framework of the joint research project \textit{QuORep}.
\section{} \begin{acknowledgments} This work was supported in part by the Center for Excitonics, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Award Number DE-SC0001088. Device fabrication was carried out at the Center for Functional Nanomaterials, Brookhaven National Laboratory, which is supported by the U.S. Department of Energy, Office of Basic Energy Sciences, under Contract No. DE-AC02-98CH10886, and at WPAFB, Ohio. Graphene assembly was supported by the Center for Re-Defining Photovoltaic Efficiency Through Molecule Scale Control, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Science under Award Number DE-SC0001085. The authors thank Nicholas Harris for the helpful discussions. \end{acknowledgments}
\section*{Introduction} With the advancement of computer technologies, lattice QCD calculations have become an essential tool to tackle problems of strong interaction physics. For many applications nowadays the statistical uncertainties have been reduced to such a degree that a full understanding of all systematic errors is more important to reach high precision than a further increase of statistics. One of these systematic errors comes from the lattice itself: The symmetry group on the lattice is the hypercubic group $H(4)$, a subgroup of $O(4)$. This introduces lattice artifacts for any lattice momentum but zero which only disappear in the continuum limit. Of course, the severity of these hypercubic lattice artifacts depends on the lattice spacing $a$, but since lattice calculations are currently done for values of $a$ between $0.04$ and $0.09\,\text{fm}$ it may be important to know these artifacts for one's favorite observable on a quantitative rather than qualitative level. In general, discretization effects are most visible at large lattice momentum \(a^2p^2\) with $p_\mu = 2\pi k_\mu/ aN_\mu$ \big(with $k_\mu\in(-N_\mu/2,N_\mu/2]$\big). However, this dependence is non-monotonous as lattice observables transform as functions of the four \(H(4)\) invariants \(\{p^2, p^{[4]}, p^{[6]}, p^{[8]}\}\) with \(p^{[n]} \equiv \sum_\mu p_\mu^n\) rather than just of $p^2$ as in the continuum. There are mainly two approaches that have been followed in the past to treat these artifacts. One is the so-called $H(4)$-method (see, e.g., \cite{Becirevic:1999uc,*Becirevic:1999hj,*Boucaud2003,*Soto2007}), which relies on fitting coefficients of an expansion in the hypercubic invariants. Another is calculating hypercubic corrections in lattice perturbation theory (LPT). This has been used and developed for precision determinations of renormalization constants for hadronic operators \cite{Gockeler:2010yr,*Constantinou:2013ada,*Constantinou:2013ova} where an efficient treatment of such artifacts is crucial. For these calculations one often uses plane-wave sources for the inversion of the fermion matrix to keep the statistical noise at a minimum and also restricts to momenta close to the lattice diagonal to minimize hypercubic lattice artifacts. To further reduce discretization errors, one subtracts the difference $\Delta O(p)$ of an operator $O$ in LPT at finite lattice spacing~$a$ and its value in the corresponding limit $a\to0$. Unfortunately, the calculation of \(\Delta O(p)\) even at 1-loop order can be rather involved in LPT for certain operators and actions. Furthermore, in some cases a 1-loop correction does not suffice. Numerical stochastic lattice perturbation theory (NSPT) provides a way out: It allows for a non-diagrammatic but stochastic treatment of LPT to high loop orders and so to get an estimate for $\Delta O(p)$ beyond 1-loop.\footnote{For a NSPT calculation of renormalization constants see for example \cite{Brambilla:2013sua}.} We will test the feasibility and precision of such an approach here for the gluon~($Z$) and ghost~($J$) dressing functions and in particular for the Minimal MOM coupling in Landau gauge~\cite{vonSmekal:1997is,*vonSmekal:2009ae} \begin{equation} \alpha_s^{\mathsf{MM}}(p) = \frac{g_0^2(a)}{4\,\pi} Z(a, p) J^2(a, p)\text{,}\qquad a \rightarrow 0\text{,} \label{eq:minimom} \end{equation} which serves as an ideal benchmark for this: Its renorma\-lization-group invariance is slightly broken on the lattice at larger momenta by the hypercubic lattice artifacts of the gluon and ghost dressing functions and a subtraction of the exact 1-loop correction was shown to be insufficient to repair this in the data for $\alpha_s^{\mathsf{MM}}$~\cite{Sternbeck2012}. It also allows us to focus first on quenched QCD, keeping the additions of fermions to the NSPT calculations for later. \section*{$Z$ and $J$ from Numerical Stochastic Perturbation Theory} \begin{figure*} \centering \mbox{\includegraphics[height=6cm]{gluondressing.pdf} \includegraphics[height=6cm]{ghostdressing.pdf}} \caption{NSPT data for the gluon (left) and ghost (right) dressing functions for diagonal lattice momenta up to 3-loop order. Curves represent the $a\to0$ result (see text). Arrows mark the hypercubic lattice artifacts.} \label{fig:hypcor} \end{figure*} To get the desired hypercubic corrections, we have to calculate the Landau-gauge gluon and ghost propagators in NSPT. Details for this can be found in \cite{DiRenzo:2009ni,*DiRenzo2010cs}. Here it suffices to mention that their tensor structure (for the Wilson gauge action) is of the familiar form \begin{equation} D_{\mu\nu}^{ab} = \left(\delta_{\mu\nu}-\frac{\hat{p}_\mu\,\hat{p}_\nu}{\hat{p}^2}\right)\delta^{ ab} \frac{Z}{\hat{p}^2}\qquad\text{and}\qquad G_{\mu\nu}^{ab} = \delta_{\mu\nu}\delta^{ab}\frac{J}{\hat{p}^2} \label{eq:dressingfn} \end{equation} where \(\hat{p}_\mu = \frac{2}{a}\sin(a\,p_\mu/2)\) denotes the eigenvalues of the free lattice Laplacian. The results are obtained in orders of \(g_0^2\), because the \(SU(3)\) link fields \(U_{x,\mu}\) are kept as an expansion in powers of \(\beta=2N_c/g_0^2\) which also requires algebraic operations to be performed with respect to this perturbative structure for operators $A$ and $B$: \begin{equation} U_{x,\mu} = 1 + \sum_{k \ge 1} \beta^{-k/2} U_{x,\mu}^{(k)}\text{,}\qquad (A+B)^{(k)} = A^{(k)}+B^{(k)}\text{,}\qquad (A\,B)^{(k)} = \sum_{j=0}^k A^{(k)}\,B^{(k-j)}\text{.} \end{equation} Furthermore, the framework of NSPT relies on stochastic quantization. This is implemented by adding an artificial time $\tau$ to the gauge fields whose evolution in $\tau$ is governed by a Langevin equation, which is solved numerically. A suitable Euler scheme reads \begin{equation} U_{x,\mu}(\tau+\epsilon) = \exp\{-i\,\epsilon\,\nabla_{U_{x,\mu}}S[U] + \sqrt{\epsilon}\,\eta_x(\tau)\}\,U_{x,\mu}(\tau)\text{,} \end{equation} where \(\eta = \eta^a t^a\) is a random gaussian distributed \(su(3)\) field (\(t^a\) being the generators of \(SU(3)\)) and \(S[U]\) (in our case) the Wilson plaquette action. \(\nabla_{U_{x,\mu}}\) is the discretized left Lie derivative with respect to \(U_{x,\mu}\). Since the Euler time step $\epsilon$ has to be finite, simulations have to be performed for several \(\epsilon\) such that at the end the limit \(\epsilon\rightarrow 0\) can be taken. We choose $\epsilon=0.03$, 0.02 and 0.01 for all our simulations and collect data for $J^{(l)}$ and $Z^{(l)}$ for $l=1,\ldots,4$ loops after fixing $U_{x,\mu}$ to Landau gauge. We will focus here on diagonal lattice momenta and our lattice sizes are $8^4$, $16^4$, $24^4$ and $32^4$. Data for a $48^4$ lattice is in progress. In Figure~\ref{fig:hypcor} we show our NSPT data for the gluon and ghost dressing functions up to 3-loop order. As we are interested in the hypercubic lattice corrections for diagonal lattice momenta we only show points for these momenta and compare it to the momentum dependence \begin{equation} J^{(i)}_{cont}(a^2\,p^2) = j^{(i)}_{0} + \sum_{k=1}^{i} j^{(i)}_{k}\left(\log(a^2\,p^2)\right)^k\text{.} \end{equation} expected in the limit $a\to0$ (for simplicity we write \(J\) meaning both \(J\) and \(Z\)). For the 1-loop curve we use the well-known coefficients, $j^{(1)}_0$ and $j^{(1)}_{1}$, from \cite{Kawai:1980ja}, while for the 2-loop and 3-loop curves we use the coefficients $j^{(i=2,3)}_{k>0}$ multiplying the divergent logarithms from 3-loop continuum perturbation theory \cite{Gracey2003}, and the finite \(j^{(i)}_{0}\) from our fits (see below). Within errors our values for the latter agree with those found in \cite{DiRenzo:2009ni,*DiRenzo2010cs}. For a better illustration, we have marked the momentum-dependent hypercubic lattice corrections $\Delta Z$ and $\Delta J$ in Figure~\ref{fig:hypcor} which we aim to quantify: \begin{equation} \Delta J^{(i)} = J^{(i)}_{lat}(a^2p^{2}, a^4p^{[4]}, a^6p^{[6]}, a^8p^{[8]}) - J^{(i)}_{cont}(a^2\,p^2) \text{.} \label{eq:hypcor} \end{equation}\vspace{-4ex} \begin{figure*} \mbox{\includegraphics[height=6cm]{gluonfit.pdf} \includegraphics[height=6cm]{ghostfit.pdf}}\vspace*{-1ex} \caption{Hypercubic lattice artifacts at 1-loop order for the gluon (left) and ghost (right) dressing functions. $pL$-corrected (uncorrected) data is shown by full (open) symbols. Green curves are the exact 1-loop LPT results from~\cite{Sternbeck2012}. Red bands are from a fit to the data.} \label{fig:lptnspt} \medskip \mbox{\includegraphics[height=6cm]{gluoncor.pdf} \includegraphics[height=6cm]{ghostcor.pdf}}\vspace*{-1ex} \caption{Hypercubic lattice artifacts up to 3-loop order for the gluon (left) and ghost (right) dressing functions at \(\beta=6.92\). Bands result from a fit to our NSPT data.} \label{fig:glughocor} \end{figure*} \section*{Removal of finite size effects} Although a first look at Figure~\ref{fig:hypcor} suggests we will get $\Delta Z$ and $\Delta J$ straightaway, a closer inspection of the raw data reveals we have to carefully treat finite size effects first. These are visible in particular for the ghost propagator at small momenta (see, e.g., the open symbols in Figure~\ref{fig:lptnspt}). To treat these effects we follow a procedure proposed in \cite{DiRenzo:2009ni,*DiRenzo2010cs}, albeit in a slightly different manner. It starts with the observation that an observable \(O\) on the lattice only depends on dimensionless quantities, namely on $ap$ (in the way as mentioned above) and on $pL$ which comes in due to the finite lattice extend $N=L/a$: \vspace{-1ex} \begin{equation} O(p\,a, p\,L) = O(p\,a, \infty) + \delta O(p\,a,p\,L)\text{.} \end{equation} To get $O$ in the infinite volume limit one can either extrapolate data for several volumes or calculate and subtract $\delta O(p\,a,p\,L)$ from the data. We choose the latter, which is more robust in our case, and determine \(\delta O(p\,a,p\,L)\) in the following way, similarly to \cite{DiRenzo:2009ni,*DiRenzo2010cs}: We assume we can neglect the influence of additional (hypercubic) \(p\,a\)-corrections to $\delta O$, i.e., \(\delta O(p\,a, p\,L) \approx \delta O(0, p\,L)\). This quantity then only depends on the integer tuple \(k\) that defines the momentum $ap$, because \( p_\mu\,L = \frac{2\,\pi\,k_\mu}{N\,a} N\,a = 2\,\pi\,k_\mu\text{.}\) Furthermore, an observable \(O(pa,\infty)\) is expected to be equal for equal physical momenta, i.e., for equal \(k/N\). That is, the difference must be due to the finite-size effect and we can write: \begin{enumerate} \item \(\delta O(k, N_1) = \delta O(k, N_2)\) \item \(\delta O(k_1, N_1) = O(k_1, N_1) - \left[O(k_2, N_2) - \delta O(k_2, N_2)\right]\) \qquad \text{if}\quad \(k_1/N_1 = k_2/N_2\)\,. \end{enumerate} This allows us to go through our data along sequences of links (paths), each connecting two data points of either constant \(k/N\) or \(k\), and to get the \(p\,L\)-error for each path recursively. As a starting point for this recursion we choose \(\delta O(k_{max}, N_{max})=0\) which should be valid for large \(N_{max}\). By that procedure we can reach nearly all pairs of \((k, N)\). Most of them can be accessed by many different paths over which we average. The \(pL\)-error of points that are disconnected from the anchor \((k_{max}, N_{max})\) is interpolated linearly between neighbouring values of \(k/N\). This method of finite-size corrections turns out to be very successful. To demonstrate that we show in Figure \ref{fig:lptnspt} (right) our 1-loop data for the uncorrected (open symbols) and the $pL$-corrected (full symbols) difference $\Delta J$ (Eq.~\eqref{eq:hypcor}). There we see, finite size errors have quite an effect at low momenta and it is necessary to remove them, but after the $pL$-correction the $\Delta J$ data from different lattices falls on top of the curve one knows exactly from 1-loop LPT (from \cite{Sternbeck2012}). \begin{figure*} \vspace*{-0.5cm} \centering \includegraphics[height=5.7cm]{coupling.pdf} \caption{Coupling constant in the Minimal MOM scheme for quenched QCD in Landau gauge. Data is for a fixed physical volume (from \cite{Sternbeck2012}), before (left) and after (right) correcting for discretization effects.} \label{fig:coupling} \end{figure*} The gluon propagator data is currently too noisy for a removal of $pL$-effects and it thus remains uncorrected in what follows. Though, within errors, points lie on the 1-loop LPT curve already. \section*{Results} Now, that we have checked that our 1-loop NSPT results conform with the exact from LPT, we can quantify the hypercubic lattice artifacts (Eq.~\eqref{eq:hypcor}) up to 3-loop order. To ease their later use we will provide them in a parametrization independent of the lattice coupling \(\beta\). For this we fit each order separately to a polynomial of the \(H(4)\) invariants: \begin{equation} \Delta J^{(i)}(p^{2}, p^{[4]}, p^{[6]}, p^{[8]})=c_2\,p^2 + c_4\,p^{[4]} + c_6\,p^{[6]} + c_8\,p^{[8]}\text{.} \end{equation} The (summed-up) hypercubic lattice artifacts are then obtained for each value of $\beta$ from: \begin{equation} \Delta J(\beta,p^{2}, p^{[4]}, p^{[6]}, p^{[8]}) = \sum_{i=1}^{3} \beta^{-i} \Delta J^{(i)}(p^{2}, p^{[4]}, p^{[6]}, p^{[8]})\text{.} \end{equation} A comparison with the exact 1-loop LPT result from \cite{Sternbeck2012} shows that these fits describe the overall momentum dependence quite well, even though it cannot describe it perfectly (see Figure~\ref{fig:lptnspt}). There the fit is dominated by the many points at higher momentum while being constrained to zero at \(a^2p^2=0\). So there are small deviations for small $a^2p^2$. We therefore regard this fit model not as best but as best usable to parametrize the leading hypercubic artifacts. The final results for $\Delta Z$ and $\Delta J$ can be seen in Figure~\ref{fig:glughocor}. There we choose $\beta\equiv 6.92$ and one sees that even for this fine lattice (\(a\approx 0.026\,\text{fm}\)) the 3-loop order correction still contributes rather significantly to the gluon dressing function. The ghost dressing function needs improvement to at least 2-loop order. This explains why a 1-loop correction of the data for $\alpha_s^{\mathsf{MM}}$ is insufficient as seen in \cite{Sternbeck2012}. Our new results allows us now to correct the hypercubic lattice artifacts up to 3-loop order, and, as evidenced in Figure~\ref{fig:coupling}, with these we are successful: The corrected $\alpha_s^{\mathsf{MM}}$ data (from \cite{Sternbeck2012}) for the different lattice spacings falls on top of each other for \emph{all} momenta. Renormalization-group invariance is thus restored. \section*{Comparison with the H(4)-Method} \begin{floatingfigure}[r] \centering \vspace*{-0.2cm} \parbox{6.7cm}{% \includegraphics[width=7.cm]{h4vsnspt_diagmom.pdf}} \caption{Hypercubic corrections for the gluon dressing function for diagonal momenta ($64^4$ lattice at \(\beta=6.92\)). Blue crosses are from the \(H(4)\)-method, using all available orbits (not shown). The curve is the 3-loop correction from NSPT. \label{fig:h4method}\vspace*{-1cm}} \end{floatingfigure} It is interesting to compare our approach to the above-mentioned \(H(4)\)-method. With this, one tries to fit the coefficients \(c_i\) of the hypercubic expansion of $O(ap)$, e.g., truncated at \(\mathcal{O}(a^4)\), \begin{equation} O(a\,\hat{p}) = O(a^2\,p^2) + c_2\,a^2\frac{p^{[4]}}{p^2} + c_4\,a^4\,p^{[4]} + \cdots\text{.} \label{eq:h4fit} \end{equation} regarding \(O(a^2p^2)\) as the discretization-artifacts-free operator~\cite{Becirevic:1999uc,*Becirevic:1999hj,*Boucaud2003,*Soto2007}. A simultaneous fit of all the $c_i$'s in \eqref{eq:h4fit} requires many degenerate \(H(4)\) or\-bits for nearby $a^2p^2$, that is this method works best for intermediate momenta on large lattices. For a fair comparison, we therefore choose a $64^4$ lattice, for which we have data for $Z$ for many different orbits. Looking again at diagonal momenta, we find the $H(4)$ correction scatter, where available, around our 3-loop NSPT correction (see Figure~\ref{fig:h4method}). \section*{Conclusion} We have developed a new way to determine and remove discretization errors which are present in any lattice observable due to the hypercubic lattice symmetry. Our method is based on a perturbative calculation in the framework of NSPT, and a first application to quenched data for the Minimal MOM coupling in Landau gauge has been very successful (see Figure~\ref{fig:coupling}). Our method can easily be extended, for instance, to aid calculations of renormalization constants for hadronic operators. It is planned to test this next. More details on our approach will be given in a forthcoming article.\par \acknowledgments We thank Francesco Di Renzo for insights on the determination of finite size effects. This work is supported by the European Union under the Grant Agreement IRG 256594. Grants of time on the Linux-Cluster of the Leibniz Rechenzentrum in Munich (Germany) are acknowledged. {\small
\section{Introduction} \label{sec:introduction} The focus of this article is the search for evidence of continuous gravitational waves, as might be radiated by nearby, rapidly spinning neutron stars, in data from the Laser Interferometer Gravitational-wave Observatory (LIGO) \cite{Abbott:2007kv}. The data used in this paper were produced during LIGO's fifth science run (S5) that started on November 4, 2005 and ended on October 1, 2007. Spinning neutron stars are promising sources of gravitational wave signals in the LIGO frequency band. These objects may generate continuous gravitational waves through a variety of mechanisms including non-axisymmetric distortions of the neutron star, unstable oscillation modes in the fluid part of the star and free precession \cite{Bildsten:1998ey, Ushomirsky:2000ax, Cutler:2002nw, Melatos:2005ez, Owen:2005fn}. Independently of the specific mechanism, the emitted signal is a quasi-periodic wave whose frequency changes slowly during the observation time due to energy loss through gravitational wave emission, and possibly other mechanisms. At an Earth-based detector the signal exhibits amplitude and phase modulations due to the motion of the Earth with respect to the source. A number of searches have been carried out previously in LIGO data \cite{Abbott:2005pu, Abbott:2006vg, Abbott:2007ce, :2007tda, Abbott:2008fx, Abbott:2008uq, :2008rg, Collaboration:2009nc, Collaboration:2009rfa, Abadie:2011md, Abadie:2011wj, Aasi:2012fw} including: targeted searches in which precise pulsar ephemerides from radio, X-ray or $\gamma$-ray observations can be used in a coherent integration over the full observation span; directed searches in which the direction of the source is known precisely, but for which little or no frequency information is known; and all-sky searches in which there is no information about location or frequency. All-sky searches for unknown neutron stars must cope with a very large parameter space volume. Optimal methods based on coherent integration over the full observation time are completely unfeasible since the template bank spacing decreases dramatically with observation time, and even for a coherent time baseline of just few days, a wide-frequency-band all-sky search is computationally extremely challenging. Therefore hierarchical approaches have been proposed~\cite{Brady:1998nj, Papa:2000wg, Krishnan:2004sv, Cutler:2005pn, Prix:2012yu, Shaltev:2013kqa} which incorporate semi-coherent methods into the analysis. These techniques are less sensitive for the same observation time but are computationally inexpensive. The Hough transform~\cite{ Krishnan:2004sv, Abbott:2005pu, Sintes:2006uc, :2007tda, Palomba:2005fp} is an example of such a method and has been used in previous wide-parameter-space searches published by the LIGO and Virgo Collaborations. Moreover it has also been used in the hierarchical approach for Einstein$@$Home searches, as the incoherent method to combine the information from coherently analyzed segments \cite{Collaboration:2009nc, Aasi:2012fw}. In this paper we report the results of an all-sky search making use of the `weighted Hough' method~\cite{Sintes:2006uc, :2007tda, Palomba:2005fp}. The `weighted Hough' was developed to improve the sensitivity of the `standard Hough' search~\cite{Abbott:2005pu, Krishnan:2004sv} and allows us to analyze data from multiple detectors, taking into account the different sensitivities. The work presented here achieves improved sensitivity compared to previous Hough searches~\cite{Abbott:2005pu, :2007tda} by splitting the run into two year--long portions and requiring consistency between signal levels in the two separate years for each candidate event, in addition to incorporating a $\chi^2$--test~\cite{SanchodelaJordana:2008dc}. This new pipeline is efficient at rejecting background, allowing us to lower the event threshold and achieve improved sensitivity. The parameter space searched in our analysis covers the frequency range $50<f<1000\,\mathrm{Hz}$ and the frequency time--derivative range $-8.9\times10^{-10}<\dot{f}<0$ Hz/s. We detect no signals, so our results are presented as strain amplitudes $h_0$ excluded at 95$\%$ confidence, marginalized over the above $\dot{f}$ interval. Through the use of significant distributed computing resources \cite{Einstweb}, another search \cite{Aasi:2012fw} has achieved better sensitivity on the same data as the search described here. But the Einstein$@$Home production run on the second year of S5 LIGO data required about 9.5 months, used a total of approximately 25000 CPU (central processing unit) years \cite{Aasi:2012fw}, and required five weeks for the post-processing on a cluster with 6720 CPU cores. The search presented in this paper used only 500 CPU months to process each of the two years of data, representing a computational cost more than two orders of magnitude smaller. This is also an order of magnitude smaller than the computational cost of the semi-coherent `PowerFlux' search reported in a previous paper \cite{Abadie:2011wj}. The significance of our analysis is through offering an independent analysis to cross-check these results, and a method that allows the attainment of sensitivity close to that of the Einstein$@$Home search at substantially reduced computational burden. This technique will be particularly important in the advanced LIGO and Virgo detector when applied to ``quick-look'' searches for nearby sources that may have detectable electromagnetic counterparts. Moreover, the Hough transform is more robust than other computationally efficient semi-coherent methods with respect to noise spectral disturbances \cite{:2007tda} and phase modeling of the signal. In particular, it is also more robust than the Einstein$@$Home search to the non inclusion of second order frequency derivatives. An important feature to note is that the sensitivity of the Hough search is proportional to $1/( N^{1/4} \sqrt{T_{\textrm{\mbox{\tiny{coh}}}}})$ or $N^{1/4}/ \sqrt{T_{\textrm{\mbox{\tiny{obs}}}}}$, assuming $T_{\textrm{\mbox{\tiny{obs}}}} = N T_{\textrm{\mbox{\tiny{coh}}}}$, being $N$ the number of data segments coherently integrated over a time baseline $T_{\textrm{\mbox{\tiny{coh}}}}$ and combined using the Hough transform over the whole observation time $T_{\textrm{\mbox{\tiny{obs}}}}$, while for a coherent search over the whole observation time, the sensitivity is proportional to $1/ \sqrt{T_{\textrm{\mbox{\tiny{obs}}}}}$. This illustrates the lost of sensitivity introduced combining the different data segments incoherently but, of course, this is compensated by the lesser computational requirements of the semi-coherent method. For sufficiently short segments ($T_{\textrm{\mbox{\tiny{coh}}}}$ of the order of 30 minutes or less), the signal remains within a single Fourier frequency bin in each segment. In this case a simple Fourier transform can be applied as a coherent integration method. As the segment duration $T_{\textrm{\mbox{\tiny{coh}}}}$ is increased, it becomes necessary to account for signal modulations within each segment by computing the so-called ${\cal F}$-statistic \cite{Jaranowski:1998qm} over a grid in the space of phase evolution parameters, whose spacing decreases dramatically with time baseline $T_{\textrm{\mbox{\tiny{coh}}}}$. This results in a significant increase in the computational requirements of the search and also limits the significant thresholds for data points selection and the ultimate sensitivity of the search. The search presented here is based on 30 minute long coherent integration times, being this the reason for the significant reduction of the computational time compared to the Einstein$@$Home search \cite{Aasi:2012fw} in which the span of each segment was set equal to 25 hours. For an in-depth discussion on how to estimate and optimize the sensitivity of wide area searches for spinning neutron stars at a given computational cost, we refer the reader to \cite{Cutler:2005pn, Prix:2012yu, Wette:2011eu}. This paper is organized as follows: Section \ref{sec:data} briefly describes the LIGO interferometers and the data from LIGO's fifth science run. Section \ref{sec:waveform} defines the waveforms we seek and the associated assumptions we have made. In section \ref{sec:houghtransform} we briefly review the Hough-transform method. Section \ref{sec:chi2veto} describes the $\chi^2$ test implemented for the analysis of the full S5 data. Section \ref{sec:pipeline} gives a detailed description of the search pipeline and results. Upper limit computations are provided in section \ref{sec:upperlimits}. The study of some features related to the $\chi^2$-veto is presented in section \ref{sec:veto2}. Section \ref{sec:sensitivities} discusses variations, further improvements and capabilities of alternative searches. Section \ref{sec:end} concludes with a summary of the results. \section{Data from the LIGO's fifth science run} \label{sec:data} During LIGO's fifth science run the LIGO detector network consisted of a 4-km interferometer in Livingston, Louisiana (called L1) and two interferometers in Hanford, Washington, one a 4-km and another 2-km (H1 and H2, respectively). The fifth science run spanned a nearly two-year period of data acquisition. This run started at 16:00~UTC on November 4, 2005 at Hanford and at 16:00~UTC on November 14, 2005 at Livingston Observatory; the run ended at 00:00~UTC on October 1, 2007. During this run, all three LIGO detectors had displacement spectral amplitudes very near their design goals of $1.1\times 10^{ -19}$~m$\cdot$Hz$^{-1/2}$ in their most sensitive frequency band near $150$~Hz for the 4-km detectors and, in terms of gravitational-wave strain, the H2 interferometer was roughly a factor of two less sensitive than the other two over most of the relevant band. \begin{table} \caption{\label{T.times}The reference GPS initial and final time for the data collected during the LIGO's fifth science run, together with the number of hours of data used for the analysis. } \begin{indented} \item[]\begin{tabular}{@{}lllllll} \br & 1st year && & 2nd year && \\ Detector&start &end & hours &start & end& hours \\ \mr H1& 815410991& 846338742 & 5710 & 846375384 & 877610329 & 6295 \\ H2& 815201292 & 846340583 & 6097.5 & 846376386 & 877630716 & 6089\\ L1 & 816070323 & 846334700 & 4349& 846387978& 877760976 & 5316.5\\ \br \end{tabular} \end{indented} \end{table} The data were acquired and digitized at a rate of $16384$~Hz. Data acquisition was periodically interrupted by disturbances such as seismic transients (natural or anthropogenic), reducing the net running time of the interferometers. In addition, there were 1-2 week commissioning breaks to repair equipment and address newly identified noise sources. The resulting duty factors for the interferometers, defined as the fraction of the total run time when the interferometer was locked (i.e., all the interferometer control servos operating in their linear regime) and in its low configuration, were approximately $69\%$ for H1, $77\%$ for H2, and $57\%$ for L1 during the first eight months. A nearby construction project degraded the L1 duty factor significantly during this early period of the S5 run. By the end of the S5 run, the cumulative duty factors had improved to $78\%$ for H1, $79\%$ for H2, and $66\%$ for L1. In the paper the data from each of the three LIGO detectors is used to search for continuous gravitational wave signals. In table \ref{T.times} we provide the reference GPS initial and final times for the data collected for each detector, together with the number of hours of data used for the analysis, where each data segment used was required to contain at least 30 minutes of continuous interferometer operation. \section{The waveform model} \label{sec:waveform} Spinning neutron stars may generate continuous gravitational waves (GW) through a variety of mechanisms. Independently of the specific mechanism, the emitted signal is a quasi-periodic wave whose frequency changes slowly during the observation time due to energy loss through gravitational wave emission, and possibly other mechanisms. The form of the received signal at the detector is \begin{equation} h(t)=F_+\left( t,\psi \right) h_+ \left( t \right) + F_\times \left( t,\psi \right) h_\times \left( t \right) \end{equation} where $t$ is time in the detector frame, $\psi$ is the polarization angle of the wave and $F_{+,\times}$ characterize the detector responses for the two orthogonal polarizations \cite{Bonazzola:1995rb, Jaranowski:1998qm}. For an isolated quadrupolar gravitational-wave emitter, characterized by a rotating triaxial-ellipsoid mass distribution, the individual components $h_{+,\times}$ have the form \begin{equation} h_+ = h_0 \frac{1+\cos^2\iota}{2} \cos\Phi(t) \quad \textrm{and} \quad h_\times = h_0 \,\cos\iota\, \sin\Phi(t), \end{equation} where $\iota$ describes the inclination of the source's rotation axis to the line of sight, $h_0$ is the wave amplitude and $\Phi(t)$ is the phase evolution of the signal. % For such a star, the gravitational wave frequency, $f$, is twice the rotation frequency and the amplitude $h_0$ is given by \begin{equation} \label{eq:GWampl} h_0 = \frac{4\pi^2G}{c^4}\frac{I_{zz}f^2\varepsilon}{d}, \end{equation} where $d$ is the distance to the star, $I_{zz}$ is the principal moment of inertia with respect to its spin axis, $\varepsilon$ the equatorial ellipticity of the star, $G$ is Newton's constant and $c$ is the speed of light. Note that the search method used in this paper is sensitive to periodic signals from any type of isolated gravitational-wave source, though we present upper limits in terms of $h_0$. Because we use the Hough method, only the instantaneous signal frequency in the detector frame, $2\pi f(t) = d \Phi(t)/dt $, needs to be calculated. This is given, to a very good approximation, by the non-relativistic Doppler expression: \begin{equation} \label{eq:Doppler} f(t)-\hat{f}(t) = \hat{f}(t) \frac{\textbf{v}(t)\cdot \textbf{n}}{c}, \end{equation} where $\hat{f}(t)$ is the instantaneous signal frequency in the Solar System Barycenter (SSB), $\textbf{v}(t)$ is the detector velocity with respect to the SSB frame and $\textbf{n}$ is the unit-vector corresponding to the sky location of the source. In this analysis, we search for $\hat{f}(t)$ signals well described by a nominal frequency $f_0$ at the start time of the S5 run $t_0$ and a constant first time derivative $\dot f$, such that \begin{equation} \hat{f}(t) = f_0+ \dot f (t-t_0). \end{equation} These equations ignore corrections to the time interval $t-t_0$ at the detector compared with that at the SSB and relativistic corrections. These corrections are negligible for the search described here. \section{The Hough transform} \label{sec:houghtransform} The Hough transform is a well known method for pattern recognition that has been applied to the search for continuous gravitational waves. In this case the Hough transform is used to find hypothetical signals whose time-frequency evolution fits the pattern produced by the Doppler modulation of the detected frequency, due to the Earth's rotational and orbital motion with respect to the Solar System Barycenter, and the time derivative of the frequency intrinsic to the source. Further details can be found in \cite{Krishnan:2004sv, Sintes:2006uc, Palomba:2005fp, Abbott:2005pu}; here we only give a brief summary. The starting point for the Hough transform are $N$ short Fourier transforms (SFTs). Each of these SFTs is digitized by setting a threshold $\rho_\textrm{\mbox{\tiny{th}}}$ on the normalized power \begin{equation} \label{eq:normpower} \rho_k = \frac{2|\tilde{x}_k|^2}{T_{\textrm{\mbox{\tiny{coh}}}} S_n(f_k)} \,. \end{equation} Here ${\tilde{x}_k}$ is the discrete Fourier transform of the data, the frequency index $k$ corresponds to a physical frequency of $f_k= k/T_{\textrm{\mbox{\tiny{coh}}}}$, $S_n(f_k)$ is the single sided power spectral density of the detector noise and $T_{\textrm{\mbox{\tiny{coh}}}}$ is the time baseline of the SFT. The $k^{th}$ frequency bin is selected if $\rho_k \geq \rho_\textrm{\mbox{\tiny{th}}}$, and rejected otherwise. In this way, each SFT is replaced by a collection of zeros and ones called a peak-gram. This is the simplest method of selecting frequency bins, for which the optimal choice of the threshold $\rho_\textrm{\mbox{\tiny{th}}}$ is 1.6~\cite{Krishnan:2004sv}. Alternative conditions could be imposed \cite{doi:10.1142/S0218271800000438, Astone:2005fj, PhysRevD.66.102003}, that might be more robust against spectral disturbances. For our choice, the probability that a frequency bin is selected is $q = e^{-\rho_\textrm{\mbox{\tiny{th}}}}$ for Gaussian noise and $\eta$, given by \begin{equation} \label{eq:eta} \eta = q\left\{1+\frac{\rho_\textrm{\mbox{\tiny{th}}}}{2}\lambda_k + \mathcal{O}(\lambda_k^2) \right\} \end{equation} is the corresponding probability in the presence of a signal. $\lambda_k$ is the signal to noise ratio (SNR) within a single SFT, and for the case when there is no mismatch between the signal and the template: \begin{equation} \label{eq:lambda} \lambda_k = \frac{4|\tilde{h}(f_k)|^2}{T_{\textrm{\mbox{\tiny{coh}}}} S_n(f_k)} \end{equation} with $\tilde{h}(f)$ being the Fourier transform of the signal $h(t)$. Several flavors of the Hough transform have been developed \cite{Krishnan:2004sv, Sintes:2006uc, Antonucci:2008jp} and used for different searches \cite{Abbott:2005pu, :2007tda, Collaboration:2009nc}. The Hough transform is used to map points from the time-frequency plane of our data (understood as a sequence of peak-grams) into the space of the source parameters. Each point in parameter space corresponds to a pattern in the time-frequency plane, and the Hough number count $n$ is the weighted sum of the ones and zeros, $n_k^{(i)} $, of the different peak-grams along this curve. For the `weighted Hough' this sum is computed as \begin{equation} \label{eq:2} n = \sum_{i=0}^{N-1} w^{(i)}_{k} n^{(i)}_{k}\,. \end{equation} where the the choice of weights is optimal, in the sense of \cite{Sintes:2006uc}, if defined as \begin{equation} \label{eq:3} w^{(i)}_{k} \propto \frac{1}{S^{(i)}_{k}}\left\{ \left(F_{+1/2}^{(i)}\right)^2 + \left(F_{\times 1/2}^{(i)}\right)^2\right\}, \end{equation} where $F_{+1/2}^{(i)}$ and $F_{\times 1/2}^{(i)}$ are the values of the beam pattern functions at the mid point of the $i^{th}$ SFT and are normalized according to \begin{equation} \label{eq:4} \sum_{i=0}^{N-1} w^{(i)}_{k} = N\,. \end{equation} The natural detection statistic is the \textit{significance} (or critical ratio) defined as: \begin{equation} s= \frac{n-\langle n\rangle}{\sigma} \, , \end{equation} where $\langle n\rangle$ and $\sigma$ are the expected mean and standard deviation for pure noise. Furthermore, the relation between the significance and the false alarm probability $\alpha$, in the Gaussian approximation \cite{Krishnan:2004sv}, is given by \begin{equation} \label{eq:sth} s_\textrm{\mbox{\tiny{th}}}= \sqrt{2}\textrm{erfc}^{-1}(2\alpha)\,. \end{equation} \section{The $\chi^2$ veto} \label{sec:chi2veto} $\chi^2$ time-frequency discriminators are commonly used for gravitational wave detection. Originally, they were designed for broadband signals with a known waveform in a data stream \cite{Allen:2004gu}. But they can be adapted for narrowband continuous signals, as those expected from rapidly rotating neutron stars. The essence of these tests is to ``break up" the data (in time or frequency domain) and to see if the response in each chunk is consistent with what would be expected from the purported signal. In this paper, a chi-square test is implemented as a veto, in order to reduce the number of candidates in the analysis of the full S5 data. The idea for this $\chi^2$ discriminator is to split the data into $p$ non-overlapping chunks, each of them containing a certain number of SFTs $\{N_1, N_2,\ldots,N_p\}$, such that \begin{equation} \sum_{j=1}^p N_j =N \, , \end{equation} and analyze them separately, obtaining the Hough number-count $n_j$ which, for the same pattern across the different chunks, would then satisfy \begin{equation} \sum_{j=1}^p n_j =n \, , \end{equation} where $n$ is the total number-count for a given point in parameter space. The $\chi^2$ statistic will look along the different chunks to see if the number count accumulates in a way that is consistent with the properties of the signal and the detector noise. Small values of $\chi^2$ are consistent with the hypothesis that the observed significance arose from a detector output which was a linear combination of Gaussian noise and the continuous wave signal. Large values of $\chi^2$ indicate either the signal did not match the template or that the detector noise was non-Gaussian. In the following subsections we derive a $\chi^2$ discriminator for the different implementations of the Hough transform and show how the veto curve was derived for LIGO S5 data. \subsection{The standard Hough} \label{sec:chi2SH} In the simplest case in which all weights are set to unity, the expected value and variance of the number count are \begin{equation} \langle n\rangle =N\eta \, , \qquad \sigma^2_n =N\eta(1-\eta) \, , \end{equation} \begin{equation} \langle n_j\rangle= N_j\eta= N_j\frac{\langle n\rangle}{N} \, , \qquad \sigma^2_{n_j} =N_j\eta(1-\eta) \, . \end{equation} Consider the $p$ quantities defined by \begin{equation} \Delta n_j\equiv n_j-\frac{N_j}{N}n \, . \end{equation} With this definition, it holds true that \begin{equation}\label{eq:chi=0} \langle \Delta n_j\rangle=0 \, , \qquad \sum_{j=1}^p \Delta n_j=0 \, , \qquad \langle n_j n\rangle= \frac{N_j}{N} \langle n^2\rangle \,, \end{equation} and the expectation value of the square of $\Delta n_j$ is \begin{equation} \langle (\Delta n_j)^2\rangle =\left( 1-\frac{N_j}{N}\right) N_j\eta(1-\eta) \, . \end{equation} Therefore we can define the $\chi^2$ discriminator statistic by \begin{equation} \label{eq:chi3} \chi^2(n_1, \ldots, n_p) = \sum_{j=1}^p {\frac{(\Delta n_j)^2}{\sigma^2_{n_j}} }= \sum_{j=1}^p {\frac{\left(n_j-nN_j/N\right)^2}{N_j\eta(1-\eta)} }\, . \end{equation} This corresponds to a $\chi^2$-distribution with $p-1$ degrees of freedom. To implement this discriminator, we need to measure, for each point in parameter space, the total number-count $n$, the partial number-counts $n_j$ and assume a constant value of $\eta=n/N$. \subsection{The weighted Hough} \label{sec:chi2WH} In the case of the weighted Hough the result given by equation (\ref{eq:chi3}) can be generalized. Let $I_j$ be the set of SFT indices for each different $p$ chunks, thus the mean and variance of the number-count become \begin{equation} \langle n_j \rangle= \sum_{i\in I_j} w_i\eta_i \qquad \langle n \rangle= \sum_{j=1}^p \langle n_j \rangle \qquad \sigma^2_{n_j} = \sum_{i\in I_j} w_i^2\eta_i(1-\eta_i) \end{equation} and we can define \begin{equation} \Delta n_j\equiv n_j-n\frac{\sum_{i\in I_j} w_i\eta_i}{\sum_{i=1}^N w_i\eta_i} \, , \end{equation} so that $\langle \Delta n_j\rangle=0$, $\sum_{j=1}^p \Delta n_j=0$. Hence, the $\chi^2$ discriminator would now be: \begin{equation} \chi^2 = \sum_{j=1}^p {\frac{(\Delta n_j)^2}{\sigma^2_{n_j}} } = \sum_{j=1}^p {\frac{\left(n_j-n (\sum_{i\in I_j} w_i\eta_i)/(\sum_{i=1}^N w_i\eta_i)\right)^2} {\sum_{i\in I_j} w_i^2\eta_i(1-\eta_i)} }\, . \label{eq:chi5} \end{equation} In a given search, we can compute the $\sum_{i\in I_j} w_i$, $\sum_{i\in I_j} w_i^2$ for each of the $p$ chunks, but the different $\eta_i$ values can not be measured from the data itself because they depend on the exact SNR for each single SFT as defined in equations (\ref{eq:eta}) and (\ref{eq:lambda}). For this reason, the discriminator we proposed is constructed by replacing $\eta_i\rightarrow\eta^*$, where $\eta^*=n/N$. In this way, from equation (\ref{eq:chi5}) we get \begin{equation} \label{eq:chi6} \chi^2 \approx \sum_{j=1}^p {\frac{\left(n_j-n (\sum_{i\in I_j} w_i)/N\right)^2} {\eta^*(1-\eta^*) \sum_{i\in I_j} w_i^2 } }\, . \end{equation} In principle, one is free to choose the different $p$ chunks of data as one prefers, but it is reasonable to split the data into segments in such a way that they would contribute a similar relative contribution to the total number count. Therefore we split the SFT data in such a way that the sum of the weights in each block satisfies \begin{equation} \sum_{i\in I_j} \omega_{i} \approx \frac{N}{p} . \end{equation} Further details and applications of this $\chi^2$ on LIGO S4 data can be found in ~\cite{SanchodelaJordana:2008dc}. \subsection{The S5 $\chi^2$ veto curve} We study the behavior of this $\chi^2$ discriminator in order to characterize the $\chi^2$-significance plane in the presence of signals and derive empirically the veto curve. For this purpose we use the full LIGO S5 SFT data, split into both years, in the same way that is done in the analysis, and inject a large number of Monte Carlo simulated continuous gravitational wave signals into the data, varying the amplitude, frequency, frequency derivative, sky location, as well as the nuisance parameters $\cos \iota$, $\psi$ and $\phi_0$ of the signals. Those injections are analyzed with the multi-interferometer Hough code using the same grid resolution in parameter space as is used in the search. \begin{figure}[h] \begin{center} \includegraphics[width=15pc]{SigMeanChi_v2.pdf} \includegraphics[width=15pc]{SigStdChi_v2.pdf} \\ \includegraphics[width=32pc]{veto_v2_2.pdf} \end{center} \caption{\label{fig:veto} (Top left) Mean value of the significance versus mean of the $\chi^2$ and the fitted power law curve for 177834 simulated injected signals. (Top right) Mean value of the significance versus mean $\chi^2$ standard deviation and the fitted power law curve. (Bottom) Significance-$\chi^2$ plane for the injections, together with the fitted mean curve (dot-dashed line) and the veto curve (dashed line) corresponding to the mean $\chi^2$ plus five times its standard deviation. } \end{figure} To characterize the veto curve, nine 0.25Hz bands, spread in frequency and free of known large spectral disturbances have been selected. These are: 102.5Hz, 151Hz, 190Hz, 252.25Hz, 314.1Hz, 448.5Hz, 504.1Hz, 610.25Hz, and 710.25Hz. Monte Carlo injections in those bands have been performed separately in the data from both years. Since the results were comparable for both years a single veto curved is derived. In total 177834 injections are considered with a significance value lower than 70. The results of these injections in terms of ($s$, $\chi^2$) are presented in figure \ref{fig:veto}. The $\chi^2$ values obtained correspond to those by splitting the data in $p=16$ segments. Then we proceed as follows: first we sort the points with respect to the significance, and we group them in sets containing 1000 points. For each set we compute the mean value of the significance, the mean of the $\chi^2$ and its standard deviation. With these reduced set of points we fit two power laws $p-1+a \,s^c$ and $\sqrt{2p-2}+b \, s^d$ to the (mean $s$, mean $\chi^2$) and (mean $s$, std $\chi^2$) respectively, obtaining the following coefficients (with $95\%$ confidence bounds): \begin{eqnarray} a = 0.3123 & \quad & (0.305, 0.3195) \nonumber \\ c = 1.777 & \quad & (1.77, 1.783) \nonumber \\ b = 0.1713 & \quad & (0.1637, 0.1789) \nonumber \\ d = 1.621 & \quad & (1.609, 1.633) \nonumber \end{eqnarray} The veto curve we will use in this analysis corresponds to the mean curve plus five times the standard deviation \begin{equation} \bar\chi^2= p-1 + 0.3123 \, s^{1.777} + 5 (\sqrt{2p-2} + 0.1713 \, s^{1.621}). \end{equation} This curve vetoes 25 of the 177834 injections considered with significance lower than 70, that could translate into a false dismissal rate of 0.014. In figure \ref{fig:veto} we show the fitted curves and the $ \bar\chi^2$ veto curve compared to the result of the injections. \section{Description of the all-sky search. } \label{sec:pipeline} \begin{figure}[t] \begin{center} \includegraphics[width=24pc]{Fig2.pdf} \end{center} \caption{\label{fig:pipe}Pipeline of the Hough search. } \end{figure} In this paper, we use a new pipeline to analyze the data from the fifth science run of the LIGO detectors to search for evidence of continuous gravitational waves, that might be radiated by nearby unknown rapidly spinning isolated neutron stars. Data from each of the three LIGO interferometers is used to perform the all-sky search. The key difference from previous searches is that, starting from $30~\mathrm{min}$ SFTs, we perform a multi-interferometer search analyzing separately the two years of the S5 run, and we study coincidences among the source candidates produced by the first and second years of data. Furthermore, we use a $\chi^2$ test adapted to the Hough transform searches to veto potential candidates. The pipeline is shown schematically in figure \ref{fig:pipe}. A separate search was run for each successive 0.25 Hz band within the frequency range $50\,$--$\,1000$~Hz and covering frequency time derivatives in the range $-\sci{8.9}{-10}~\mathrm{Hz}~\mathrm{s}^{-1}$ to zero. We use a uniform grid spacing equal to the size of a SFT frequency bin, $\delta f = 1/ T_{\textrm{\mbox{\tiny{coh}}}}= \sci{5.556}{-4}~\mathrm{Hz}$. The resolution $\delta \dot{f}$ is given by the smallest value of $\dot{f}$ for which the intrinsic signal frequency does not drift by more than a frequency bin during the observation time $T_{\textrm{\mbox{\tiny{obs}}}}$ in the first year: $\delta \dot{f} = \delta f / T_{\textrm{\mbox{\tiny{obs}}}} \sim \sci{1.8}{-11}~\mathrm{Hz}~\mathrm{s}^{-1}$. This yields 51 spin-down values for each frequency. $\delta \dot{f}$ is fixed to the same value for the search on the first and the second year of S5 data. The sky resolution, $\delta \theta$, is frequency dependent, as given by Eq.(4.14) of Ref.~\cite{Krishnan:2004sv}, that we increase by a factor 2. As explained in detail in Section V.B.1 of \cite{:2007tda}, the sky-grid spacing can be increased with a negligible loss in SNR, and for previous PowerFlux searches \cite{:2007tda, :2008rg, Abadie:2011wj} a factor 5 of increase was used in some frequency ranges to analyze LIGO S4 and S5 data. The set of SFTs are generated directly from the calibrated data stream, using 30-minute intervals of data for which the interferometer is operating in what is known as science mode. With this requirement, we search 32295 SFTs from the first year of S5 (11402 from H1, 12195 from H2 and 8698 from L1) and 35401 SFTs from the second year (12590 from H1, 12178 from H2 and 10633 from L1). \subsection{A two-step hierarchical Hough search} \label{subsec:hierarchicalHough} The approach used to analyze each year of data is based on a two-step hierarchical search for continuous signals from isolated neutron stars. In both steps, the weighted Hough transform is used to find signals whose frequency evolution fits the pattern produced by the Doppler shift and the spin-down in the time-frequency plane of the data. The search is done by splitting the frequency range in $0.25$~Hz bands and using the SFTs from multiple interferometers. \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{HistPertcentSFTs1stYr420Hz.pdf} \end{center} \caption{\label{fig:SFTs} Histograms of the percentage of SFTs that each detector has contributed in the first stage to the all-sky search. These figures correspond to a 0.25 Hz band at 420 Hz for the first year of S5 data. The vertical axes are the number of sky-patches. } \end{figure} In the first stage, and for each $0.25$~Hz band, we break up the sky into smaller patches with frequency dependent size in order to use the \textit{look up table} approach to compute the Hough transform, which greatly reduces the computational cost. The \textit{look up table} approach benefits from the fact that, according to the Doppler expression (\ref{eq:Doppler}), the set of sky positions consistent with a given frequency bin $f_k$ at a given time correspond to annuli on the celestial sphere centered on the velocity vector $\textbf{v}(t)$. In the \textit{look up table} approach, we precompute all the annuli for a given time and a given search frequency mapped on the sky search grid. Moreover, it turns out that the mapped annuli are relatively insensitive to changes in frequency and can therefore be reused a large number of times. The Hough map is then constructed by selecting the appropriate annuli out of all the ones that have been found and adding them using the corresponding weights. A detailed description of the \textit{look up table} approach with further details of implementation choices can be found in~\cite{Krishnan:2004sv}. But limitations on the memory of the computers constrain the volume of data (i.e., the number of SFTs) that can be analyzed at once and the parameter space (e.g., size and resolution of the sky-patches and number of spin-down values) we can search over. For this reason, in this first stage, we select the best 15000 SFTs (according to the noise floor and the beam pattern functions) for each frequency band and each sky-patch and apply the Hough transform on the selected data. The size of the sky-patches ranges from $\sim 0.4~\mathrm{rad} \times 0.4~\mathrm{rad}$ at $50$~Hz to $\sim 0.07 ~\mathrm{rad} \times 0.07~\mathrm{rad}$ at $1$~kHz and we calculate the weights only for the center of each sky-patch. This was set in order to ensure that the memory usage will never exceed the 0.8GB and this search could run on the Merlin/Morgane dual compute cluster at the Albert Einstein Institute\footnote{http://gw.aei.mpg.de/resources/computational-resources/merlin-morgane-dual-compute-cluster}. A top-list keeping the best $1000$ candidates is produced for each $0.25$~Hz band for the all-sky search. Figure \ref{fig:SFTs} shows the histograms of the percentage of SFTs that each detector contributes for the different sky locations for a band at 420~Hz for the first year of S5 data. At this particular frequency, the detector that contributes the most is H1 between 44--64$\%$, giving the maximum contribution near the poles, L1 contributes between 28.1--45.5$\%$ with its maximum around the equator, and H2 contributes at most $21.7\%$ of the SFTs. As shown in figure 1 in~\cite{delaJordana:2010za}, the maximum contribution of H2 corresponds to those sky regions where L1 contributes the least. If SFT selection had been based only upon the weights due to the noise floor, the H2 detector would not have contributed at all in this first stage. In a second stage, we compute the $\chi^2$ value for all the candidates in the top-list in each $0.25$~Hz band. This is done by dividing the data into 16 chunks and summing weighted binary zeros or ones along the expected path of the frequency evolution of a hypothetical periodic gravitational wave signal in the digitized time-frequency plane of our data. Since there are no computational limitations, we use the complete set of available SFTs from all three interferometers, and we also get a new value of the significance using all the data. In this way we reduce the mismatch of the template, since the number count is obtained without the roundings introduced by the \textit{look up table} approach and the weights are computed for the precise sky location and not for the center of the corresponding patch. All these refinements contribute also to a potential improvement of sensitivity when a threshold is subsequently applied to the recomputed significance (described below). \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{S5AllMaxSig025bands_v2.pdf} \end{center} \caption{\label{fig:MaxSig025} Maximum value of the significance for each $0.25$~Hz band for both years of LIGO S5 data. } \end{figure} Figure \ref{fig:MaxSig025} shows the maximum-significance value in each $0.25$~Hz band obtained for the first and second years of S5 data. \subsection{The post-processing} \label{subsec:postpro} After the multi-interferometer Hough search is performed on each year of S5 data between 50 and 1000 Hz, a top list keeping the best 1000 candidates is produced for each 0.25 Hz band. This step yields $3.8\times10^6$ candidates for each year. The post-processing of these results has the following steps: \begin{enumerate} \item Remove those 0.25 Hz bands that are affected by power lines or violin modes. A total of 96 bands are removed. These bands are given in table \ref{table:lines}. \begin{table} \caption{\label{table:lines} Initial frequency of the 0.25Hz bands excluded from the search.} \begin{indented} \item[]\begin{tabular}{@{}c c} \br Excluded Bands (Hz) & Description \\ \mr $\left[ n60-0.25,n60+0.25 \right]$ $n$=1 to 16 & Power lines \\ $\left[ 343.0 , 344.75 \right]$ & Violin modes \\ $\left[ 346.5 , 347.75 \right]$ & Violin modes \\ $\left[ 348.75, 349.25 \right]$ & Violin modes \\ $\left[ 685.75 , 689.75 \right]$ & Violin mode harmonics\\ $\left[ 693.0 , 695.5 \right]$ & Violin mode harmonics\\ $\left[ 697.5 , 698.75 \right]$ & Violin mode harmonics\\ \br \end{tabular} \end{indented} \end{table} \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{Chi2vetoLevel_v2.pdf} \end{center} \caption{\label{fig:Chi2Veto025} Percentage of the number of candidates vetoed due to a large $\chi^2$ value for each $0.25$~Hz band for both years of LIGO S5 data. } \end{figure} \item Remove all the 0.25~Hz bands for which the $\chi^2$ vetoes more than a 95$\%$ of the elements in the top list. Figure \ref{fig:Chi2Veto025} shows the $\chi^2$ veto level for all the frequency bands for both years. With this criterion, 144 and 131 0.25~Hz bands would be vetoed for the first and second year of data respectively. These first two steps leave a total 3548 bands in which we search for coincidence candidates and set upper limits; a total of 252 bands were discarded. \begin{figure}[h] \begin{center} \includegraphics[width=15pc]{templates_v2.pdf} \includegraphics[width=15pc]{threshold_v2.pdf} \end{center} \caption{\label{fig:templa} (Left) Number of templates analyzed in each 0.25~Hz band as a function of frequency. (Right) Significance threshold for a false alarm level of 1/(number of templates) (solid line), compared to 10/(number of templates) (dashed line) and 0.5/(number of templates) (dot-dashed line) in each band.} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{n2_S5CandidatesAfterVeto_2.pdf} \end{center} \caption{\label{fig:candidates} Surviving candidates from both years after applying the $\chi^2$ veto and setting a threshold in the significance. } \end{figure} \item Set a threshold on the significance. Given the relation of the Hough significance and the Hough false alarm probability (see equation (\ref{eq:sth})), we set a threshold on the candidate's significance that corresponds to a false alarm of 1/(number of templates) for each 0.25 Hz band. Figure \ref{fig:templa} shows the value of this threshold at different frequencies. \item Apply the $\chi^2$ veto. From the initial $3.8\times10^6$ elements in the top list for each year, after excluding the noisy bands, applying the $\chi^2$ veto and setting a threshold on the significance, the number of candidates remaining in the 3548 `clean' bands are 31427 for the 1st year and 50832 for the 2nd year. Those are shown in figure \ref{fig:candidates}. \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{n2b_CoincidentCandidates.pdf} \end{center} \caption{\label{fig:coincidences} Significance of the coincidence candidates from the two years of LIGO S5 data. The upper and lower plots correspond to the first and second year respectively. } \end{figure} \item Selection of coincident candidates. For each of the four parameters: frequency, spin-down and sky location, we set the coincidence window with a size equal to five times the grid spacing used in the search and centered on the values of the candidates parameters. Therefore the coincidence window always contains 625 cells in parameter space, with frequency-dependent size, according to the search grid. This window is computed for each of the candidates selected from the first year of data and then we look for coincidences among the candidates of the second year, making sure to translate their frequency to the reference time of the starting time of the run, taking into account their spin-down values. Extensive analysis of software injected signals, in different frequency bands, have been used to determine the size of this coincidence window. This was done by comparing the parameters of the most significant candidates of the search, using the same pipeline, in both years of data. With this procedure, we obtain 135728 coincidence pairs, corresponding to 5823 different candidates of the first year that have coincidences with 7234 different ones of the second year. Those are displayed in figure \ref{fig:coincidences}. All those candidates cluster in frequency in 34 groups. The most significant outlier at 108.857 Hz corresponds to a simulated pulsar signal injected into the instrument as a test signal. The most significant events in each cluster are shown in tables \ref{table:candidates1y} and \ref{table:candidates2y}. Notice how with this coincidence step the overall number of candidates has been reduced by a factor 5.4 for the first year and a factor 7.0 for the second year. Furthermore, without the coincidence step, the candidates are spread over all frequencies, whereas the surviving coincident candidates are clustered in a few small regions, illustrating the power of this procedure on real data. \begin{table} \caption{\label{table:candidates1y} Summary of 1st year coincidence candidates, including the frequency band, the number of candidates in each cluster and showing the details of the most significant candidate in each of the 34 clusters. Shown are the significance $s$, the $\chi^2$ value, the detected frequency at the start of the run (SSB frame) $f_0$, the spin down $\dot f$, and the sky position (RA, dec).} {\footnotesize \begin{tabular}{@{}r c r r r r r r r } \br & Band (Hz) & Num. & $s$ & $\chi^2$ & $f_0$ (Hz) & $\dot f$ (Hz s$^{-1}$) &RA (rad) &dec (rad) \\ \mr 1 & 50.001 - 50.003 & 12 & 7.103 & 64.115 & 50.0022 & 0 & -1.84 & 0.69 \\ 2 & 50.997 - 51.004 & 7 & 5.431 & 57.322 & 51.0028 & 0 & -2.73 & 0.94 \\ 3 & 52.000 - 52.016 & 44 & 6.886 & 54.108 & 52.0139 & -21.4e-11 & -1.99 & -0.08 \\ 4 & 52.786 - 52.793 & 6 & 6.094 & 40.089 & 52.7911 & 0 & 0 & 1.33 \\ 5 & 53.996 - 54.011 & 1136 & 11.880 & 60.583 & 54.0039 & -10.7e-11 & 2.68 & -1.21 \\ 6 & 54.996 - 55.011 & 82 & 6.995 & 48.289 & 55.0067 & 0 & 3.00 & -0.47 \\ 7 & 55.749 - 55.749 & 7 & 5.940 & 37.902 & 55.7489 & 0 & -0.23 & 1.12 \\ 8 & 56.000 - 56.016 & 167 & 8.085 & 66.130 & 56.0056 & -7.1e-11 & 0.20 & 0.22 \\ 9 & 56.997 - 57.011 & 1370 & 24.751 & 175.459 & 57.0028 & -12.5e-11 & -2.10 & 1.40 \\ 10 & 58.000 - 58.015 & 89 & 7.901 & 61.360 & 58.0128 & -35.7e-11 & -2.32 & 0.43 \\ 11 & 61.994 - 62.000 & 195 & 6.597 & 26.445 & 61.9989 & 0 & 0.98 & 0.21 \\ 12 & 62.996 - 63.009 & 1324 & 23.188 & 131.619 & 63.0028 & -8.9e-11 & 0 & -1.40 \\ 13 & 64.996 - 65.000 & 101 & 6.252 & 52.374 & 64.9994 & -5.3e-11 & -0.17 & 0.03 \\ 14 & 65.378 - 65.381 & 3 & 5.586 & 29.000 & 65.3806 & -16.1e-11 & -2.17 & 1.15 \\ 15 & 65.994 - 66.013 & 919 & 11.662 & 35.587 & 65.9994 & -7.1e-11 & 0.15 & 1.38 \\ 16 & 67.006 - 67.006 & 1 & 5.801 & 11.950 & 67.0056 & -17.8e-11 & -1.13 & -0.20 \\ 17 & 67.993 - 68.009 & 39 & 6.061 & 44.319 & 68.0017 & -14.3e-11 & -1.55 & 0.66 \\ 18 & 72.000 - 72.000 & 4 & 5.604 & 17.668 & 72.0000 & -1.8e-11 & 1.57 & -1.11 \\ 19 & 86.002 - 86.024 & 14 & 6.786 & 47.054 & 86.0150 & -17.8e-11 & 1.78 & 1.09 \\ 20 & 90.000 - 90.000 & 2 & 5.554 & 58.338 & 90.0000 & -3.6e-11 & 1.54 & -1.05 \\ 21 & 108.857 - 108.860 & 50 & 64.850 & 552.161 & 108.8570 & 0 & 3.10 & -0.60 \\ 22 & 111.998 - 111.998 & 1 & 5.673 & 46.493 & 111.9980 & -7.1e-11 & -0.54 & 1.18 \\ 23 & 118.589 - 118.613 & 18 & 7.072 & 52.181 & 118.5990 & -57.1e-11 & 2.90 & -0.46 \\ 24 & 160.000 - 160.000 & 1 & 5.571 & 18.847 & 160.0000 & 0 & 1.56 & -1.15 \\ 25 & 178.983 - 179.026 & 21 & 7.483 & 43.380 & 179.0010 & -3.6e-11 & -1.59 & 1.17 \\ 26 & 181.000 - 181.038 & 8 & 6.309 & 13.174 & 181.0170 & -8.9e-11 & -1.12 & -0.89 \\ 27 & 192.000 - 192.002 & 5 & 7.976 & 41.042 & 192.0000 & -1.8e-11 & -1.51 & 1.17 \\ 28 & 341.763 - 341.765 & 3 & 6.292 & 39.340 & 341.7630 & -1.8e-11 & -1.63 & 1.21 \\ 29 & 342.680 - 342.684 & 5 & 6.113 & 36.893 & 342.6800 & -3.6e-11 & 1.47 & -1.14 \\ 30 & 345.721 - 345.724 & 17 & 6.835 & 38.540 & 345.7230 & -12.5e-11 & -1.07 & 1.44 \\ 31 & 346.306 - 346.316 & 9 & 6.973 & 30.298 & 346.3070 & -3.6e-11 & 1.32 & -1.13 \\ 32 & 394.099 - 394.100 & 3 & 10.617 & 95.063 & 394.1000 & 0 & -1.58 & 1.16 \\ 33 & 575.163 - 575.167 & 57 & 26.058 & 146.929 & 575.1640 & -1.8e-11 & -2.53 & 0.06 \\ 34 & 671.728 - 671.733 & 101 & 13.878 & 132.197 & 671.7290 & 0 & 1.54 & -1.17 \\ \br \end{tabular} } \end{table} \begin{table} \caption{\label{table:candidates2y} Summary of 2nd year coincidence candidates, showing the details of the most significant candidate in each of the 34 clusters.} {\footnotesize \begin{tabular}{@{}r c r r r r r r r } \br & Band (Hz) & Num. & $s$ & $\chi^2$ & $f_0$ (Hz) & $\dot f$ (Hz s$^{-1}$) & RA (rad) & dec (rad) \\ \mr 1 & 50.001 - 50.004 & 10 & 9.345 & 77.005 & 50.0006 & -5.3e-11 & 1.25 & -1.05 \\ 2 & 50.993 - 51.003 & 16 & 7.106 & 17.670 & 51.0011 & -1.8e-11 & -1.95 & 1.32 \\ 3 & 52.000 - 52.012 & 38 & 8.800 & 30.888 & 52.0094 & -17.8e-11 & -2.17 & -0.08 \\ 4 & 52.784 - 52.792 & 14 & 13.269 & 128.801 & 52.7911 & -7.1e-11 & 0.77 & 1.31 \\ 5 & 53.996 - 54.006 & 1380 & 23.431 & 69.473 & 54.0033 & -14.3e-11 & -1.92 & 1.31 \\ 6 & 54.995 - 55.008 & 405 & 11.554 & 73.521 & 55.0056 & -5.3e-11 & -2.98 & 0.13 \\ 7 & 55.748 - 55.749 & 46 & 7.102 & 34.151 & 55.7489 & -1.8e-11 & -0.10 & 0.13 \\ 8 & 56.000 - 56.008 & 161 & 11.478 & 64.077 & 56.0006 & -1.8e-11 & 1.58 & -1.13 \\ 9 & 56.996 - 57.004 & 1478 & 47.422 & 412.817 & 56.9989 & 0 & -0.31 & 1.43 \\ 10 & 58.000 - 58.008 & 166 & 11.675 & 65.764 & 58.0000 & 0 & 1.59 & -1.13 \\ 11 & 61.993 - 62.001 & 400 & 9.361 & 56.108 & 61.9989 & 0 & -1.16 & 1.02 \\ 12 & 62.997 - 63.004 & 1138 & 42.432 & 360.615 & 63.0039 & -5.3e-11 & -1.51 & 1.41 \\ 13 & 64.994 - 65.000 & 288 & 12.557 & 119.361 & 64.9978 & 0 & 0.61 & -1.14 \\ 14 & 65.377 - 65.378 & 2 & 5.777 & 34.580 & 65.3772 & -14.3e-11 & -1.84 & 1.03 \\ 15 & 65.995 - 66.006 & 1162 & 20.864 & 67.221 & 66.0022 & -5.3e-11 & -1.51 & 1.40 \\ 16 & 66.999 - 66.999 & 5 & 6.207 & 19.546 & 66.9989 & -17.8e-11 & -1.08 & 0.03 \\ 17 & 67.994 - 68.007 & 199 & 11.682 & 82.662 & 68.0006 & -1.8e-11 & 1.65 & -1.14 \\ 18 & 71.999 - 72.000 & 10 & 6.802 & 45.727 & 72.0000 & -1.8e-11 & 1.41 & -1.14 \\ 19 & 86.002 - 86.013 & 15 & 7.368 & 54.707 & 86.0094 & -10.7e-11 & 2.10 & 0.75 \\ 20 & 89.999 - 90.000 & 4 & 9.008 & 86.486 & 90.0000 & 0 & 1.54 & -1.19 \\ 21 & 108.857 - 108.858 & 27 & 77.157 & 1613.230 & 108.8580 & -5.3e-11 & 2.99 & -0.71 \\ 22 & 111.996 - 111.996 & 1 & 5.775 & 59.202 & 111.9960 & -7.1e-11 & -0.37 & 1.00 \\ 23 & 118.579 - 118.589 & 19 & 7.398 & 69.372 & 118.5820 & -41.0e-11 & 2.79 & -0.68 \\ 24 & 160.000 - 160.000 & 2 & 7.898 & 15.079 & 160.0000 & 0 & 1.60 & -1.17 \\ 25 & 178.984 - 179.014 & 24 & 7.089 & 33.229 & 179.0010 & -10.7e-11 & -1.79 & 1.35 \\ 26 & 180.998 - 181.019 & 8 & 6.587 & 28.605 & 181.0180 & -28.5e-11 & -0.39 & 1.07 \\ 27 & 191.999 - 192.001 & 8 & 6.582 & 45.301 & 192.0000 & -3.6e-11 & 1.51 & -1.19 \\ 28 & 341.762 - 341.764 & 19 & 7.859 & 71.736 & 341.7630 & -3.6e-11 & -1.63 & 1.17 \\ 29 & 342.677 - 342.680 & 19 & 7.569 & 45.255 & 342.6790 & -14.3e-11 & 1.47 & -1.11 \\ 30 & 345.718 - 345.721 & 12 & 7.890 & 44.570 & 345.7200 & -12.5e-11 & -0.91 & 1.38 \\ 31 & 346.303 - 346.309 & 14 & 7.803 & 45.756 & 346.3070 & 0 & 1.46 & -1.21 \\ 32 & 394.099 - 394.100 & 2 & 9.877 & 95.232 & 394.1000 & 0 & -1.58 & 1.16 \\ 33 & 575.163 - 575.165 & 53 & 40.576 & 415.830 & 575.1640 & -1.8e-11 & -2.53 & 0.06 \\ 34 & 671.729 - 671.732 & 87 & 13.600 & 116.261 & 671.7320 & -1.8e-11 & 1.59 & -1.15 \\ \br \end{tabular} } \end{table} \end{enumerate} Noise lines were identified by previously performed searches (\cite{:2008rg, Collaboration:2009nc, Aasi:2012fw, Abadie:2011wj, Abbott:2009ws}) as well as the search described in this paper. Several techniques were used to identify the causes of outliers, including the calculation of the coherence between the interferometer output channel and physical environment monitoring channels and the computation of high resolution spectra. A dedicated analysis code ``FScan" \cite{Coughlin} was also created specifically for identification of instrumental artifacts. Problematic noise lines were recorded and monitored throughout S5. In addition, a number of particular checks were performed on the coincidence outliers, including: a detailed study of the full top-list results, for those 0.25Hz bands where the candidates were found --in order to check if candidates are more dominant in a given year, or if they cluster in certain regions of parameter space; and a second search using the data of the two most sensitive detectors, H1 and L1 separately --in order to see if artifacts could be associated to a given detector, consistent with the observed spectra. All of the 34 outliers were investigated and were all traced to instrumental artifacts or hardware injections (see details in table \ref{table:description}). Hence the search did not reveal any true continuous gravitational wave signals. \begin{table} \caption{\label{table:description} Description of the coincidence outliers, together with the maximum value of the significance in both years.} {\footnotesize \begin{tabular}{@{}r c r r c} \br & Bands (Hz) & s 1y & s 2y & Comment \\ \mr 1 & 50.001 - 50.004 & 7.103 & 9.345 & L1 1 Hz Harmonic from control/data acquisition system \\ 2 & 50.993 - 51.004 & 5.431 & 7.106 & L1 1 Hz Harmonic from control/data acquisition system \\ 3 & 52.000 - 52.016 & 6.886 & 8.800 & L1 1 Hz Harmonic from control/data acquisition system \\ 4 & 52.784 - 52.793 & 6.094 & 13.269 & Instrumental line in H1 \\ 5 & 53.996 - 54.011 & 11.880 & 23.431 & Pulsed heating sideband on 60 Hz mains \\ 6 & 54.995 - 55.011 & 6.995 & 11.554 & 1 Hz Harmonic from control/data acquisition system \\ 7 & 55.748 - 55.749 & 5.940 & 7.102 & Instrumental line in L1\\ 8 & 56.000 - 56.016 & 8.085 & 11.478 & L1 1 Hz Harmonic from control/data acquisition system \\ 9 & 56.996 - 57.011 & 24.751 & 47.422 & Pulsed heating sideband on 60 Hz mains\\ 10 & 58.000 - 58.015 & 7.901 & 11.675 & L1 1 Hz Harmonic from control/data acquisition system \\ 11 & 61.993 - 62.001 & 6.597 & 9.361 & L1 1 Hz Harmonic from control/data acquisition system \\ 12 & 62.996 - 63.009 & 23.188 & 42.432 & Pulsed heating sideband on 60 Hz mains \\ 13 & 64.994 - 65.000 & 6.252 & 12.557 & L1 1 Hz Harmonic from control/data acquisition system \\ 14 & 65.377 - 65.381 & 5.586 & 5.777 & Instrumental line in L1 -- member of offset 1Hz comb \\ 15 & 65.994 - 66.013 & 11.662 & 20.864 & Pulsed sideband on 60 Hz mains \\ 16 & 66.999 - 67.006 & 5.801 & 6.207 & L1 1 Hz Harmonic from control/data acquisition system \\ 17 & 67.993 - 68.009 & 6.061 & 11.682 & 1 Hz Harmonic from control/data acquisition system \\ 18 & 71.999 - 72.000 & 5.604 & 6.802 & 1 Hz Harmonic from control/data acquisition system \\ 19 & 86.002 - 86.024 & 6.786 & 7.368 & Instrumental line in H1 \\ 20 & 89.999 - 90.000 & 5.554 & 9.008 & Instrumental line in H1 \\ 21 & 108.857 - 108.860 & 64.850 & 77.157 & Hardware injection of simulated signal (ip3) \\ 22 & 111.996 - 111.998 & 5.673 & 5.775 & 16 Hz harmonic from data acquisition system \\ 23 & 118.579 - 118.613 & 7.072 & 7.398 & Sideband of mains at 120 Hz \\ 24 & 160.000 - 160.000 & 5.571 & 7.898 & 16 Hz harmonic from data acquisition system \\ 25 & 178.983 - 179.026 & 7.483 & 7.089 & Sideband of mains at 180 Hz \\ 26 & 180.998 - 181.038 & 6.309 & 6.587 & Sideband of mains at 180 Hz \\ 27 & 191.999 - 192.002 & 7.976 & 6.582 & 16 Hz harmonic from data acquisition system\\ 28 & 341.762 - 341.765 & 6.292 & 7.859 & Sideband of suspension wire resonance in H1 \\ 29 & 342.677 - 342.684 & 6.113 & 7.569 & Sideband of suspension wire resonance in H1 \\ 30 & 345.718 - 345.724 & 6.835 & 7.890 & Sideband of suspension wire resonance in H1 \\ 31 & 346.303 - 346.316 & 6.973 & 7.803 & Sideband of suspension wire resonance in H1 \\ 32 & 394.099 - 394.100 & 10.617 & 9.877 & Sideband of calibration line at 393.1 Hz in H1 \\ 33 & 575.163 - 575.167 & 26.058 & 40.576 & Hardware injection of simulated signal (ip2) \\ 34 & 671.728 - 671.733 & 13.878 & 13.600 & Instrumental line in H1 \\ \br \end{tabular} } \end{table} \section{Upper limits estimation and astrophysical reach} \label{sec:upperlimits} The analysis of the Hough search presented here has not identified any convincing continuous gravitational wave signal. Hence, we proceed to set upper limits on the maximum intrinsic gravitational wave strain $h_0$ that is consistent with our observations for a population of signals described by an isolated triaxial rotating neutron star. As in the previous S2 and S4 searches \cite{Abbott:2005pu, :2007tda}, we set a population-based frequentist upper limit, assuming random positions in the sky, in the gravitational wave frequency range $[50,1000]$ Hz and with spin-down values in the range $-\sci{8.9}{-10}~\mathrm{Hz}~\mathrm{s}^{-1}$ to zero. The rest of the nuisance parameters, $\cos \iota$, $\psi$ and $\phi_0$, are assumed to be uniformly distributed. As commonly done in all-sky, all-frequency searches, the upper limits are given in different frequency sub-bands, here chosen to be 0.25 Hz wide. Each upper limit is based on the most significant event from each year in its 0.25 Hz band. Our goal is to find the value of $h_0$ (denoted $h_0^{95\%}$) such that $95\%$ of the signal injections at this amplitude would be recovered by our search and are more significant than the most significant candidate from the actual search in that band, thus providing the $95\%$ confidence all-sky upper limit on $h_0$. \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{scale_ratio_v2.pdf} \end{center} \caption{\label{fig:scalefactor} Ratio of the upper limits measured by means of Monte-Carlo injections in the multi-interferometer Hough search to the quantity $h_0^{95\%}/C$ as defined in Equation~(\ref{eq:ul}). The top figure corresponds to the first year of S5 data and the bottom one to the second year. The comparison is performed by doing 500 Monte-Carlo injections for 10 different amplitude in several small frequency bands. 153 and 144 frequency bands have been used for the first and second year respectively. } \end{figure} \begin{figure}[h] \begin{center} \includegraphics[width=34pc]{marks_UL95.pdf} \end{center} \caption{\label{fig:UL} The $95\%$ confidence all-sky upper limits on $h_0$ from the hierarchical Hough multi-interferometer search together with excluded frequency bands. The best upper limits correspond to $1.0\times 10^{-24}$ for the S5 first year in the $158-158.25$~Hz band, and $8.9 \times 10^{-25}$ for the S5 second year in the $146.5-146.75$~Hz band. } \end{figure} Our procedure for setting upper-limits uses partial Monte Carlo signal injection studies, using the same search pipeline as described above, together with an analytical sensitivity estimation. As in the previous S4 Hough search \cite{:2007tda}, upper limits can be computed accurately without extensive Monte Carlo simulations. Up to a constant factor $C$, that depends on the grid resolution in parameter space, they are given by \begin{equation} \label{eq:ul} h_0^{95\%} = C \left( \frac{1}{ \sum_{i=0}^{N-1}(S^{}_{i})^{-2}}\right)^{1/4}\sqrt{\frac{ {\cal S}}{T_{\textrm{\mbox{\tiny{coh}}}}}} \,. \end{equation} where \begin{equation} \label{eq:S} {\cal S}= \textrm{erfc}^{-1}(2\alpha_\textrm{\mbox{\tiny{H}}}) + \textrm{erfc}^{-1}(2\beta_\textrm{\mbox{\tiny{H}}}) \,, \end{equation} $S_i$ is the average value of the single sided power spectral noise density of the $i^{th}$ SFT in the corresponding frequency sub-band, $\alpha_\textrm{\mbox{\tiny{H}}}$ is the false alarm and $\beta_\textrm{\mbox{\tiny{H}}}$ the false dismissal probability. The utility of this fit is that having determined the value of $C$ in a small frequency range, it can be extrapolated to cover the full bandwidth without performing any further Monte Carlo simulations. Figure~\ref{fig:scalefactor} shows the value of the constant $C$ for a number of $0.1\,$Hz frequency bands. More precisely, this is the ratio of the upper limits measured by means of Monte-Carlo injections in the multi-interferometer Hough search to the quantity $h_0^{95\%}/C$ as defined in equation~(\ref{eq:ul}). The value of ${\cal S}$ is computed using the false alarm $\alpha_\textrm{\mbox{\tiny{H}}}$ corresponding to the observed loudest event, in a given frequency band, and a false dismissal rate $\beta_\textrm{\mbox{\tiny{H}}}=0.05$, in correspondence to the desired confidence level of 95$\%$, i.e., ${\cal S}\rightarrow s^*/ \sqrt{2} + \textrm{erfc}^{-1}(0.1)$, where $s^*$ is the highest significance value in the frequency band. % This yields a scale factor $C$ of $8.32\pm0.19$ for the first year and $8.25\pm0.16$ for the second year of S5. With these values we proceed to set the upper limits for all the frequency bands. The validity of equation~(\ref{eq:ul}) was studied in \cite{:2007tda} using LIGO S4 data. In that paper upper limits were measured for each 0.25 Hz frequency band from 100 Hz to 1000 Hz using Monte Carlo injections and compared with those prescribed by this analytical approximation. Such comparison study showed that the values obtained using equation~(\ref{eq:ul}) have an error smaller than $5\%$ for bands free of large instrumental disturbances. For an in-depth study of how to analytically estimate the sensitivity of wide parameter searches for gravitational-wave pulsars, we refer the reader to \cite{Wette:2011eu}. The 95$\%$ confidence all-sky upper limits on $h_0$ from this multi-interferometer search for each year of S5 data are shown in figure \ref{fig:UL}. The best upper limits correspond to $1.0\times 10^{-24}$ for the first year of S5 in the $158-158.25$~Hz band, and $8.9 \times 10^{-25}$ for the second year in the $146.5-146.75$~Hz band. There is an overall 15\%\ calibration uncertainty on these upper limits. No upper limits are provided in the 252 vetoed bands, that were excluded from the coincidence analysis, since the analytical approximation would not be accurate enough. These excluded frequency bands are marked in the figure. \begin{figure}[t] \begin{center} \includegraphics[width=30pc]{Distance.pdf}\\ \includegraphics[width=30pc]{Ellipticity.pdf} \end{center} \caption{\label{fig:reach} These plots represent the distance range (in kpc) and the maximum ellipticity, respectively, as a function of frequency. Both plots are valid for neutron stars spinning down solely due to gravitational radiation and assuming a spin-down value of $-8.9 \times 10^{-10}$~Hz/s. In the upper plot, the excluded frequency bands for which no upper limits are provided have not been considered. } \end{figure} Figure \ref{fig:reach} provides the maximum astrophysical reach of our search for each year of the S5 run. The top panel shows the maximum distance to which we could have detected a source emitting a continuous wave signal with strain amplitude $h_0^{95\%}$. The bottom panel does not depend on any result from the search. It shows the corresponding ellipticity values as a function of frequency. For both plots the source is assumed to be spinning down at the maximum rate considered in the search $-8.9 \times 10^{-10}$~Hz/s, and emitting in gravitational waves all the energy lost. This follows formulas in paper \cite{:2007tda} and assumes the canonical value of $10^{38}$ kg m$^2$ for $I_{zz}$ in equation (\ref{eq:GWampl}). Around the frequencies of greatest sensitivity, we are sensitive to objects as far away as 1.9 and 2.2 kpc for the first and second year of S5 and with an ellipticity $\varepsilon$ around $10^{-4}$. Normal neutron stars are expected to have $\varepsilon$ less than $10^{-5}$ \cite{JohnsonMcDaniel:2012wg, Horowitz:2009ya}. Such plausible value of $\varepsilon$ could be detectable by a search like this if the object were emitting at 350 Hz and at a distance no further than 750 pc. For a source of fixed ellipticity and frequency, this search had a bit less range than the Einstein$@$Home search on the same data \cite{Aasi:2012fw}. \section{Applications of the $\chi^2$ veto and hardware-injected signals. } \label{sec:veto2} \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{n2_violin350_2Y_2.pdf} \\ \includegraphics[width=32pc]{n2_violin690_2Y_2.pdf} \end{center} \caption{\label{fig:violins} Significance and $\chi^2$ values obtained for all the elements in the top list for the second year of S5 data for the frequency bands 325-355 Hz and 685-699 Hz. Those two frequency bands include violin modes. Marked in dark red appear all the elements vetoed by the $\chi^2$ test. The solid line corresponds to the veto curve. } \end{figure} A novel feature of the search presented here is the implementation of the $\chi^2$ veto. It is worth mentioning that this discriminator has been able to veto all the violin modes present in the data and many other narrow instrumental artifacts. Figure \ref{fig:violins} demonstrates how well the $\chi^2$ veto used works on those frequency bands affected by violin modes. \begin{figure}[h] \begin{center} \includegraphics[width=32pc]{n2_P3_2Y.pdf} \end{center} \caption{\label{fig:P3} Significance and $\chi^2$ values obtained for all the elements in the top list for the second year of S5 data for 0.25 Hz band starting at 108.75 Hz that contains a hardware injected simulated pulsar signal. Marked in dark red appear all the elements vetoed by the $\chi^2$ test. The solid line corresponds to the veto curve. } \end{figure} As part of the testing and validation of search pipelines and analysis code, simulated signals are added into the interferometer length control system to produce mirror motions similar to what would be generated if a gravitational wave signal were present. These are the so-called hardware-injected pulsars. During the S5 run ten artificial pulsars were injected. Four of these pulsars: P2, P3, P5 and P8, at frequencies 575.16, 108.85, 52.81 and 193.4 Hz respectively, were strong enough to be detected by the multi-interferometer Hough search (see table III in \cite{Aasi:2012fw} for the detailed parameters). The hardware injections were not active all the time, having a duty factor of about $60\%$. The fact that these signals were not continuously present in the data caused the $\chi^2$ test to veto most of the templates associated with them, since they did not behave like the signals we were looking for. In particular, for the second year of S5, the elements of the top-list in frequency band containing P8 were vetoed by the $\chi^2$ test at the $99.4\%$ level, and therefore that band was excluded from the analysis. The bands containing injected pulsars P2 and P3 were vetoed at the $87.7\%$ and $94.5\%$ level respectively, including the most significant events. Figure \ref{fig:P3} shows the behavior of the $\chi^2$ veto for the 0.25 Hz band starting at 108.75 Hz that contains pulsar P3. In the frequency band 52.75-53.0 Hz, the candidates in the top-list were all produced by the 52.79 Hz instrumental artifact present in H1 and consequently the search failed to detect P5. This suggests that, in future analysis, smaller frequency intervals should be used to produce the top list of candidates, to prevent missing gravitational wave signals due to the presence of instrumental line-noise closeby. \section{Alternative strategies and future improvements } \label{sec:sensitivities} The search presented in this paper is more robust than but not as sensitive as the hierarchical all-sky search performed by the Einstein$@$Home distributed computing project on the same data \cite{Aasi:2012fw}, which for example, in the 0.5 Hz-wide band at 152.5 Hz, excluded the presence of signals with a $h_0$ greater than $7.6 \times 10^{-25}$ at a 90\% confidence level. This later run used the Hough transform method to combine the information from coherent searches on a time scale of about a day and it was very computationally intensive. At the same time, the Einstein$@$Home search, due to its larger coherent baseline, is more sensitive to the fact that the second spin-down is not included in the search. The Hough transform method has also proven to be more robust against transient spectral disturbances than the StackSlide or PowerFlux semi-coherent methods \cite{:2007tda}. Other strategies can be applied to perform all-sky multi-interferometer searches using the Hough transform operating on successive short Fourier transforms. In this section we estimate the sensitivity of the semi-coherent Hough search for two hypothetical searches to illustrate its capabilities, by either varying the duration of the total amount of data used in the multi-interferometer search, or by lowering the threshold for candidate selection. \begin{figure}[t] \begin{center} \includegraphics[width=32pc]{SensitivityBothYears.pdf} \\ \includegraphics[width=32pc]{Sensitivity2Year10C.pdf} \\ \end{center} \caption{\label{fig:sensit} Projected sensitivities at different confidence levels for (top) a combined search over the full S5 data using the same false alarm and (bottom) sensitivity of the second year of S5 but increasing the false alarm rate.} \end{figure} In the first case we consider a search over the full S5 data with the same criteria of selecting candidates as presented here, i.e. setting a threshold in the significance for a false alarm level equivalent to one candidate per 0.25 Hz band, but using the full data. This first search would be more sensitive since we increase the number of SFTs to search over. In this case the sensitivity can be estimated from equation (\ref{eq:ul}) and using the desired significance threshold as the 'loudest' event. In this case we should take into account that the number of templates for a two years search is double that for a single year, because of the increase of spin down values that are resolvable. This corresponds to the dot-dashed line in the significance threshold in figure \ref{fig:templa}. Different sensitivity confidence levels can also be provided by modifying the false dismissal rate in equation (\ref{eq:S}) accordingly: \begin{eqnarray} {\cal S}_{95\%}= s_\textrm{\mbox{\tiny{th}}}/\sqrt{2} + \textrm{erfc}^{-1}(0.1) \,, \nonumber\\ {\cal S}_{90\%}= s_\textrm{\mbox{\tiny{th}}}/\sqrt{2} + \textrm{erfc}^{-1}(0.2) \,, \nonumber\\ {\cal S}_{50\%}= s_\textrm{\mbox{\tiny{th}}}/\sqrt{2} + \textrm{erfc}^{-1}(1) \,. \nonumber \end{eqnarray} In the second case, we consider only the data from second year of S5 data but lower the threshold in the significance such that the false alarm would be 10 candidates per 0.25 Hz band (see dashed line in figure \ref{fig:templa}). In figure \ref{fig:sensit} we show the projected sensitivities for these two searches for different confidence levels. The best sensitivity would be for the search performed on the combined full two years of S5 data. For example in the frequency interval 159.75-160.0 Hz the estimated sensitivity levels are of $8.1 \times 10^{-25}$, $7.9 \times 10^{-25}$ and $7.1 \times 10^{-25}$ at the $95\%$, $90\%$, and $50\%$ confidence level respectively. In the second case, corresponding to an increase in false alarm rate but with a reduced amount of data, the best sensitivities are of $8.8 \times 10^{-25}$, $8.5 \times 10^{-25}$ and $7.6 \times 10^{-25}$ at the $95\%$, $90\%$, and $50\%$ confidence level respectively. Notice that although both strategies are, in principle, more sensitive than the search presented in this paper, they would produce many more candidates. These would need eliminating either by demanding coincidence between two searches of comparable sensitivity, or by follow-up using a more sensitive, computationally intensive, search. Coincidence analysis will be explored in future searches owing to the efficiency with which background noise is removed. Follow up studies will always be computationally limited. Therefore the follow up capacity will actually limit the event threshold for candidate selection. Without the inclusion of a coincidence analysis, the event threshold will have to be set higher, therefore compromising the potential sensitivity of the search itself. Moreover, for the first hypothetical case, in order to achieve the projected sensitivity, one would need to perform the search over the entire 67696 SFTs available at once. If one wanted to use the ``look up table" approach to compute the Hough transform over the two years of data, the computational cost would increase by a factor of nine with respect to the one year search presented in this paper, and the memory usage would increase from 0.8 GB to 7.2 GB, for the same sky-patch size. The memory needs could be reduced by decreasing the sky-patch size, but at additional computational cost. Another consequence of analyzing both years together is that the spin down step size in the production search would have had to be reduced significantly. There are a number of areas where further refinements could improve the sensitivity of the Hough search. In particular, one could decrease the grid spacing in parameter space in order to reduce the maximum mismatch allowed, increase the duration of the SFTs to increase the signal to noise ratio within a single SFT, the development of further veto strategies to increase the overall efficiency of the analysis, as well as the tracking and establishing of appropriate data-cleaning strategies to remove narrow-band disturbances present in the peak-grams \cite{Coughlin,1742-6596-363-1-012037,1742-6596-363-1-012024}. Several of these ideas are being addressed and will be implemented in the ``Frequency Hough all-sky search" using data from the Virgo second and fourth science runs to analyze data between 20 and 128 Hz. \section{Conclusions} \label{sec:end} In summary, we have reported the results of an all-sky search for continuous, nearly monochromatic gravitational waves on data from LIGO's fifth science run, using a new detection pipeline based on the Hough transform. The search covered the frequency range $50\,$--$\,1000$~Hz and with the frequency's time derivative in the range $-8.9 \times 10^{-10}$~Hz/s to zero. Since no evidence for gravitational waves has been observed, we have derived upper limits on the intrinsic gravitational wave strain amplitude using a standard population-based method. The best upper limits correspond to $1.0\times 10^{-24}$ for the first year of S5 in the $158-158.25$~Hz band, and $8.9 \times 10^{-25}$ for the second year in the $146.5-146.75$~Hz band (see figure \ref{fig:UL}). This new search pipeline has allowed to process outliers down to significance from 5.10 at 50 Hz to 6.13 at 1000 Hz permitting deeper searchers than in previous Hough all-sky searches \cite{:2007tda}. A set of new features have been included into the multi-detector Hough search code to be able to cope with large amounts of data and the memory limitations on the machines. In addition, a $\chi^2$ veto has been applied for the first time for continuous gravitational wave searches. This veto might be very useful for the analysis of the most recent set of data produced by the LIGO and Virgo interferometers (science runs S6, VSR2 and VSR4) whose data at lower frequencies are characterized by larger contamination of non-Gaussian noise than for S5. Although the search presented here is not the most sensitive one on the same S5 data, this paper shows the potential of the new pipeline given the advantage of the lower computational cost of the Hough search and its robustness compared to other methods, and suggests further improvements to increase the sensitivity and overall efficiency of the analysis. \ack \input L-V_ack_Feb2012.tex This document has been assigned LIGO Laboratory document number LIGO-P1300071. \section{References}
\section{Introduction} Structured classification involves learning a classifier to predict objects like trees, graphs and image segments. Such objects are usually composed of several components with complex interactions, and are hence called ``structured''. Typical structured classification techniques learn from a set of labeled training examples $\{({ \bf {x} }_i,{ \bf {y} }_i)\}_{i=1}^{l}$, where the instances ${ \bf {x} }_i$ are from an input space $\mathscr X$ and the corresponding labels ${ \bf {y} }_i$ belong to a structured output space $\mathscr Y$. Several efficient algorithms are available for fully supervised structured classification (see for \textit{e.g.}~\cite{joachims_cp, sdm}). In many practical applications, however, obtaining the label of every training example is a tedious task and we are often left with only a very few labeled training examples. When the training set contains only a few training examples with labels, and a large number of unlabeled examples, a common approach is to use semi-supervised learning methods~\citep{chapellebook}. For a set of labeled training examples $\{({ \bf {x} }_i,{ \bf {y} }_i)\}_{i=1}^{l}$, ${ \bf {x} }_i \in \mathscr X$, ${ \bf {y} }_i \in \mathscr Y$ and a set of unlabeled examples $\{{ \bf {x} }_j\}_{j=l+1}^{l+u}$, we consider the following semi-supervised learning problem: \begin{align} \min_{{{\bf {w}}}, { \bf {y} }_j^* \in \mathscr Y} \;\; \frac{1}{2} {\| {{\bf {w}}} \|^2} + \frac{C_l}{l} \sum_{i=1}^{l} L_s({ \bf {x} }_i, { \bf {y} }_i;{\bf {w}}) \nonumber \\ + \frac{C_u}{u} \sum_{j=l+1}^{l+u} L_u({ \bf {x} }_j, { \bf {y} }_j^*;{\bf {w}}) \nonumber \\ {\rm s.t.} \;\; { \bf {y} }_j^* \in \mathscr W, \; \forall \; j=l+1,\ldots,l+u, \; \nonumber \\ \mathscr W = \bigcup_{k} \mathscr W_k \nonumber \\ \label{general_semissvm} \end{align} where $L_s(\cdot)$ and $L_u(\cdot)$ denote the loss functions corresponding to the labeled and unlabeled set of examples respectively. In addition to minimizing the loss functions, we also want to ensure that the predictions ${ \bf {y} }_j^*$ over the unlabeled data satisfy a certain set of constraints: $\mathscr W=\bigcup_{k} \mathscr W_k$, determined using domain knowledge. Unlike binary or multi-class classification problem, the solution of~\eqref{general_semissvm} is hard due to each ${ \bf {y} }_j^*$ having combinatorial possibilities. Further, the constraints play a key role in semi-supervised structured output learning, as demonstrated in~\citep{codl, daso}. \cite{codl} also provide a list of constraints (Table 1 in their paper), useful in sequence labeling. Each of these constraints can be expressed using a function, $\Phi: \mathscr X \times \mathscr Y \rightarrow T$, where $T=\{0,1\}$ for hard constraint or $T=\mathbb{R}$ for soft constraints. For example, the constraint that ``a citation can only start with author'' is a hard constraint and the violation of this constraint can be denoted as $\Phi_1({ \bf {x} }_j,{ \bf {y} }_j^*)=1$. On the other hand. the constraint that ``each output at has at least one author'' can be expressed as $\Phi_2({ \bf {x} }_j, { \bf {y} }_j^*) \geq 1$. Violation of constraints can be penalized by using an appropriate constraint loss function $\mathscr C(\Phi_k({ \bf {x} }_j,{ \bf {y} }_j^*)-c)$ in the objective function. The domain constraints $\Phi(\mathscr X,\mathscr Y)$ can be further divided into two broad categories, namely the \textit{instance level} constraints, which are imposed over individual training examples, and the \textit{corpus level} constraints, imposed over the entire corpus. By extending the binary transductive SVM algorithm proposed in~\citep{joachims_transsvm} and constraining the relative frequencies of the output symbols, ~\cite{zien_trans_SSVM} proposed to solve the transductive SVM problem for structured outputs. Small improvement in performance over purely supervised learning was observed, possibly because of lack of domain dependent prior knowledge. \cite{codl} proposed a constraint-driven learning algorithm (CODL) by incorporating domain knowledge in the constraints and using a perceptron style learning algorithm. This approach resulted in high performance learning using significantly less training data.~\cite{alt_proj_bellare} proposed an alternating projection method to optimize an objective function, which used auxiliary expectation constraints. ~\cite{ganchev_PR} proposed a posterior regularization (PR) method to optimize a similar objective function. \cite{chun_nam_yu} considered transductive structural SVMs and used a convex-concave procedure to solve the resultant non-convex problem. Closely related to our proposed method is the Deterministic Annealing for Structured Output (DASO) approach proposed in~\citep{daso}. It deals with the combinatorial nature of the label space by using relaxed labeling on unlabeled data. It was found to perform better than the approaches like CODL and PR. However, DASO has not been explored for large-margin methods. Moreover, dealing with combinatorial label space is not straightforward for large-margin methods and the relaxation idea proposed in~\citep{daso} cannot be easily extended to handle large-margin formulation. This paper has the following important contributions in the context of semi-supervised large-margin structured output learning. \textbf{Contributions:} In this paper, we propose an efficient algorithm to solve semi-supervised structured output learning problem~\eqref{general_semissvm} in the large-margin setting. Alternating optimization steps (fix ${ \bf {y} }_j^*$ and solve for ${\bf {w}}$, and then fix ${\bf {w}}$ and solve for ${ \bf {y} }_j^*$) are used to solve the problem. While solving~\eqref{general_semissvm} for ${\bf {w}}$ can be done easily using any known algorithm, finding optimal ${ \bf {y} }_j^*$ for a fixed ${\bf {w}}$ requires combinatorial search. We propose an efficient and effective hill-climbing method to solve the combinatorial label switching problem. Deterministic Annealing is used in conjunction with alternating optimization to avoid poor local minima. Numerical experiments on two real-world datasets demonstrate that the proposed algorithm gives comparable or better results with those reported in~\citep{daso} and~\citep{chun_nam_yu}, thereby making the proposed algorithm a useful alternative for semi-supervised structured output learning. The paper is organized as follows. The next section discusses related work on semi-supervised learning techniques for structured output learning. Section~\ref{da_section} explains the deterministic annealing solution framework for semi-supervised training of structural SVMs with domain constraints. The label-switching procedure is elaborated in section~\ref{labelswitch_section}. Empirical results on two benchmark datasets are presented in section~\ref{experiments_section}. Section~\ref{conclusion_section} concludes the paper. \section{Related Work} \label{related_work_section} A related work to our approach is the transductive SVM (TSVM) for multi-class and hierarchical classification by~\citep{keerthi_coling}, where the idea of TSVMs in~\citep{joachims_transsvm} was extended to multi-class problems. The main challenge for multi-class problems was in designing an efficient procedure to handle the combinatorial optimization involving the labels $y_j^*$ for unlabeled examples. Note that for multi-class problems, $y_j^* \in \{1,2,\ldots,k\}$ for some $k\geq3$. \cite{keerthi_coling} showed that the combinatorial optimization for multi-class label switching results in an integer program, and proposed a transportation simplex method to solve it approximately. However, the transportation simplex method turned out to be in-efficient and an efficient label-switching procedure was given in~\citep{keerthi_coling}. A deterministic annealing method and domain constraints in the form of class-ratios were also used in the training. We note however that a straightforward extension of TSVM to structured output learning is hindered by the complexity of solving the associated label switching problem. Extending the label switching procedure to structured outputs is much more challenging, due to their complex structure and the large cardinality of the output space. Semi-supervised structural SVMs considered in ~\citep{zien_trans_SSVM} avoid the combinatorial optimization of the structured output labels, and instead consider a working set of labels. We also note that the combinatorial optimization of the label space is avoided in the recent work on transductive structural SVMs by~\citep{chun_nam_yu}; instead, a working set of cutting planes is maintained. The other related work DASO~\citep{daso} too does not consider the combinatorial problem of the label space directly; rather, it solves a problem of the following form: \begin{align} \min_{{\bf {w}},a} R({\bf {w}}) + {\mathbb{E}}_{a} L(\mathscr Y;\mathscr X, {\bf {w}}) + \mathscr C({\mathbb{E}}_{a} [\Phi(\mathscr X,\mathscr Y)] - c) \label{daso_problem} \end{align} where $a$ denotes a distribution $a(\mathscr Y)$ over the label space and $L(\mathscr Y;\mathscr X, {\bf {w}})$ is considered to be the log-linear loss. Including the distribution $a(\mathscr Y)$ avoids dealing with the original label space $\mathscr Y$ and side-steps the combinatorial problem over label space. Hence, to the best of our knowledge, no prior work exists, which tackles directly the combinatorial optimization problem involving structured outputs. \section{Semi-supervised learning of structural SVMs} \label{da_section} We consider the sequence labeling problem as a running example throughout this paper. The sequence labeling problem is a well-known structured classification problem, where a sequence of entities ${ \bf {x} }=(x^1, x^2, \ldots, x^M)$ is labeled using corresponding sequence of labels ${ \bf {y} }=(y^1, y^2, \ldots, y^M)$. The labels $\{y^j\}_{j=1}^{M}$ are assumed to be from a fixed alphabet $\Omega$ of size $|\Omega|$. Consider a structured input-output space pair $(\mathscr X,\mathscr Y)$. Given a training set of labeled examples $\{({ \bf {x} }_i,{ \bf {y} }_i)\}_{i=1}^{l} \; \in (\mathscr X \times \mathscr Y)$, structural SVMs learn a parametrized classification rule $h:\mathscr X \rightarrow \mathscr Y$ of the form $h({ \bf {x} }; {\bf {w}})=\arg\max_{{ \bf {y} }} {\bf {w}}^T f({ \bf {x} },{ \bf {y} })$ by solving the following convex optimization problem: \begin{align} \min_{{{\bf {w}}}, \xi_i \geq 0} \;\;\dsp \frac{1}{2} {\| {{\bf {w}}} \|^2} + C\sum_{i=1}^{l} \xi_i({ \bf {x} }_i,{ \bf {y} }_i,{\bf {w}}) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_i \geq \delta_i({\by}_i,{\by}) -{{\bf {w}}}^T {\Delta f_i ({\by}_i,{\by})}, \nonumber \\ \forall \;i=1,\ldots,l, \; \forall \; { \bf {y} } \in \mathscr Y. \nonumber \\ \label{ssvm_primal} \end{align} An input ${ \bf {x} }_i$ is associated with the output ${ \bf {y} }_i$ of the same length and this association is captured using the feature vector $f({ \bf {x} }_i,{ \bf {y} }_i)$. The notation $\Delta f_i ({\by}_i,{\by})$ in~\eqref{ssvm_primal} is given by $f({ \bf {x} }_i,{ \bf {y} }_i) - f({ \bf {x} }_i,{ \bf {y} })$, the difference between the feature vectors corresponding to ${ \bf {y} }_i$ and ${ \bf {y} }$, respectively. The notation $\delta_i({\by}_i,{\by})=\delta({ \bf {x} }_i,{ \bf {y} }_i,{ \bf {y} })$ is a suitable loss term. For sequence labeling applications, $\delta_i({\by}_i,{\by})$ can be chosen to be the Hamming loss function. $C>0$ is a regularization constant. With the availability of a set of unlabeled examples $\{{ \bf {x} }_j\}_{j=l+1}^{l+u} \in \mathscr X$, the semi-supervised learning problem for structural SVMs is given by: \begin{align} \min_{{{\bf {w}}}, \xi_i \geq 0, \xi_j^* \geq 0} \;\;\dsp \frac{1}{2} {\| {{\bf {w}}} \|^2} + \frac{C_l}{l} \sum_{i=1}^{l} \xi_i({ \bf {x} }_i,{ \bf {y} }_i,{\bf {w}}) \nonumber \\ + {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_i \geq \delta_i({\by}_i,{\by}) -{{\bf {w}}}^T {\Delta f_i ({\by}_i,{\by})} \; \nonumber \\ \forall \;i=1,\cdots,l, \; \forall \; { \bf {y} } \in \mathscr Y, \nonumber \\ \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \nonumber \\ \; \forall \;j=l+1,\cdots,l+u. \nonumber \\ \label{zien_s3svm_primal} \end{align} The problem~\eqref{zien_s3svm_primal} was considered in~\citep{zien_trans_SSVM}, where a working-set idea was proposed to handle the optimization with respect to $\xi_j^*$. However, domain knowledge was not incorporated into the problem~\eqref{zien_s3svm_primal}. We consider the following non-convex problem for semi-supervised learning of structural SVMs, which contain the domain constraints: \begin{align} \min_{{{\bf {w}}}, \xi_i \geq 0, \xi_j^* \geq 0} \;\;\dsp \frac{1}{2} {\| {{\bf {w}}} \|^2} + \frac{C_l}{l} \sum_{i=1}^{l} \xi_i({ \bf {x} }_i,{ \bf {y} }_i,{\bf {w}}) + \nonumber \\ {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) + \mathscr C (\Phi(\mathscr X,\mathscr Y) - c) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_i \geq \delta_i({\by}_i,{\by}) -{{\bf {w}}}^T {\Delta f_i ({\by}_i,{\by})} \nonumber \\ \forall \;i=1,\cdots,l, \; \forall \; { \bf {y} } \in \mathscr Y, \nonumber \\ \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \nonumber \\ \forall \;j=l+1,\cdots,l+u. \nonumber \\ \label{s3svm_primal} \end{align} Note that the problem~\eqref{s3svm_primal} is an extension of the semi-supervised learning problems associated with binary~\citep{joachims_transsvm} and multi-class~\citep{keerthi_coling} outputs. $C_u$, the regularization constant associated with the unlabeled examples, is chosen using an annealing procedure, which gradually increases the influence of unlabeled examples in the training~\citep{daso,keerthi_coling}. The objective function in~\eqref{s3svm_primal} is non-convex and hence we resort to an alternating optimization approach, which is an extension of the procedure given in~\citep{keerthi_coling} for semi-supervised multi-class and hierarchical classification. We note however that this extension is not easy. This will become clear when we describe the constraint matching problem. A supervised learning problem is solved to obtain an initial model ${\bf {w}}$ before the alternating optimization is performed. This supervised learning is done only with the labeled examples, by solving~\eqref{ssvm_primal}. With an initial estimate of ${\bf {w}}$ available in hand, the alternating optimization procedure starts by solving the constraint matching problem with respect to $\xi_j^*$: \begin{align} \min_{\xi_j^* \geq 0} \;\;\dsp {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) + \mathscr C (\Phi(\mathscr X,\mathscr Y) - c) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \nonumber \\ \forall \;j=l+1,\cdots,l+u. \nonumber \\ \label{constraintmatchprob} \end{align} Note that the objective function in the problem~\eqref{constraintmatchprob} is linear in $\xi_j^*$. However, it also contains the penalty term on domain constraints $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$. The constraints involving $\xi_j^*$ are not simple, and finding the quantity \begin{align} \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \label{xijstar_eqn} \end{align} becomes difficult when the domain constraints $\Phi(\mathscr X,\mathscr Y)$ are also considered. Hence an efficient procedure needs to be designed to solve~\eqref{constraintmatchprob}. In the next section, we propose an efficient label-switching procedure to find a set of candidate outputs $\{{ \bf {y} }_j^*\}_{j=l+1}^{l+u}$ for the constraint matching problem~\eqref{constraintmatchprob}. Once we find $\{{ \bf {y} }_j^*\}_{j=l+1}^{l+u}$, the alternating optimization moves on to solve~\eqref{s3svm_primal} for ${\bf {w}}$, with $\{{ \bf {y} }_j^*\}_{j=l+1}^{l+u}$ remaining fixed. After solving~\eqref{s3svm_primal}, the alternating optimization solves~\eqref{constraintmatchprob} again, for a fixed ${\bf {w}}$. This procedure is repeated until some suitable stopping criterion is satisfied. The alternating optimization procedure described above, is carried out within a deterministic annealing framework. The regularization constant $C_u$ associated with the unlabeled examples is slowly varied in the annealing step. Maintaining a fixed value of $C_u$ is often not useful, and using cross-validation to find a suitable value for $C_u$ is also not possible. Hence, the deterministic annealing framework provides a useful way to get a reasonable value of $C_u$. In our experiments, $C_u$ was varied over the range $\{10^{-4},3\times10^{-4},10^{-3},3\times10^{-3},\ldots,1\}$. We describe the alternating optimization along with the deterministic annealing procedure in Algorithm~\ref{ssl_algo}. \begin{algorithm} \caption{\textit{A Deterministic Annealing -Alternating Optimization algorithm to solve~\eqref{s3svm_primal}}} \label{ssl_algo} \begin{algorithmic}[1] \renewcommand{\algorithmicrequire}{\textbf{Input:}} \STATE Input labeled examples $\{({{ \bf {x} }}_i, {{ \bf {y} }}_i)\}^n_{i=1}$ and unlabeled examples $\{{{ \bf {x} }}_j\}^{l+u}_{j=l+1}$. \STATE Input $C_l$, the regularization constant and $\Omega$, the alphabet for labels. \STATE Set maxiter = 1000. \STATE Obtain ${\bf {w}}$ by solving the supervised learning problem in \eqref{ssvm_primal}. \FOR{$C_u = 10^{-4},3\times10^{-4},10^{-3},3\times10^{-3},\ldots,1$} \STATE iter:=1 \REPEAT \STATE Obtain ${\{{ \bf {y} }_j^*\}}_{j=l+1}^{l+u}$ for unlabeled examples by solving the constraint matching problem~\eqref{constraintmatchprob} using Algorithm~\ref{labelswitch_algo}. \label{const_match_step} \STATE Obtain ${\bf {w}}$ by solving~\eqref{s3svm_primal} with ${ \bf {y} }_j^*$ as the actual labels for ${ \bf {x} }_j, \; j=l+1,\ldots,l+u$. \STATE iter:=iter+1 \UNTIL{the labeling ${ \bf {y} }_j^*$ does not change for unlabeled examples or iter$>$maxiter} \ENDFOR \end{algorithmic} \end{algorithm} \section{Label switching algorithm to solve~\eqref{constraintmatchprob}} \label{labelswitch_section} In this section, we describe an efficient approach to solve the constraint matching problem given by~\eqref{constraintmatchprob}. We note that finding $\xi_j^*$ using the relation~\eqref{xijstar_eqn} involves solving a complex combinatorial optimization problem. Hence, a useful heuristic is to fix a ${ \bf {y} }_j^* \in \mathscr Y$, and find a corresponding ${ \bf {y} } \in \mathscr Y$, such that the quantity $\delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})}$ is maximized with respect to ${ \bf {y} }_j^*$. An important step is to find a good candidate for ${ \bf {y} }_j^*$, such that the constraint violation $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$ is minimized. We propose to choose ${ \bf {y} }_j^*$ by an iterative label-switching procedure, which we describe in detail below. Since the constraint matching problem (Step~8 in Algorithm~\ref{ssl_algo}) always follows a supervised learning step, we have an estimate of ${\bf {w}}$, before we solve the constraint matching problem. With this current estimate of ${\bf {w}}$, we can find an initial candidate ${ \bf {y} }_j^*$ for an unlabeled example ${ \bf {x} }_j$ as: \begin{align} { \bf {y} }_j^* = \arg\max_{{ \bf {y} } \in \mathscr Y} {{\bf {w}}}^T f({ \bf {x} }_j,{ \bf {y} }). \label{byjstar_initial} \end{align} The availability of candidate ${ \bf {y} }_j^*$ for all unlabeled examples $j=l+1,\ldots,l+u$, makes it possible to compute the objective term in~\eqref{constraintmatchprob}, where $\xi_j^*$ for the fixed ${ \bf {y} }_j^*$ is obtained using \begin{align} \xi_j^* = \arg\max_{{ \bf {y} } \in \mathscr Y} \delta_j({\byjstar},{\by}) - {{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})}. \label{xijstar_fixedyjstar} \end{align} Let us denote the objective value in~\eqref{constraintmatchprob} by ${\mathscr O}^*$ and let the length of the sequence ${ \bf {x} }_j=(x_j^1,x_j^2,\ldots,x_j^M)$ be $M$. We now iteratively pass over the components $x_j^m$, $\forall \; m=1,2,\ldots,M$, and switch the label components $y_j^m$. Recall that the label components $y_j^m$ are from a finite alphabet $\Omega$ of size $|\Omega|$. The switching is done randomly by replacing the label component $y_j^m$ with a new label $y^r \in \Omega$. With this replacement, we compute the constraint violation $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$ and the slack term $\xi_j^*$. If the replacement causes a decrease in the objective value ${\mathscr O}^*$, then we keep the new label $y^r$ for the $m$-th component and move on to the next component. If there is no decrease in the objective value, we ignore the replacement and keep the original label. Note that the constraint matching is handled for each replacement as follows. Whenever a new label is considered for the $m$-th component, instance level constraint violation can be checked with the new label in a straightforward way. Any corpus level constraint violation is usually decomposable over the instances and can also be handled in a simple way. Hence the constraint violation term $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$ can be computed for each replacement. This ensures that by switching the labels, we do not violate the constraints too much. Apart from choosing the label $y^r$ for the $m$-th component randomly, the component $m$ itself was randomly selected in our implementation. The switching procedure is stopped when there is no sufficient decrease in the objective value term in~\eqref{constraintmatchprob} or when a prescribed upper limit on the number of label switches is exceeded. The overall procedure is illustrated in Algorithm~\ref{labelswitch_algo}. \begin{algorithm} \caption{\textit{A label switching algorithm to solve constraint matching problem~\eqref{constraintmatchprob}}} \label{labelswitch_algo} \begin{algorithmic}[1] \renewcommand{\algorithmicrequire}{\textbf{Input:}} \STATE Input unlabeled example ${ \bf {x} }_j$ \STATE Input ${\bf {w}}$, $C_u$ \STATE Set maxswitches = 1000, numswitches=0 \FOR{$j=l+1,\ldots,l+u$} \STATE Find initial candidate ${ \bf {y} }_j^*$ by~\eqref{byjstar_initial} \STATE Compute slack $\xi_j^*$ and constraint violation $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$. \ENDFOR \STATE Calculate objective value \begin{align} {\mathscr O}^* = {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) + \mathscr C (\Phi(\mathscr X,\mathscr Y) - c). \nonumber \end{align} \FOR{$j=l+1,\ldots,l+u$} \STATE $\hat{ \bf {y} }_j = { \bf {y} }_j^*$, $M = length(\hat{ \bf {y} })$, $mincost = {\mathscr O}^*$ \FOR{$m=1,\dots,M$} \STATE $y^m$$=$$m^{th}$ label component of $\hat{ \bf {y} }$, $mincostlabel$$=$$y^m$ \FOR{$y^r \in \Omega$ and $y^r \neq y^m$} \STATE Replace $y^m$ with $y^r$. \STATE Compute $\mathscr C(\phi(\mathscr X,\mathscr Y)$$-$$c)$, the constraint violation. \STATE Find violator for ${\hat{ \bf {y} }}_j$ as \begin{align} {\bar{{ \bf {y} }}}=\arg\max_{{ \bf {y} }} \{ {\bf {w}}^T f({ \bf {x} }_j,{ \bf {y} }) + \delta_j(\hat{ \bf {y} }_j,{ \bf {y} }) \}. \nonumber \end{align} \STATE Compute slack \begin{align} \xi(\hat{ \bf {y} }_j) = \max(0, {\bf {w}}^T \Delta f_j(\hat{ \bf {y} }_j, \bar{{ \bf {y} }}) + \delta_j(\hat{ \bf {y} }_j,\bar{{ \bf {y} }})). \nonumber \end{align} \STATE Compute objective value of~\eqref{constraintmatchprob} as $\hat{\mathscr O}$. \IF {$\hat{\mathscr O}$$<$$mincost$} \STATE $mincost$ = $\hat{\mathscr O}$, \; $mincostlabel = y^r$ \ENDIF \ENDFOR \STATE Replace $m^{th}$ label of ${ \bf {y} }_j^*$ with $mincostlabel$ \STATE numswitches = numswitches + 1 \IF{numswitches $>$ maxswitches} \STATE Goto Step 30 \ENDIF \ENDFOR \ENDFOR \STATE Output ${\{{ \bf {y} }_j^*\}}_{j=l+1}^{l+u}$. \end{algorithmic} \end{algorithm} \section{Experiments and Results} \label{experiments_section} We performed experiments with the proposed semi-supervised structured classification algorithm on two benchmark sequence labeling datasets; the citations and apartment advertisements. These datasets and were originally introduced in~\citep{field_segmentation} and contain manually-labeled training and test examples. The datasets also contain a set of unlabeled examples, which is proportionally large when compared to the training set. The annealing schedule was performed as follows: the annealing temperature was started at $10^{-4}$ and increased in small steps and stopped at $1$. The evaluation is done in terms of labeling accuracy on the test data obtained by the model at the end of training. We used the sequential dual method (SDM)~\citep{sdm} for supervised learning (Step 4 in Algorithm~\ref{ssl_algo}) and compared the following methods for our experiments: \begin{itemize} \item Semi-supervised structural SVMs proposed in this paper (referred to as SSVM-SDM) \item Constraint-driven Learning~\citep{codl} (referred to as CODL) \item Deterministic Annealing for semi-supervised structured classification~\citep{daso} (referred to as DASO) \item Posterior-Regularization~\citep{ganchev_PR} (referred to as PR) \item Transductive structural SVMs~\citep{chun_nam_yu} (referred to as Trans-SSVM) \end{itemize} The apartments dataset contains 300 sequences from craigslist.org. These sequences are labeled using a set of 12 labels like features, rent, contact, photos, size and restriction. The average sequence length is 119. The citations dataset contains 500 sequences, which are citations of computer science papers. The labeling is done from a set of 13 labels like author, title, publisher, pages, and journal. The average sequence length for citation dataset is 35. The description and split sizes of the datasets are given in Table~\ref{datatable}. The partitions of citation data was taken to be the same as considered in~\cite{daso}. For apartments dataset, we considered datasets of 5, 20 and 100 labeled examples. We generated 5 random partitions for each case and provide the averaged results over these partitions. \begin{table}[!ht] \caption{\textbf{Dataset Characteristics}} \label{datatable} \begin{center} \begin{tabular}{|l|c|c|c|c|} \hline % Dataset & $n_{labeled}$ & $n_{dev}$ & $n_{unlabeled}$ & $n_{test}$ \\\hline citation & 5; 20; 300 & 100 & 1000 & 100 \\\hline apartments & 5; 20; 100 & 100 & 1000 & 100 \\\hline \end{tabular} \end{center} \end{table} \begin{table*}[t] \caption{\textbf{Comparison of average test accuracy(\%) obtained from SSVM-SDM with results in \citep{daso} (denoted by $^{\$}$) and \citep{chun_nam_yu} (denoted by $^{*}$) for Citation Dataset.} (I) denotes inductive setting, in which test examples were not used as unlabeled examples for training. (no I) denotes the setting where test examples were used as unlabeled examples for training. $^{*}$Note that different set of features were considered in~\citep{chun_nam_yu}.} \label{citations_comparison_table} \begin{center} \begin{tabular}{|l|l|l|l|l|l|l|l|} \hline % $n_{labeled}$ & Baseline CRF$^{\$}$ & Baseline SDM & DASO$^{\$}$ & SSVM-SDM & PR$^{\$}$ & CODL$^{\$}$ & Trans-SSVM$^{*}$ \\ & & & (I)& (I) & (I) & (I) & (no I) \\\hline\hline 5 & 63.1 & 66.82 & \textbf{75.2} & 74.74 & 62.7 & 71 & 72.8 \\\hline 20 & 79.1 & 78.25 & 84.9 & \textbf{86.2} & 76 & 79.4 & 81.4 \\\hline 300 & 89.9 & 91.54 & 91.1 & \textbf{92.92} & 87.29 & 88.8 & 92.8 \\\hline \end{tabular} \end{center} \end{table*} \begin{table*}[t] \caption{\textbf{Comparison of average test accuracy(\%) obtained from SSVM-SDM with results in \citep{daso} (denoted by $^{\$}$) and \citep{chun_nam_yu} (denoted by $^{*}$) for Apartments Dataset.} (I) denotes inductive setting, in which test examples were not used as unlabeled examples for training. (no I) denotes the setting where test examples were used as unlabeled examples for training. $^{*}$Note that different set of features and split sizes were considered in~\citep{chun_nam_yu}.} \label{apartments_comparison_table} \begin{center} \begin{tabular}{|l|l|l|l|l|l|l|l|} \hline % $n_{labeled}$ & Baseline CRF$^{\$}$ & Baseline SDM & DASO$^{\$}$& SSVM-SDM & PR$^{\$}$ & CODL$^{\$}$ & Trans-SSVM$^{*}$ \\ & & & (I) & (I) & (I) & (I) & (no I) \\\hline\hline 5 & 65.1 & 64.06 & 67.9 & \textbf{68.28} & 66.5 & 66 & Not Available \\\hline 20 & 72.7 & 73.63 & 76.2 & \textbf{76.37} & 74.9 & 74.6 & Not Available \\\hline 100 & 76.4 & 79.95 & 80 & \textbf{81.93} & 79 & 78.6 & 78.6 \\\hline \end{tabular} \end{center} \end{table*} \subsection{Description of the constraints} We describe the instance level and corpus level constraints considered for citations data. A similar description holds for those of apartment dataset. We used the same set of constraints given in~\citep{codl, daso}. The constraints considered are of the form $\Phi(\mathscr X,\mathscr Y)-c$, which are further sub-divided into \textit{instance level} constraints of the form \begin{align} \Phi_I(\mathscr X,\mathscr Y) = \phi_I({ \bf {x} }_j,{ \bf {y} }_j^*), \; \; j=l+1,\cdots,l+u \end{align} and the \textit{corpus level} constraints of the form $\Phi_D(\mathscr X,\mathscr Y)-c_D$ where \begin{align} \Phi_D(\mathscr X,\mathscr Y) = \sum_{j=l+1}^{l+u}\phi_D({ \bf {x} }_j,{ \bf {y} }_j^*). \end{align} We consider the following examples for instance level domain constraints. 1. \textit{AUTHOR label list can only appear at most once in each citation sequence} : For this instance level constraint, we could consider \begin{align} \phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) = \text{Number of $AUTHOR$ label lists in ${ \bf {y} }_j^*$} \nonumber \end{align} and the corresponding $c_I$ to be 1. Hence the instance level domain constraint is of the form $\phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) \leq 1$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) - 1) = r | \phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) - 1 |^2 \nonumber \end{align} where $r$ is a suitable penalty scaling factor. We used $r=1000$ for our experiments. 2. \textit{The word CA is LOCATION} : For this instance level constraint, we could consider \begin{align} \phi_{I_2}({ \bf {x} }_j,{ \bf {y} }_j^*) = \mathscr I\text{(Label for word CA in ${ \bf {y} }_j^*$} \nonumber \\ \text{== LOCATION)} \nonumber \end{align} where $\mathscr I(z)$ is the indicator function which is 1 if $z$ is true and $0$ otherwise. The corresponding $c_I$ for this constraint is set to 1. Hence the instance level domain constraint is of the form $\phi_{I_2}({ \bf {x} }_j,{ \bf {y} }_j^*) = 1$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{I_2}({ \bf {x} }_j,{ \bf {y} }_j^*) - 1) = r \nonumber \end{align} 3. \textit{Each label must be a consecutive list of words and can occur atmost only once} : For this instance level constraint, we could consider \begin{align} \phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*) = \text{Number of labels which appear} \nonumber \\ \text{more than once as disjoint lists in ${ \bf {y} }_j^*$} \nonumber \end{align} The corresponding $c_I$ for this constraint is set to 0. Hence the instance level domain constraint is of the form $\phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*) = 0$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*) ) = r |\phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*)|^2 \nonumber \end{align} Next, we consider some corpus level constraints. 1. \textit{30\% of tokens should be labeled AUTHOR} : For this corpus level constraint, we could consider \begin{align} \Phi_{D_1}(\mathscr X,\mathscr Y) = \text{Percentage of $AUTHOR$ labels in $\mathscr Y$} \nonumber \end{align} and the corresponding $c_I$ to be 30. Hence the corpus level domain constraint is of the form $\phi_{D_1}(\mathscr X,\mathscr Y) = 30$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{D_1}(\mathscr X,\mathscr Y) -30 ) = r | \phi_{D_1}(\mathscr X,\mathscr Y) - 30 |^2 \nonumber \end{align} 2. \textit{Fraction of label transitions that occur on non-punctuation characters is 0.01} : For this corpus level constraint, we could consider \begin{align} \phi_{D_2}(\mathscr X,\mathscr Y) = \text{Fraction of label transitions} \nonumber \\ \text{that occur on non-punctuation characters} \nonumber \end{align} The corresponding $c_I$ for this constraint is set to 0.01. Hence the corpus level domain constraint is of the form $\phi_{D_2}(\mathscr X,\mathscr Y) = 0.01$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{D_2}(\mathscr X,\mathscr Y) - 0.01 ) = r | \phi_{D_2}(\mathscr X,\mathscr Y) - 0.01 |. \nonumber \end{align} \subsection{Experiments on the citation data} We considered the citation dataset with 5, 20 and 300 labeled examples, along with 1000 unlabeled examples and measured the performance on a test set of 100 examples. The parameter $C_l$ was tuned using a development dataset of 100 examples. The average performance on the test set was computed by training on five different partitions for each case of 5, 20 and 300 labeled examples. The average test set accuracy comparison is presented in Table~\ref{citations_comparison_table}. The results for CODL, DASO and PR are quoted from the Inductive setting in~\citep{daso}, as the same set of features and constraints in~\citep{daso} are used for our experiments and test examples were not considered for our training. With respect to Trans-SSVMs, we quote the results for non-Inductive setting from~\citep{chun_nam_yu}, in which constraints were not used for prediction. However, we have the following important differences with Trans-SSVM in terms of the features and constraints. The feature set used for Trans-SSVM is not the same as that used for our experiments. Test examples were used for training Trans-SSVMs, which is not done for SSVM-SDM. Hence, the comparison results in Table~\ref{citations_comparison_table} for Trans-SSVM are only indicative. From the results in Table~\ref{citations_comparison_table}, we see that, for citations dataset with 5 labeled examples, the performance of SSVM-SDM is slightly worse when compared to that obtained for DASO. However, for other datasets, SSVM-SDM achieves a comparable performance. We present the plots on test accuracy and primal objective value for the partitions containing 5, 20 and 300 examples, in Figure~\ref{citation_plots}. These plots indicate that as the annealing temperature increases, the generalization performance increases initially and then continues to drop. This drop in generalization performance might possibly be the result of over-fitting caused by an inappropriate weight $C_u$ for unlabeled examples. Similar observation has been made in other semi-supervised structured output learning work using deterministic annealing~\citep{kai_wei_chang}. These observations suggest that finding a suitable stopping criterion for semi-supervised structured output learning in the deterministic annealing framework requires further study. For our comparison results, we considered the maximum test accuracy obtained from the experiments. This is indicated by a square marker in the test accuracy plots in Figure~\ref{citation_plots}. \begin{figure}[ht] \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_5_1_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_5_1_C1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_20_0_Cpoint1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_20_0_Cpoint1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_300_Cpoint1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_300_Cpoint1_testacc_withmarker.png} \end{minipage} } \caption{\textbf{Primal objective value and Test accuracy behaviour for a partition of citations dataset.} The rows correspond to 5, 20 and 300 labeled examples in that order. The square marker in the test accuracy plots denotes the best generalization performance.} \label{citation_plots} \end{figure} \begin{figure}[ht] \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_5_2_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_5_2_C1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_20_0_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_20_0_C1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_300_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_300_C1_testacc_withmarker.png} \end{minipage} } \caption{\textbf{Primal objective value and Test accuracy behaviour for a partition of apartments dataset.} The rows correspond to 5, 20 and 100 labeled examples in that order.The square marker in the test accuracy plots denotes the best generalization performance.} \label{apartments_plots} \end{figure} \subsection{Experiments on the apartments data} Experiments were performed on the apartments dataset with five partitions each for 5,20 and 100 labeled examples. 1000 unlabeled examples were considered and a test set of 100 examples was used to measure the generalization performance. The parameter $C_l$ was tuned using a development dataset of 100 examples. The average test set accuracy comparison is presented in Table~\ref{apartments_comparison_table}. For apartments dataset, though the features and constraints used in our experiments were the same as those considered in~\citep{daso}, our data partitions differ from those used in their paper. However, the comparison of mean test accuracy over the 5 different partitions for various split sizes is justified. Note also that we do not include the results with respect to Trans-SSVM for some of our experiments, as different split-sizes are considered for Trans-SSVM in~\citep{chun_nam_yu}. In particular,~\cite{chun_nam_yu} considered splits of 10, 25 and 100 labeled examples for their experiments. The results in Table~\ref{apartments_comparison_table} show that SSVM-SDM achieves a comparable average performance with DASO on all datasets. The plots on test accuracy and primal objective value for various partition sizes are given in Figure~\ref{apartments_plots}. The plots show a similar performance as seen for the citation datasets. \section{Conclusion} \label{conclusion_section} In this paper, we considered semi-supervised structural SVMs and proposed a simple and efficient algorithm to solve the resulting optimization problem. This involves solving two sub-problems alternately. One of the sub-problems is a simple supervised learning, performed by fixing the labels of the unlabeled training examples. The other sub-problem is the constraint matching problem in which suitable labeling for unlabeled examples are obtained. This was done by an efficient and effective hill-climbing procedure, which ensures that most of the domain constraints are satisfied. The alternating optimization was coupled with deterministic annealing to avoid poor local minima. The proposed algorithm is easy to implement and gives comparable generalization performance. Experimental results on real-world datasets demonstrated that the proposed algorithm is a useful alternative for semi-supervised structured output learning. The proposed label-switching method can also be used to handle complex constraints, which are imposed over only parts of the structured output. We are currently investigating this extension. \bibliographystyle{apa} \section{Introduction} Structured classification involves learning a classifier to predict objects like trees, graphs and image segments. Such objects are usually composed of several components with complex interactions, and are hence called ``structured''. Typical structured classification techniques learn from a set of labeled training examples $\{({ \bf {x} }_i,{ \bf {y} }_i)\}_{i=1}^{l}$, where the instances ${ \bf {x} }_i$ are from an input space $\mathscr X$ and the corresponding labels ${ \bf {y} }_i$ belong to a structured output space $\mathscr Y$. Several efficient algorithms are available for fully supervised structured classification (see for \textit{e.g.}~\cite{joachims_cp, sdm}). In many practical applications, however, obtaining the label of every training example is a tedious task and we are often left with only a very few labeled training examples. When the training set contains only a few training examples with labels, and a large number of unlabeled examples, a common approach is to use semi-supervised learning methods~\citep{chapellebook}. For a set of labeled training examples $\{({ \bf {x} }_i,{ \bf {y} }_i)\}_{i=1}^{l}$, ${ \bf {x} }_i \in \mathscr X$, ${ \bf {y} }_i \in \mathscr Y$ and a set of unlabeled examples $\{{ \bf {x} }_j\}_{j=l+1}^{l+u}$, we consider the following semi-supervised learning problem: \begin{align} \min_{{{\bf {w}}}, { \bf {y} }_j^* \in \mathscr Y} \;\; \frac{1}{2} {\| {{\bf {w}}} \|^2} + \frac{C_l}{l} \sum_{i=1}^{l} L_s({ \bf {x} }_i, { \bf {y} }_i;{\bf {w}}) \nonumber \\ + \frac{C_u}{u} \sum_{j=l+1}^{l+u} L_u({ \bf {x} }_j, { \bf {y} }_j^*;{\bf {w}}) \nonumber \\ {\rm s.t.} \;\; { \bf {y} }_j^* \in \mathscr W, \; \forall \; j=l+1,\ldots,l+u, \; \nonumber \\ \mathscr W = \bigcup_{k} \mathscr W_k \nonumber \\ \label{general_semissvm} \end{align} where $L_s(\cdot)$ and $L_u(\cdot)$ denote the loss functions corresponding to the labeled and unlabeled set of examples respectively. In addition to minimizing the loss functions, we also want to ensure that the predictions ${ \bf {y} }_j^*$ over the unlabeled data satisfy a certain set of constraints: $\mathscr W=\bigcup_{k} \mathscr W_k$, determined using domain knowledge. Unlike binary or multi-class classification problem, the solution of~\eqref{general_semissvm} is hard due to each ${ \bf {y} }_j^*$ having combinatorial possibilities. Further, the constraints play a key role in semi-supervised structured output learning, as demonstrated in~\citep{codl, daso}. \cite{codl} also provide a list of constraints (Table 1 in their paper), useful in sequence labeling. Each of these constraints can be expressed using a function, $\Phi: \mathscr X \times \mathscr Y \rightarrow T$, where $T=\{0,1\}$ for hard constraint or $T=\mathbb{R}$ for soft constraints. For example, the constraint that ``a citation can only start with author'' is a hard constraint and the violation of this constraint can be denoted as $\Phi_1({ \bf {x} }_j,{ \bf {y} }_j^*)=1$. On the other hand. the constraint that ``each output at has at least one author'' can be expressed as $\Phi_2({ \bf {x} }_j, { \bf {y} }_j^*) \geq 1$. Violation of constraints can be penalized by using an appropriate constraint loss function $\mathscr C(\Phi_k({ \bf {x} }_j,{ \bf {y} }_j^*)-c)$ in the objective function. The domain constraints $\Phi(\mathscr X,\mathscr Y)$ can be further divided into two broad categories, namely the \textit{instance level} constraints, which are imposed over individual training examples, and the \textit{corpus level} constraints, imposed over the entire corpus. By extending the binary transductive SVM algorithm proposed in~\citep{joachims_transsvm} and constraining the relative frequencies of the output symbols, ~\cite{zien_trans_SSVM} proposed to solve the transductive SVM problem for structured outputs. Small improvement in performance over purely supervised learning was observed, possibly because of lack of domain dependent prior knowledge. \cite{codl} proposed a constraint-driven learning algorithm (CODL) by incorporating domain knowledge in the constraints and using a perceptron style learning algorithm. This approach resulted in high performance learning using significantly less training data.~\cite{alt_proj_bellare} proposed an alternating projection method to optimize an objective function, which used auxiliary expectation constraints. ~\cite{ganchev_PR} proposed a posterior regularization (PR) method to optimize a similar objective function. \cite{chun_nam_yu} considered transductive structural SVMs and used a convex-concave procedure to solve the resultant non-convex problem. Closely related to our proposed method is the Deterministic Annealing for Structured Output (DASO) approach proposed in~\citep{daso}. It deals with the combinatorial nature of the label space by using relaxed labeling on unlabeled data. It was found to perform better than the approaches like CODL and PR. However, DASO has not been explored for large-margin methods. Moreover, dealing with combinatorial label space is not straightforward for large-margin methods and the relaxation idea proposed in~\citep{daso} cannot be easily extended to handle large-margin formulation. This paper has the following important contributions in the context of semi-supervised large-margin structured output learning. \textbf{Contributions:} In this paper, we propose an efficient algorithm to solve semi-supervised structured output learning problem~\eqref{general_semissvm} in the large-margin setting. Alternating optimization steps (fix ${ \bf {y} }_j^*$ and solve for ${\bf {w}}$, and then fix ${\bf {w}}$ and solve for ${ \bf {y} }_j^*$) are used to solve the problem. While solving~\eqref{general_semissvm} for ${\bf {w}}$ can be done easily using any known algorithm, finding optimal ${ \bf {y} }_j^*$ for a fixed ${\bf {w}}$ requires combinatorial search. We propose an efficient and effective hill-climbing method to solve the combinatorial label switching problem. Deterministic Annealing is used in conjunction with alternating optimization to avoid poor local minima. Numerical experiments on two real-world datasets demonstrate that the proposed algorithm gives comparable or better results with those reported in~\citep{daso} and~\citep{chun_nam_yu}, thereby making the proposed algorithm a useful alternative for semi-supervised structured output learning. The paper is organized as follows. The next section discusses related work on semi-supervised learning techniques for structured output learning. Section~\ref{da_section} explains the deterministic annealing solution framework for semi-supervised training of structural SVMs with domain constraints. The label-switching procedure is elaborated in section~\ref{labelswitch_section}. Empirical results on two benchmark datasets are presented in section~\ref{experiments_section}. Section~\ref{conclusion_section} concludes the paper. \section{Related Work} \label{related_work_section} A related work to our approach is the transductive SVM (TSVM) for multi-class and hierarchical classification by~\citep{keerthi_coling}, where the idea of TSVMs in~\citep{joachims_transsvm} was extended to multi-class problems. The main challenge for multi-class problems was in designing an efficient procedure to handle the combinatorial optimization involving the labels $y_j^*$ for unlabeled examples. Note that for multi-class problems, $y_j^* \in \{1,2,\ldots,k\}$ for some $k\geq3$. \cite{keerthi_coling} showed that the combinatorial optimization for multi-class label switching results in an integer program, and proposed a transportation simplex method to solve it approximately. However, the transportation simplex method turned out to be in-efficient and an efficient label-switching procedure was given in~\citep{keerthi_coling}. A deterministic annealing method and domain constraints in the form of class-ratios were also used in the training. We note however that a straightforward extension of TSVM to structured output learning is hindered by the complexity of solving the associated label switching problem. Extending the label switching procedure to structured outputs is much more challenging, due to their complex structure and the large cardinality of the output space. Semi-supervised structural SVMs considered in ~\citep{zien_trans_SSVM} avoid the combinatorial optimization of the structured output labels, and instead consider a working set of labels. We also note that the combinatorial optimization of the label space is avoided in the recent work on transductive structural SVMs by~\citep{chun_nam_yu}; instead, a working set of cutting planes is maintained. The other related work DASO~\citep{daso} too does not consider the combinatorial problem of the label space directly; rather, it solves a problem of the following form: \begin{align} \min_{{\bf {w}},a} R({\bf {w}}) + {\mathbb{E}}_{a} L(\mathscr Y;\mathscr X, {\bf {w}}) + \mathscr C({\mathbb{E}}_{a} [\Phi(\mathscr X,\mathscr Y)] - c) \label{daso_problem} \end{align} where $a$ denotes a distribution $a(\mathscr Y)$ over the label space and $L(\mathscr Y;\mathscr X, {\bf {w}})$ is considered to be the log-linear loss. Including the distribution $a(\mathscr Y)$ avoids dealing with the original label space $\mathscr Y$ and side-steps the combinatorial problem over label space. Hence, to the best of our knowledge, no prior work exists, which tackles directly the combinatorial optimization problem involving structured outputs. \section{Semi-supervised learning of structural SVMs} \label{da_section} We consider the sequence labeling problem as a running example throughout this paper. The sequence labeling problem is a well-known structured classification problem, where a sequence of entities ${ \bf {x} }=(x^1, x^2, \ldots, x^M)$ is labeled using corresponding sequence of labels ${ \bf {y} }=(y^1, y^2, \ldots, y^M)$. The labels $\{y^j\}_{j=1}^{M}$ are assumed to be from a fixed alphabet $\Omega$ of size $|\Omega|$. Consider a structured input-output space pair $(\mathscr X,\mathscr Y)$. Given a training set of labeled examples $\{({ \bf {x} }_i,{ \bf {y} }_i)\}_{i=1}^{l} \; \in (\mathscr X \times \mathscr Y)$, structural SVMs learn a parametrized classification rule $h:\mathscr X \rightarrow \mathscr Y$ of the form $h({ \bf {x} }; {\bf {w}})=\arg\max_{{ \bf {y} }} {\bf {w}}^T f({ \bf {x} },{ \bf {y} })$ by solving the following convex optimization problem: \begin{align} \min_{{{\bf {w}}}, \xi_i \geq 0} \;\;\dsp \frac{1}{2} {\| {{\bf {w}}} \|^2} + C\sum_{i=1}^{l} \xi_i({ \bf {x} }_i,{ \bf {y} }_i,{\bf {w}}) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_i \geq \delta_i({\by}_i,{\by}) -{{\bf {w}}}^T {\Delta f_i ({\by}_i,{\by})}, \nonumber \\ \forall \;i=1,\ldots,l, \; \forall \; { \bf {y} } \in \mathscr Y. \nonumber \\ \label{ssvm_primal} \end{align} An input ${ \bf {x} }_i$ is associated with the output ${ \bf {y} }_i$ of the same length and this association is captured using the feature vector $f({ \bf {x} }_i,{ \bf {y} }_i)$. The notation $\Delta f_i ({\by}_i,{\by})$ in~\eqref{ssvm_primal} is given by $f({ \bf {x} }_i,{ \bf {y} }_i) - f({ \bf {x} }_i,{ \bf {y} })$, the difference between the feature vectors corresponding to ${ \bf {y} }_i$ and ${ \bf {y} }$, respectively. The notation $\delta_i({\by}_i,{\by})=\delta({ \bf {x} }_i,{ \bf {y} }_i,{ \bf {y} })$ is a suitable loss term. For sequence labeling applications, $\delta_i({\by}_i,{\by})$ can be chosen to be the Hamming loss function. $C>0$ is a regularization constant. With the availability of a set of unlabeled examples $\{{ \bf {x} }_j\}_{j=l+1}^{l+u} \in \mathscr X$, the semi-supervised learning problem for structural SVMs is given by: \begin{align} \min_{{{\bf {w}}}, \xi_i \geq 0, \xi_j^* \geq 0} \;\;\dsp \frac{1}{2} {\| {{\bf {w}}} \|^2} + \frac{C_l}{l} \sum_{i=1}^{l} \xi_i({ \bf {x} }_i,{ \bf {y} }_i,{\bf {w}}) \nonumber \\ + {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_i \geq \delta_i({\by}_i,{\by}) -{{\bf {w}}}^T {\Delta f_i ({\by}_i,{\by})} \; \nonumber \\ \forall \;i=1,\cdots,l, \; \forall \; { \bf {y} } \in \mathscr Y, \nonumber \\ \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \nonumber \\ \; \forall \;j=l+1,\cdots,l+u. \nonumber \\ \label{zien_s3svm_primal} \end{align} The problem~\eqref{zien_s3svm_primal} was considered in~\citep{zien_trans_SSVM}, where a working-set idea was proposed to handle the optimization with respect to $\xi_j^*$. However, domain knowledge was not incorporated into the problem~\eqref{zien_s3svm_primal}. We consider the following non-convex problem for semi-supervised learning of structural SVMs, which contain the domain constraints: \begin{align} \min_{{{\bf {w}}}, \xi_i \geq 0, \xi_j^* \geq 0} \;\;\dsp \frac{1}{2} {\| {{\bf {w}}} \|^2} + \frac{C_l}{l} \sum_{i=1}^{l} \xi_i({ \bf {x} }_i,{ \bf {y} }_i,{\bf {w}}) + \nonumber \\ {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) + \mathscr C (\Phi(\mathscr X,\mathscr Y) - c) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_i \geq \delta_i({\by}_i,{\by}) -{{\bf {w}}}^T {\Delta f_i ({\by}_i,{\by})} \nonumber \\ \forall \;i=1,\cdots,l, \; \forall \; { \bf {y} } \in \mathscr Y, \nonumber \\ \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \nonumber \\ \forall \;j=l+1,\cdots,l+u. \nonumber \\ \label{s3svm_primal} \end{align} Note that the problem~\eqref{s3svm_primal} is an extension of the semi-supervised learning problems associated with binary~\citep{joachims_transsvm} and multi-class~\citep{keerthi_coling} outputs. $C_u$, the regularization constant associated with the unlabeled examples, is chosen using an annealing procedure, which gradually increases the influence of unlabeled examples in the training~\citep{daso,keerthi_coling}. The objective function in~\eqref{s3svm_primal} is non-convex and hence we resort to an alternating optimization approach, which is an extension of the procedure given in~\citep{keerthi_coling} for semi-supervised multi-class and hierarchical classification. We note however that this extension is not easy. This will become clear when we describe the constraint matching problem. A supervised learning problem is solved to obtain an initial model ${\bf {w}}$ before the alternating optimization is performed. This supervised learning is done only with the labeled examples, by solving~\eqref{ssvm_primal}. With an initial estimate of ${\bf {w}}$ available in hand, the alternating optimization procedure starts by solving the constraint matching problem with respect to $\xi_j^*$: \begin{align} \min_{\xi_j^* \geq 0} \;\;\dsp {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) + \mathscr C (\Phi(\mathscr X,\mathscr Y) - c) \nonumber \\ {\rm s.t.} \;\;\dsp \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \nonumber \\ \forall \;j=l+1,\cdots,l+u. \nonumber \\ \label{constraintmatchprob} \end{align} Note that the objective function in the problem~\eqref{constraintmatchprob} is linear in $\xi_j^*$. However, it also contains the penalty term on domain constraints $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$. The constraints involving $\xi_j^*$ are not simple, and finding the quantity \begin{align} \xi_j^* = \min_{{ \bf {y} }_j^* \in \mathscr Y} \max_{{ \bf {y} } \in \mathscr Y} \; \delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})} \label{xijstar_eqn} \end{align} becomes difficult when the domain constraints $\Phi(\mathscr X,\mathscr Y)$ are also considered. Hence an efficient procedure needs to be designed to solve~\eqref{constraintmatchprob}. In the next section, we propose an efficient label-switching procedure to find a set of candidate outputs $\{{ \bf {y} }_j^*\}_{j=l+1}^{l+u}$ for the constraint matching problem~\eqref{constraintmatchprob}. Once we find $\{{ \bf {y} }_j^*\}_{j=l+1}^{l+u}$, the alternating optimization moves on to solve~\eqref{s3svm_primal} for ${\bf {w}}$, with $\{{ \bf {y} }_j^*\}_{j=l+1}^{l+u}$ remaining fixed. After solving~\eqref{s3svm_primal}, the alternating optimization solves~\eqref{constraintmatchprob} again, for a fixed ${\bf {w}}$. This procedure is repeated until some suitable stopping criterion is satisfied. The alternating optimization procedure described above, is carried out within a deterministic annealing framework. The regularization constant $C_u$ associated with the unlabeled examples is slowly varied in the annealing step. Maintaining a fixed value of $C_u$ is often not useful, and using cross-validation to find a suitable value for $C_u$ is also not possible. Hence, the deterministic annealing framework provides a useful way to get a reasonable value of $C_u$. In our experiments, $C_u$ was varied over the range $\{10^{-4},3\times10^{-4},10^{-3},3\times10^{-3},\ldots,1\}$. We describe the alternating optimization along with the deterministic annealing procedure in Algorithm~\ref{ssl_algo}. \begin{algorithm} \caption{\textit{A Deterministic Annealing -Alternating Optimization algorithm to solve~\eqref{s3svm_primal}}} \label{ssl_algo} \begin{algorithmic}[1] \renewcommand{\algorithmicrequire}{\textbf{Input:}} \STATE Input labeled examples $\{({{ \bf {x} }}_i, {{ \bf {y} }}_i)\}^n_{i=1}$ and unlabeled examples $\{{{ \bf {x} }}_j\}^{l+u}_{j=l+1}$. \STATE Input $C_l$, the regularization constant and $\Omega$, the alphabet for labels. \STATE Set maxiter = 1000. \STATE Obtain ${\bf {w}}$ by solving the supervised learning problem in \eqref{ssvm_primal}. \FOR{$C_u = 10^{-4},3\times10^{-4},10^{-3},3\times10^{-3},\ldots,1$} \STATE iter:=1 \REPEAT \STATE Obtain ${\{{ \bf {y} }_j^*\}}_{j=l+1}^{l+u}$ for unlabeled examples by solving the constraint matching problem~\eqref{constraintmatchprob} using Algorithm~\ref{labelswitch_algo}. \label{const_match_step} \STATE Obtain ${\bf {w}}$ by solving~\eqref{s3svm_primal} with ${ \bf {y} }_j^*$ as the actual labels for ${ \bf {x} }_j, \; j=l+1,\ldots,l+u$. \STATE iter:=iter+1 \UNTIL{the labeling ${ \bf {y} }_j^*$ does not change for unlabeled examples or iter$>$maxiter} \ENDFOR \end{algorithmic} \end{algorithm} \section{Label switching algorithm to solve~\eqref{constraintmatchprob}} \label{labelswitch_section} In this section, we describe an efficient approach to solve the constraint matching problem given by~\eqref{constraintmatchprob}. We note that finding $\xi_j^*$ using the relation~\eqref{xijstar_eqn} involves solving a complex combinatorial optimization problem. Hence, a useful heuristic is to fix a ${ \bf {y} }_j^* \in \mathscr Y$, and find a corresponding ${ \bf {y} } \in \mathscr Y$, such that the quantity $\delta_j({\byjstar},{\by})-{{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})}$ is maximized with respect to ${ \bf {y} }_j^*$. An important step is to find a good candidate for ${ \bf {y} }_j^*$, such that the constraint violation $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$ is minimized. We propose to choose ${ \bf {y} }_j^*$ by an iterative label-switching procedure, which we describe in detail below. Since the constraint matching problem (Step~8 in Algorithm~\ref{ssl_algo}) always follows a supervised learning step, we have an estimate of ${\bf {w}}$, before we solve the constraint matching problem. With this current estimate of ${\bf {w}}$, we can find an initial candidate ${ \bf {y} }_j^*$ for an unlabeled example ${ \bf {x} }_j$ as: \begin{align} { \bf {y} }_j^* = \arg\max_{{ \bf {y} } \in \mathscr Y} {{\bf {w}}}^T f({ \bf {x} }_j,{ \bf {y} }). \label{byjstar_initial} \end{align} The availability of candidate ${ \bf {y} }_j^*$ for all unlabeled examples $j=l+1,\ldots,l+u$, makes it possible to compute the objective term in~\eqref{constraintmatchprob}, where $\xi_j^*$ for the fixed ${ \bf {y} }_j^*$ is obtained using \begin{align} \xi_j^* = \arg\max_{{ \bf {y} } \in \mathscr Y} \delta_j({\byjstar},{\by}) - {{\bf {w}}}^T {\Delta f_j ({\byjstar},{\by})}. \label{xijstar_fixedyjstar} \end{align} Let us denote the objective value in~\eqref{constraintmatchprob} by ${\mathscr O}^*$ and let the length of the sequence ${ \bf {x} }_j=(x_j^1,x_j^2,\ldots,x_j^M)$ be $M$. We now iteratively pass over the components $x_j^m$, $\forall \; m=1,2,\ldots,M$, and switch the label components $y_j^m$. Recall that the label components $y_j^m$ are from a finite alphabet $\Omega$ of size $|\Omega|$. The switching is done randomly by replacing the label component $y_j^m$ with a new label $y^r \in \Omega$. With this replacement, we compute the constraint violation $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$ and the slack term $\xi_j^*$. If the replacement causes a decrease in the objective value ${\mathscr O}^*$, then we keep the new label $y^r$ for the $m$-th component and move on to the next component. If there is no decrease in the objective value, we ignore the replacement and keep the original label. Note that the constraint matching is handled for each replacement as follows. Whenever a new label is considered for the $m$-th component, instance level constraint violation can be checked with the new label in a straightforward way. Any corpus level constraint violation is usually decomposable over the instances and can also be handled in a simple way. Hence the constraint violation term $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$ can be computed for each replacement. This ensures that by switching the labels, we do not violate the constraints too much. Apart from choosing the label $y^r$ for the $m$-th component randomly, the component $m$ itself was randomly selected in our implementation. The switching procedure is stopped when there is no sufficient decrease in the objective value term in~\eqref{constraintmatchprob} or when a prescribed upper limit on the number of label switches is exceeded. The overall procedure is illustrated in Algorithm~\ref{labelswitch_algo}. \begin{algorithm} \caption{\textit{A label switching algorithm to solve constraint matching problem~\eqref{constraintmatchprob}}} \label{labelswitch_algo} \begin{algorithmic}[1] \renewcommand{\algorithmicrequire}{\textbf{Input:}} \STATE Input unlabeled example ${ \bf {x} }_j$ \STATE Input ${\bf {w}}$, $C_u$ \STATE Set maxswitches = 1000, numswitches=0 \FOR{$j=l+1,\ldots,l+u$} \STATE Find initial candidate ${ \bf {y} }_j^*$ by~\eqref{byjstar_initial} \STATE Compute slack $\xi_j^*$ and constraint violation $\mathscr C (\Phi(\mathscr X,\mathscr Y) - c)$. \ENDFOR \STATE Calculate objective value \begin{align} {\mathscr O}^* = {\frac{C_u}{u}} \sum_{j=l+1}^{l+u} \xi_j^*({ \bf {x} }_j,{ \bf {y} }_j^*,{\bf {w}}) + \mathscr C (\Phi(\mathscr X,\mathscr Y) - c). \nonumber \end{align} \FOR{$j=l+1,\ldots,l+u$} \STATE $\hat{ \bf {y} }_j = { \bf {y} }_j^*$, $M = length(\hat{ \bf {y} })$, $mincost = {\mathscr O}^*$ \FOR{$m=1,\dots,M$} \STATE $y^m$$=$$m^{th}$ label component of $\hat{ \bf {y} }$, $mincostlabel$$=$$y^m$ \FOR{$y^r \in \Omega$ and $y^r \neq y^m$} \STATE Replace $y^m$ with $y^r$. \STATE Compute $\mathscr C(\phi(\mathscr X,\mathscr Y)$$-$$c)$, the constraint violation. \STATE Find violator for ${\hat{ \bf {y} }}_j$ as \begin{align} {\bar{{ \bf {y} }}}=\arg\max_{{ \bf {y} }} \{ {\bf {w}}^T f({ \bf {x} }_j,{ \bf {y} }) + \delta_j(\hat{ \bf {y} }_j,{ \bf {y} }) \}. \nonumber \end{align} \STATE Compute slack \begin{align} \xi(\hat{ \bf {y} }_j) = \max(0, {\bf {w}}^T \Delta f_j(\hat{ \bf {y} }_j, \bar{{ \bf {y} }}) + \delta_j(\hat{ \bf {y} }_j,\bar{{ \bf {y} }})). \nonumber \end{align} \STATE Compute objective value of~\eqref{constraintmatchprob} as $\hat{\mathscr O}$. \IF {$\hat{\mathscr O}$$<$$mincost$} \STATE $mincost$ = $\hat{\mathscr O}$, \; $mincostlabel = y^r$ \ENDIF \ENDFOR \STATE Replace $m^{th}$ label of ${ \bf {y} }_j^*$ with $mincostlabel$ \STATE numswitches = numswitches + 1 \IF{numswitches $>$ maxswitches} \STATE Goto Step 30 \ENDIF \ENDFOR \ENDFOR \STATE Output ${\{{ \bf {y} }_j^*\}}_{j=l+1}^{l+u}$. \end{algorithmic} \end{algorithm} \section{Experiments and Results} \label{experiments_section} We performed experiments with the proposed semi-supervised structured classification algorithm on two benchmark sequence labeling datasets; the citations and apartment advertisements. These datasets and were originally introduced in~\citep{field_segmentation} and contain manually-labeled training and test examples. The datasets also contain a set of unlabeled examples, which is proportionally large when compared to the training set. The annealing schedule was performed as follows: the annealing temperature was started at $10^{-4}$ and increased in small steps and stopped at $1$. The evaluation is done in terms of labeling accuracy on the test data obtained by the model at the end of training. We used the sequential dual method (SDM)~\citep{sdm} for supervised learning (Step 4 in Algorithm~\ref{ssl_algo}) and compared the following methods for our experiments: \begin{itemize} \item Semi-supervised structural SVMs proposed in this paper (referred to as SSVM-SDM) \item Constraint-driven Learning~\citep{codl} (referred to as CODL) \item Deterministic Annealing for semi-supervised structured classification~\citep{daso} (referred to as DASO) \item Posterior-Regularization~\citep{ganchev_PR} (referred to as PR) \item Transductive structural SVMs~\citep{chun_nam_yu} (referred to as Trans-SSVM) \end{itemize} The apartments dataset contains 300 sequences from craigslist.org. These sequences are labeled using a set of 12 labels like features, rent, contact, photos, size and restriction. The average sequence length is 119. The citations dataset contains 500 sequences, which are citations of computer science papers. The labeling is done from a set of 13 labels like author, title, publisher, pages, and journal. The average sequence length for citation dataset is 35. The description and split sizes of the datasets are given in Table~\ref{datatable}. The partitions of citation data was taken to be the same as considered in~\cite{daso}. For apartments dataset, we considered datasets of 5, 20 and 100 labeled examples. We generated 5 random partitions for each case and provide the averaged results over these partitions. \begin{table}[!ht] \caption{\textbf{Dataset Characteristics}} \label{datatable} \begin{center} \begin{tabular}{|l|c|c|c|c|} \hline % Dataset & $n_{labeled}$ & $n_{dev}$ & $n_{unlabeled}$ & $n_{test}$ \\\hline citation & 5; 20; 300 & 100 & 1000 & 100 \\\hline apartments & 5; 20; 100 & 100 & 1000 & 100 \\\hline \end{tabular} \end{center} \end{table} \begin{table*}[t] \caption{\textbf{Comparison of average test accuracy(\%) obtained from SSVM-SDM with results in \citep{daso} (denoted by $^{\$}$) and \citep{chun_nam_yu} (denoted by $^{*}$) for Citation Dataset.} (I) denotes inductive setting, in which test examples were not used as unlabeled examples for training. (no I) denotes the setting where test examples were used as unlabeled examples for training. $^{*}$Note that different set of features were considered in~\citep{chun_nam_yu}.} \label{citations_comparison_table} \begin{center} \begin{tabular}{|l|l|l|l|l|l|l|l|} \hline % $n_{labeled}$ & Baseline CRF$^{\$}$ & Baseline SDM & DASO$^{\$}$ & SSVM-SDM & PR$^{\$}$ & CODL$^{\$}$ & Trans-SSVM$^{*}$ \\ & & & (I)& (I) & (I) & (I) & (no I) \\\hline\hline 5 & 63.1 & 66.82 & \textbf{75.2} & 74.74 & 62.7 & 71 & 72.8 \\\hline 20 & 79.1 & 78.25 & 84.9 & \textbf{86.2} & 76 & 79.4 & 81.4 \\\hline 300 & 89.9 & 91.54 & 91.1 & \textbf{92.92} & 87.29 & 88.8 & 92.8 \\\hline \end{tabular} \end{center} \end{table*} \begin{table*}[t] \caption{\textbf{Comparison of average test accuracy(\%) obtained from SSVM-SDM with results in \citep{daso} (denoted by $^{\$}$) and \citep{chun_nam_yu} (denoted by $^{*}$) for Apartments Dataset.} (I) denotes inductive setting, in which test examples were not used as unlabeled examples for training. (no I) denotes the setting where test examples were used as unlabeled examples for training. $^{*}$Note that different set of features and split sizes were considered in~\citep{chun_nam_yu}.} \label{apartments_comparison_table} \begin{center} \begin{tabular}{|l|l|l|l|l|l|l|l|} \hline % $n_{labeled}$ & Baseline CRF$^{\$}$ & Baseline SDM & DASO$^{\$}$& SSVM-SDM & PR$^{\$}$ & CODL$^{\$}$ & Trans-SSVM$^{*}$ \\ & & & (I) & (I) & (I) & (I) & (no I) \\\hline\hline 5 & 65.1 & 64.06 & 67.9 & \textbf{68.28} & 66.5 & 66 & Not Available \\\hline 20 & 72.7 & 73.63 & 76.2 & \textbf{76.37} & 74.9 & 74.6 & Not Available \\\hline 100 & 76.4 & 79.95 & 80 & \textbf{81.93} & 79 & 78.6 & 78.6 \\\hline \end{tabular} \end{center} \end{table*} \subsection{Description of the constraints} We describe the instance level and corpus level constraints considered for citations data. A similar description holds for those of apartment dataset. We used the same set of constraints given in~\citep{codl, daso}. The constraints considered are of the form $\Phi(\mathscr X,\mathscr Y)-c$, which are further sub-divided into \textit{instance level} constraints of the form \begin{align} \Phi_I(\mathscr X,\mathscr Y) = \phi_I({ \bf {x} }_j,{ \bf {y} }_j^*), \; \; j=l+1,\cdots,l+u \end{align} and the \textit{corpus level} constraints of the form $\Phi_D(\mathscr X,\mathscr Y)-c_D$ where \begin{align} \Phi_D(\mathscr X,\mathscr Y) = \sum_{j=l+1}^{l+u}\phi_D({ \bf {x} }_j,{ \bf {y} }_j^*). \end{align} We consider the following examples for instance level domain constraints. 1. \textit{AUTHOR label list can only appear at most once in each citation sequence} : For this instance level constraint, we could consider \begin{align} \phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) = \text{Number of $AUTHOR$ label lists in ${ \bf {y} }_j^*$} \nonumber \end{align} and the corresponding $c_I$ to be 1. Hence the instance level domain constraint is of the form $\phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) \leq 1$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) - 1) = r | \phi_{I_1}({ \bf {x} }_j,{ \bf {y} }_j^*) - 1 |^2 \nonumber \end{align} where $r$ is a suitable penalty scaling factor. We used $r=1000$ for our experiments. 2. \textit{The word CA is LOCATION} : For this instance level constraint, we could consider \begin{align} \phi_{I_2}({ \bf {x} }_j,{ \bf {y} }_j^*) = \mathscr I\text{(Label for word CA in ${ \bf {y} }_j^*$} \nonumber \\ \text{== LOCATION)} \nonumber \end{align} where $\mathscr I(z)$ is the indicator function which is 1 if $z$ is true and $0$ otherwise. The corresponding $c_I$ for this constraint is set to 1. Hence the instance level domain constraint is of the form $\phi_{I_2}({ \bf {x} }_j,{ \bf {y} }_j^*) = 1$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{I_2}({ \bf {x} }_j,{ \bf {y} }_j^*) - 1) = r \nonumber \end{align} 3. \textit{Each label must be a consecutive list of words and can occur atmost only once} : For this instance level constraint, we could consider \begin{align} \phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*) = \text{Number of labels which appear} \nonumber \\ \text{more than once as disjoint lists in ${ \bf {y} }_j^*$} \nonumber \end{align} The corresponding $c_I$ for this constraint is set to 0. Hence the instance level domain constraint is of the form $\phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*) = 0$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*) ) = r |\phi_{I_3}({ \bf {x} }_j,{ \bf {y} }_j^*)|^2 \nonumber \end{align} Next, we consider some corpus level constraints. 1. \textit{30\% of tokens should be labeled AUTHOR} : For this corpus level constraint, we could consider \begin{align} \Phi_{D_1}(\mathscr X,\mathscr Y) = \text{Percentage of $AUTHOR$ labels in $\mathscr Y$} \nonumber \end{align} and the corresponding $c_I$ to be 30. Hence the corpus level domain constraint is of the form $\phi_{D_1}(\mathscr X,\mathscr Y) = 30$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{D_1}(\mathscr X,\mathscr Y) -30 ) = r | \phi_{D_1}(\mathscr X,\mathscr Y) - 30 |^2 \nonumber \end{align} 2. \textit{Fraction of label transitions that occur on non-punctuation characters is 0.01} : For this corpus level constraint, we could consider \begin{align} \phi_{D_2}(\mathscr X,\mathscr Y) = \text{Fraction of label transitions} \nonumber \\ \text{that occur on non-punctuation characters} \nonumber \end{align} The corresponding $c_I$ for this constraint is set to 0.01. Hence the corpus level domain constraint is of the form $\phi_{D_2}(\mathscr X,\mathscr Y) = 0.01$. The penalty function could then be defined as \begin{align} \mathscr C(\phi_{D_2}(\mathscr X,\mathscr Y) - 0.01 ) = r | \phi_{D_2}(\mathscr X,\mathscr Y) - 0.01 |. \nonumber \end{align} \subsection{Experiments on the citation data} We considered the citation dataset with 5, 20 and 300 labeled examples, along with 1000 unlabeled examples and measured the performance on a test set of 100 examples. The parameter $C_l$ was tuned using a development dataset of 100 examples. The average performance on the test set was computed by training on five different partitions for each case of 5, 20 and 300 labeled examples. The average test set accuracy comparison is presented in Table~\ref{citations_comparison_table}. The results for CODL, DASO and PR are quoted from the Inductive setting in~\citep{daso}, as the same set of features and constraints in~\citep{daso} are used for our experiments and test examples were not considered for our training. With respect to Trans-SSVMs, we quote the results for non-Inductive setting from~\citep{chun_nam_yu}, in which constraints were not used for prediction. However, we have the following important differences with Trans-SSVM in terms of the features and constraints. The feature set used for Trans-SSVM is not the same as that used for our experiments. Test examples were used for training Trans-SSVMs, which is not done for SSVM-SDM. Hence, the comparison results in Table~\ref{citations_comparison_table} for Trans-SSVM are only indicative. From the results in Table~\ref{citations_comparison_table}, we see that, for citations dataset with 5 labeled examples, the performance of SSVM-SDM is slightly worse when compared to that obtained for DASO. However, for other datasets, SSVM-SDM achieves a comparable performance. We present the plots on test accuracy and primal objective value for the partitions containing 5, 20 and 300 examples, in Figure~\ref{citation_plots}. These plots indicate that as the annealing temperature increases, the generalization performance increases initially and then continues to drop. This drop in generalization performance might possibly be the result of over-fitting caused by an inappropriate weight $C_u$ for unlabeled examples. Similar observation has been made in other semi-supervised structured output learning work using deterministic annealing~\citep{kai_wei_chang}. These observations suggest that finding a suitable stopping criterion for semi-supervised structured output learning in the deterministic annealing framework requires further study. For our comparison results, we considered the maximum test accuracy obtained from the experiments. This is indicated by a square marker in the test accuracy plots in Figure~\ref{citation_plots}. \begin{figure}[ht] \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_5_1_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_5_1_C1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_20_0_Cpoint1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_20_0_Cpoint1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_300_Cpoint1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/citation_300_Cpoint1_testacc_withmarker.png} \end{minipage} } \caption{\textbf{Primal objective value and Test accuracy behaviour for a partition of citations dataset.} The rows correspond to 5, 20 and 300 labeled examples in that order. The square marker in the test accuracy plots denotes the best generalization performance.} \label{citation_plots} \end{figure} \begin{figure}[ht] \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_5_2_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_5_2_C1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_20_0_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_20_0_C1_testacc_withmarker.png} \end{minipage} } \subfigure{ \begin{minipage}[b]{0.45\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_300_C1_primalobj.png} \end{minipage} \begin{minipage}[b]{0.4\linewidth} \includegraphics[scale=0.24]{plots/citation_apartment_plots/ads_300_C1_testacc_withmarker.png} \end{minipage} } \caption{\textbf{Primal objective value and Test accuracy behaviour for a partition of apartments dataset.} The rows correspond to 5, 20 and 100 labeled examples in that order.The square marker in the test accuracy plots denotes the best generalization performance.} \label{apartments_plots} \end{figure} \subsection{Experiments on the apartments data} Experiments were performed on the apartments dataset with five partitions each for 5,20 and 100 labeled examples. 1000 unlabeled examples were considered and a test set of 100 examples was used to measure the generalization performance. The parameter $C_l$ was tuned using a development dataset of 100 examples. The average test set accuracy comparison is presented in Table~\ref{apartments_comparison_table}. For apartments dataset, though the features and constraints used in our experiments were the same as those considered in~\citep{daso}, our data partitions differ from those used in their paper. However, the comparison of mean test accuracy over the 5 different partitions for various split sizes is justified. Note also that we do not include the results with respect to Trans-SSVM for some of our experiments, as different split-sizes are considered for Trans-SSVM in~\citep{chun_nam_yu}. In particular,~\cite{chun_nam_yu} considered splits of 10, 25 and 100 labeled examples for their experiments. The results in Table~\ref{apartments_comparison_table} show that SSVM-SDM achieves a comparable average performance with DASO on all datasets. The plots on test accuracy and primal objective value for various partition sizes are given in Figure~\ref{apartments_plots}. The plots show a similar performance as seen for the citation datasets. \section{Conclusion} \label{conclusion_section} In this paper, we considered semi-supervised structural SVMs and proposed a simple and efficient algorithm to solve the resulting optimization problem. This involves solving two sub-problems alternately. One of the sub-problems is a simple supervised learning, performed by fixing the labels of the unlabeled training examples. The other sub-problem is the constraint matching problem in which suitable labeling for unlabeled examples are obtained. This was done by an efficient and effective hill-climbing procedure, which ensures that most of the domain constraints are satisfied. The alternating optimization was coupled with deterministic annealing to avoid poor local minima. The proposed algorithm is easy to implement and gives comparable generalization performance. Experimental results on real-world datasets demonstrated that the proposed algorithm is a useful alternative for semi-supervised structured output learning. The proposed label-switching method can also be used to handle complex constraints, which are imposed over only parts of the structured output. We are currently investigating this extension. \bibliographystyle{apa}
\section{Introduction} We give a brief account of the $G$-strand construction, which gives rise to equations for a map $\mathbb{R}\times \mathbb{R}$ into a Lie group $G$ associated to a $G$-invariant Lagrangian. Our presentation reviews our previous works \cite{ Ho-Iv-Pe,Ho-Iv1, Ho-Iv2, FDT, HoLu2013} and is aimed to illustrate the $G$-strand construction with several simple but instructive examples. The following examples are reviewed here: (i) $SO(3)$-strand equations for the so-called continuous spin chain. The equations reduce to the integrable chiral model in their simplest (bi-invariant) case. (ii) $SO(3)$ - anisotropic chiral model, which is also integrable, (iii) ${\rm Diff}(\mathbb{R})$-strand equations. These equations are in general non-integrable; however they admit solutions in $2+1$ space-time with singular support (e.g., peakons). Peakon-antipeakon collisions governed by the ${\rm Diff}(\mathbb{R})$-strand equations can be solved \emph{analytically}, and potentially they can be applied in the theory of image registration. \section{Ingredients of Euler--Poincar\'e theory for Left $G$-Invariant Lagrangians} Let $G$ be a Lie group. A map $g(t,s): \mathbb{R}\times \mathbb{R}\to G$ has two types of tangent vectors, $\dot{g} := g_t \in T G$ and $g' :=g_s \in T G$. Assume that the Lagrangian density function $ L(g,\dot{g},g') $ is left $G$-invariant. The left $G$--invariance of $L$ permits us to define $l: \mathfrak{g}\times \mathfrak{g} \rightarrow \mathbb{R}$ by \[ L(g,\dot{g},g')=L(g^{-1}g,g^{-1}\dot{g},g^{-1}g')\equiv l(g^{-1} \dot{g} , g^{-1} g' ). \] Conversely, this relation defines for any reduced lagrangian $l=l({\sf u},{\sf v}) : \mathfrak{g}\times \mathfrak{g} \rightarrow \mathbb{R} $ a left $G$-invariant function $ L : T G\times TG \rightarrow \mathbb{R} $ and a map $g(t,s): \mathbb{R}\times \mathbb{R}\to G$ such that \[ {\sf u} (t,s) := g^{ -1} g_t (t,s) =g^{ -1}\dot{g}(t,s) \quad\hbox{and}\quad {\sf v} (t,s) := g^{ -1} g_s (t,s)= g^{ -1} g' (t,s) .\] \begin{lemma} The left-invariant tangent vectors ${\sf u} (t,s)$ and ${\sf v} (t,s)$ at the identity of $G$ satisfy \begin{equation} {\sf v}_t - {\sf u}_s = -\,{\rm ad}_{\sf u}{\sf v} \,. \label{zero-curv1} \end{equation} \end{lemma} \begin{proof} The proof is standard and follows from equality of cross derivatives $g_{ts}=g_{st}$. Equation (\ref{zero-curv1}) is usually called a {\bfseries\itshape zero-curvature relation}. \end{proof} \begin{theorem} [ Euler-Poincar\'e theorem for left-invariant Lagrangians]\label{lall}$\,$ With the preceding notation, the following two statements are equivalent: \begin{enumerate} \item [{\bf i} ] Variational principle on $T G\times TG$ $\,\,$ $ \delta \int _{t_1} ^{t_2} L(g(t,s), \dot{g} (t,s), g'(t,s) ) \,ds\,dt = 0 $ holds, for variations $\delta g(t,s)$ of $ g (t,s) $ vanishing at the endpoints in $t$ and $s$. The function $g(t,s)$ satisfies Euler--Lagrange equations for $L$ on $G$, given by \begin{equation*} \label{EL-eqns} \frac{\partial L}{\partial g} - \frac{\partial}{\partial t}\frac{\partial L}{\partial g_t} - \frac{\partial}{\partial s}\frac{\partial L}{\partial g_s} = 0. \end{equation*} \item [{\bf ii} ] The constrained variational principle% \footnote{As with the basic Euler--Poincar\'e equations, this is not strictly a variational principle in the same sense as the standard Hamilton's principle. It is more like the Lagrange d'Alembert principle, because we impose the stated constraints on the variations allowed.} \begin{equation*} \label{variationalprinciple} \delta \int _{t_1} ^{t_2} l({\sf u}(t,s), {\sf v}(t,s)) \,ds\,dt = 0 \end{equation*} holds on $\mathfrak{g}\times\mathfrak{g}$, using variations of $ {\sf u} := g^{ -1} g_t (t,s)$ and ${\sf v}:= g^{ -1} g_s(t,s) $ of the forms \begin{equation*} \label{epvariations} \delta {\sf u} = \dot{{\sf w} } + {\rm ad}_{\sf u}{\sf w} \quad\hbox{and}\quad \delta {\sf v} = {\sf w}\,' + {\rm ad}_{\sf v} {\sf w} \,, \end{equation*} where ${\sf w}(t,s) :=g^{ -1}\delta g \in \mathfrak{g}$ vanishes at the endpoints. The {\bfseries\itshape Euler--Poincar\'{e}} equations hold on $\mathfrak{g}^*\times\mathfrak{g}^*$ ({\bfseries\itshape$G$-strand equations}) \begin{align*} \frac{d}{dt} \frac{\delta l}{\delta {\sf u}} - \operatorname{ad}_{{\sf u}}^{\ast} \frac{ \delta l }{ \delta {\sf u}} + \frac{d}{ds} \frac{\delta l}{\delta {\sf v}} - \operatorname{ad}_{{\sf v}}^{\ast} \frac{ \delta l }{ \delta {\sf v}} = 0 \quad\hbox{ \& }\quad \partial_{s}{\sf u} - \partial_t{\sf v} = [\,{\sf u},\,{\sf v}\,] = {\rm ad}_{\sf u}{\sf v} \label{GSeqns} \end{align*} where $({\rm ad}^*: \mathfrak{g}\times\mathfrak{g}^*\to\mathfrak{g}^*)$ is defined via $({\rm ad}:\mathfrak{g}\times\mathfrak{g}\to\mathfrak{g})$ in the dual pairing $\langle \,\cdot\,,\,\cdot\,\rangle: \mathfrak{g}^*\times\mathfrak{g}\to\mathbb{R}$ by, \bigskip \begin{align*} \left\langle {\rm ad}^*_{\sf u}\frac{\delta \ell}{\delta{\sf u}} \,,\, {\sf v} \right\rangle_\mathfrak{g} = \left\langle \frac{\delta \ell}{\delta{\sf u}} \,,\, {\rm ad}_{\sf u}{\sf v} \right\rangle_\mathfrak{g}. \end{align*} \end{enumerate} \end{theorem} In 1901 Poincar\'e in his famous work proves that, when a Lie algebra acts locally transitively on the configuration space of a Lagrangian mechanical system, the well known Euler-Lagrange equations are equivalent to a new system of differential equations defined on the product of the configuration space with the Lie algebra. These equations are called now in his honor Euler-Poincar\'e equations. In modern language the contents of the Poincar\'e's article \cite{Poincare} is presented for example in \cite{Ho2011GM2,Marle}. English translation of the article \cite{Poincare} can be found as Appendix D in \cite{Ho2011GM2}. \section{$G$-strand equations on matrix Lie algebras} Denoting ${\sf m}:=\delta \ell/\delta{\sf u}$ and ${\sf n}:=\delta \ell/\delta{\sf v}$ in $\mathfrak{g}^*$, the $G$-strand equations become $$ {\sf m}_t + {\sf n}_{s} - {\rm ad}^*_{\sf u}{\sf m} - {\rm ad}^*_{\sf v}{\sf n} =0 \quad\hbox{and}\quad \partial_t{\sf v} -\partial_{s}{\sf u} + {\rm ad}_{\sf u}{\sf v} = 0. $$ For $G$ a semisimple \emph{matrix Lie group} and $\mathfrak{g}$ its \emph{matrix Lie algebra} these equations become \begin{equation} \label{MatrAlgEq} \begin{split} {\sf m}^{T}_t + {\sf n}^{T}_{s} + {\rm ad}_{\sf u}{\sf m}^{T} + {\rm ad}_{\sf v}{\sf n}^{T} =&0, \\ \partial_t{\sf v} -\partial_{s}{\sf u} + {\rm ad}_{\sf u}{\sf v}=& 0 \end{split} \end{equation} where the ad-invariant pairing for semisimple matrix Lie algebras is given by $$\Big\langle{{{\sf m}}}\,,\,{{{\sf n}}}\Big\rangle=\frac{1}{2}\tr({\sf m}^T{\sf n}), $$the transpose gives the map between the algebra and its dual $(\,\cdot\,)^{T}: \mathfrak{g}\to\mathfrak{g}^*$. For semisimple matrix Lie groups, the adjoint operator is the matrix commutator. Examples are studied in \cite{Ho-Iv-Pe, Ho-Iv2, FDT}. \section{Lie-Poisson Hamiltonian formulation} Legendre transformation of the Lagrangian $\ell({{{\sf u}},{{\sf v}}}):\, \mathfrak{g}\times \mathfrak{g}\to\mathbb{R}$ yields the Hamiltonian $h({{{\sf m}},{{\sf v}}}):\, \mathfrak{g}^*\times \mathfrak{g}\to\mathbb{R}$ \begin{equation} h({{{\sf m}}},{{{\sf v}}}) = \Big\langle{{{\sf m}}}\,,\,{{{\sf u}}}\Big\rangle - \ell({{{\sf u}},{{\sf v}}}) \,. \label{leglagham} \vspace{-3mm} \end{equation} Its partial derivatives imply \begin{eqnarray*} \frac{\delta l}{\delta {{\sf u}}} = {{\sf m}} \,,\quad \frac{\delta h}{\delta {{\sf m}}} = {{\sf u}} \quad\hbox{and}\quad \frac{\delta h}{\delta {{\sf v}}} = -\,\frac{\delta \ell}{\delta {{\sf v}}} = {\sf v} . \end{eqnarray*} These derivatives allow one to rewrite the Euler-Poincar\'e equation solely in terms of momentum ${{\sf m}}$ as \begin{equation} \label{hameqns-so3} \begin{split} {\partial_t} {{\sf m}} &= {\rm ad}^*_{\delta h/\delta {{\sf m}}}\, {{\sf m}} + \partial_{s} \frac{\delta h}{\delta {{\sf v}}} - {\rm ad}^*_{{\sf v}}\,\frac{\delta h}{\delta {{\sf v}}} \, ,\\ \partial_t {{\sf v}} &= \partial_{s}\frac{\delta h}{\delta {{\sf m}}} - {\rm ad}_{\delta h/\delta {{\sf m}}}\,{{\sf v}} \,. \end{split} \end{equation} Assembling these equations into Lie-Poisson Hamiltonian form gives, \begin{equation} \label{LP-Ham-struct-symbol1} \frac{\partial}{\partial t} \begin{bmatrix} {{\sf m}} \\ {{\sf v}} \end{bmatrix} = \begin{bmatrix} {\rm ad}^\ast(\,\cdot\,) {{\sf m}} &\hspace{5mm} \partial_s - {\rm ad}^*_{{\sf v}} \\ \partial_s + {\rm ad}_{{\sf v}} &\hspace{5mm} 0 \end{bmatrix} % \begin{bmatrix} \delta h/\delta{{\sf m}} \\ \delta h/\delta{{\sf v}} \end{bmatrix} \end{equation} The Hamiltonian matrix in equation (\ref{LP-Ham-struct-symbol1}) also appears in the Lie-Poisson brackets for Yang-Mills plasmas, for spin glasses and for perfect complex fluids, such as liquid crystals. \section{Example: The Euler-Poincar\'e PDEs for the $SO(3)$-strand and the chiral model. The $2$-time spatial and body angular velocities on $\mathfrak{so}(3)$} Let us make the following explicit identification: \begin{equation} {\sf u}=\left(% \begin{array}{ccc} 0 & -u_3 & u_2 \\ u_3 & 0 & -u_1 \\ -u_2 & u_1 & 0 \\ \end{array}% \right)\in \mathfrak{g}\quad \leftrightarrow \quad \boldsymbol{{{\sf u}}}\equiv\left(% \begin{array}{c} u_1 \\ u_2 \\ u_3 \\ \end{array}% \right)\in \mathbb{R}^3 \label{eqomegaL} \end{equation} and similarly for $\boldsymbol{{{\sf v}}}$. In terms of the corresponding group element $g(s,t)$, describing rotation, $ {\sf u}(t,{s})=g^{-1}\partial_t g(t,{s}) $ and $ {\sf v}(t,{s})=g^{-1}\partial_{s} g(t,{s}) $ resemble $2$ body angular velocities. For $G=SO(3)$ and Lagrangian $ \ell (\boldsymbol{{{\sf u}},\,{{\sf v}}} ): \mathbb{R}^3\times\mathbb{R}^3\to\mathbb{R}, $ in $1+1$ space-time the Euler-Poincar\'e equation becomes \begin{equation} \frac{\partial}{\partial t} \frac{\delta\ell}{\delta \boldsymbol{{{\sf u}}}} + \boldsymbol{{{\sf u}}}\times\frac{\delta\ell}{\delta \boldsymbol{{{\sf u}}} } = - \left(\frac{\partial}{\partial {s}} \frac{\delta\ell}{\delta \boldsymbol{{{\sf v}}} } + {\boldsymbol{{{\sf v}}}}\times \frac{\delta\ell}{\delta \boldsymbol{{{\sf v}}} }\right) \,, \label{2timeEP-SO3} \end{equation} and its auxiliary equation becomes \begin{equation} \frac{\partial}{\partial t} \boldsymbol{{{\sf v}}} = \frac{\partial}{\partial {s}}{\boldsymbol{{{\sf u}}}} + {\boldsymbol{{{\sf v}}}}\times{\boldsymbol{{{\sf u}}}} \,. \label{aux-eqn-2timeX} \end{equation} The Hamiltonian form of these equations on $\mathfrak{so}(3)^*$ are obtained from the Legendre transform relations \begin{eqnarray*} \frac{\delta \ell}{\delta \boldsymbol{{{\sf u}}} } = \boldsymbol{{{\sf m}}} \,,\quad \frac{\delta h}{\delta \boldsymbol{{{\sf m}}}} = \boldsymbol{{{\sf u}}} \quad\hbox{and}\quad \frac{\delta h}{\delta \boldsymbol{{{\sf v}}} } = -\,\frac{\delta \ell}{\delta \boldsymbol{{{\sf v}}} } \,. \end{eqnarray*} Hence, the Euler-Poincar\'e equation implies the Lie-Poisson Hamiltonian structure in vector form \begin{equation*} \label{LP-Ham-struct-so3} \partial_t \begin{bmatrix} \boldsymbol{{{\sf m}}} \\ \boldsymbol{{{\sf v}}} \end{bmatrix} = \begin{bmatrix} \boldsymbol{{{\sf m}}}\times & \partial_s + \boldsymbol{{{\sf v}}}\times \\ \partial_s + \boldsymbol{{{\sf v}}}\times & 0 \end{bmatrix} % \begin{bmatrix} \delta h/\delta\boldsymbol{{{\sf m}}} \\ \delta h/\delta\boldsymbol{{{\sf v}}} \end{bmatrix}. \end{equation*} This Poisson structure appears in various other theories, such as complex fluids and filament dynamics. When \begin{equation} \label{spinchainLagr}\ell=\frac12 \int (\boldsymbol{{{\sf u}}}\cdot A \boldsymbol{{{\sf u}}} +\boldsymbol{{{\sf v}}}\cdot B \boldsymbol{{{\sf v}}})\,ds\end{equation} this is the $SO(3)$ \emph{spin-chain model}, which is in general non-integrable- eq. (\ref{2timeEP-SO3}) and (\ref{aux-eqn-2timeX}) give: \begin{equation} \frac{\partial}{\partial t} A \boldsymbol{{{\sf u}}} + \boldsymbol{{{\sf u}}}\times A \boldsymbol{{{\sf u}}}+ \frac{\partial}{\partial {s}} B \boldsymbol{{{\sf v}}} + {\boldsymbol{{{\sf v}}}}\times B \boldsymbol{{{\sf v}}} =0\,, \label{SO3 spin chain1} \end{equation} \begin{equation} \frac{\partial}{\partial t} \boldsymbol{{{\sf v}}} = \frac{\partial}{\partial {s}}{\boldsymbol{{{\sf u}}}} + {\boldsymbol{{{\sf v}}}}\times{\boldsymbol{{{\sf u}}}} \,.\label{SO3 spin chain2} \end{equation} When $A=-B=1$, this is the $SO(3)$ \emph{chiral model}, which is an integrable Hamiltonian system. \begin{equation} \boldsymbol{{{\sf u}}}_t - \boldsymbol{{{\sf v}}}_s =0\,, \label{SO3 chiral1} \end{equation} \begin{equation} \boldsymbol{{{\sf v}}}_t - {\boldsymbol{{{\sf u}}}}_s + {\boldsymbol{{{\sf u}}}}\times{\boldsymbol{{{\sf v}}}}=0 \,.\label{SO3 chiral2} \end{equation} \section{Integrability} Some of the $G$-strands models are well known integrable models. They have a {\it zero-curvature representation} for two operators $L$ and $M$ of the form \begin{align} L_t - M_s + [L,M] = 0, \label{ZC-comrel} \end{align}\bigskip which is the compatibility condition for a pair of linear equations \[ \psi_s = L\psi, \quad\hbox{and}\quad \psi_t = M\psi. \] For the SO(3) chiral model for example these operators are \begin{equation} \begin{split} L&=\frac{1}{4}\left[(1+\lambda)({\sf u}-{\sf v})- \left(1+\frac{1}{\lambda}\right)({\sf u}+{\sf v}) \right],\\ M&=-\frac{1}{4}\left[(1+\lambda)({\sf u}-{\sf v})+ \left(1+\frac{1}{\lambda}\right)({\sf u}+{\sf v}) \right]. \end{split} \end{equation} Another integrable matrix example: $SO(3)$ anisotropic Chiral model \cite{Ch1981} \begin{equation}\label{XY-eqn-1a} \begin{split} \partial_t{\boldsymbol{\mathsf{v}}}(t,{s}) - \partial_{s}{\boldsymbol{\mathsf{u}}}(t,{s}) +{\boldsymbol{\mathsf{u}}}\times P{\boldsymbol{\mathsf{v}}}-{\boldsymbol{\mathsf{v}}}\times P{\boldsymbol{\mathsf{u}}}=0 \,, \\ \partial_{s}{\boldsymbol{\mathsf{v}}}(t,{s}) - \partial_t{\boldsymbol{\mathsf{u}}}(t,{s}) - {\boldsymbol{\mathsf{v}}}\times P {\boldsymbol{\mathsf{v}}}+{\boldsymbol{\mathsf{u}}}\times P{\boldsymbol{\mathsf{u}}}=0 \,. \end{split} \end{equation} $P=\text{diag}(P_1,P_2,P_3)$ is a constant diagonal matrix. Under the linear change of variables \begin{equation} \boldsymbol{\mathsf{X}} = \boldsymbol{\mathsf{u}} - \boldsymbol{\mathsf{v}} \quad\hbox{and}\quad \boldsymbol{\mathsf{Y}} = -\,\boldsymbol{\mathsf{u}} - \boldsymbol{\mathsf{v}} \label{changeXY-uv} \end{equation} equations (\ref{XY-eqn-1a}) acquire the form of the following $SO(3)$ anisotropic chiral model, \begin{equation}\label{uv-eqn-1} \begin{split} \partial_t{\boldsymbol{\mathsf{X}}}(t,{s}) +\partial_{s}{\boldsymbol{\mathsf{X}}}(t,{s}) +{\boldsymbol{\mathsf{X}}}\times P{\boldsymbol{\mathsf{Y}}}&=0 \,,\\ \partial_t{\boldsymbol{\mathsf{Y}}}(t,{s}) - \partial_{s}{\boldsymbol{\mathsf{Y}}}(t,{s}) +{\boldsymbol{\mathsf{Y}}}\times P {\boldsymbol{\mathsf{X}}}&=0 \,. \end{split} \end{equation} The system (\ref{uv-eqn-1}) represents two \emph{cross-coupled} equations for ${\boldsymbol{\mathsf{X}}}$ and ${\boldsymbol{\mathsf{Y}}}$. These equations preserve the magnitudes $|{\boldsymbol{\mathsf{X}}}|^2$ and $|{\boldsymbol{\mathsf{Y}}}|^2$, so they allow the further assumption that the vector fields $(\boldsymbol{\mathsf{X}},\boldsymbol{\mathsf{Y}})$ take values on the product of unit spheres $\mathbb{S}^2 \times \mathbb{S}^2 \subset \mathbb{R}^3 \times\mathbb{R}^3$. The anisotropic chiral model is an integrable system and its Lax pair in terms of $(\boldsymbol{\mathsf{u}},\boldsymbol{\mathsf{v}})$ utilizes the following isomorphism between $\mathfrak{so}(3) \oplus \mathfrak{so}(3)$ and $\mathfrak{so}(4)$: \begin{equation} A({\boldsymbol{\mathsf{u}}},{\boldsymbol{\mathsf{v}}})=\left( \begin{matrix} 0 & u_3&-u_2 & v_1\\ -u_3 & 0&u_1 & v_2\\ u_2 & -u_1&0 & v_3 \\ -v_1 & -v_2&-v_3 & 0 \end{matrix} \right) . \label{Mso4-def} \end{equation} The system (\ref{XY-eqn-1a}) can be recovered as a compatibility condition of the operators \begin{eqnarray} L&=&\partial_{s}-A({\boldsymbol{\mathsf{v}}},{\boldsymbol{\mathsf{u}}})(\lambda\,{\rm Id}+J),\\ M&=&\partial_t-A({\boldsymbol{\mathsf{u}}},{\boldsymbol{\mathsf{v}}})(\lambda\,{\rm Id}+J), \label{L-Mpair} \end{eqnarray} where the diagonal matrix $J$ is defined by \begin{equation} J = -\frac{1}{2}\text{diag}(P_1,P_2,P_3,P_1+P_2+P_3). \label{J-def} \end{equation} This Lax pair is due to Bordag and Yanovski \cite{BoYa1995}. The $O(3)$ anisotropic chiral model can be derived as an Euler-Poincar\'e equation from a Lagrangian with quadratic kinetic and potential energy. The details are presented in \cite{Ho-Iv-Pe}. \begin{remark} If ${\sf P}={\rm Id}$, equations (\ref{XY-eqn-1a}) recover the $SO(3)$ chiral model. \end{remark} \section{ The ${\rm Diff}(\mathbb{R})$-strand system} The constructions described briefly in the previous sections can be easily generalized in cases where the Lie group is the group of the Diffeomorphisms. Consider Hamiltonian which is a right-invariant bilinear form given by the $H^1$ Sobolev inner product \begin{equation} H(u,v)\equiv \frac{1}{2}\int _{#1}{\mathbf{ #1}{M}}(uv+u_xv_x) dx. \end{equation} The manifold $#1}{\mathbf{ #1}{M}$ is $\mathbb{S}^1$ or in the case when the class of smooth functions vanishing rapidly at $\pm \infty$ is considered, we will allow $#1}{\mathbf{ #1}{M} \equiv \mathbb{R}$. Let us introduce the notation $u(g(x))\equiv u\circ g$. If $g(x)\in G$, where $G\equiv \text{Diff}(#1}{\mathbf{ #1}{M})$, then $$H(u,v)=H(u\circ g, v\circ g)$$ is a right-invariant $H^1$ metric. Let us further consider an one-parametric family of diffeomprphisms, $g(x,t)$ by defining the $t$ - evolution as \begin{equation} \dot{g}=u(g(x,t),t), \qquad g(x,0)=x, \qquad \text{i.e.} \qquad \dot{g}=u\circ g \in T_g G; \end{equation} $u=\dot{g}\circ g^{-1}\in \mathfrak{g}$, where $\mathfrak{g}$, the corresponding Lie-algebra is the algebra of vector fields, $\text{Vect}(#1}{\mathbf{ #1}{M})$. Now we recall the following result: \begin{theorem}(A. Kirillov, 1980, \cite{K81, K93}) The dual space of $\mathfrak{g}$ is a space of distributions but the subspace of local functionals, called the regular dual $\mathfrak{g}^*$ is naturally identified with the space of quadratic differentials $m(x)dx^2$ on $#1}{\mathbf{ #1}{M}$. The pairing is given for any vector field $u\partial_x\in \text{Vect}(#1}{\mathbf{ #1}{M})$ by $$\langle mdx^2, u\partial_x\rangle=\int_{#1}{\mathbf{ #1}{M}}m(x)u(x)dx$$ \end{theorem} The coadjoint action coincides with the action of a diffeomorphism on the quadratic differential: $$\text{Ad}_g^*:\quad mdx^2\mapsto m(g)g_x^2dx^2$$ and $$\text{ad}_{u}^*=2u_x+u\partial_x$$ Indeed, a simple computation shows that \begin{eqnarray} \langle\text{ad}_{u\partial_x}^* mdx^2,v\partial_x\rangle&=&\langle mdx^2,[u\partial_x,v\partial_x]\rangle=\int_{#1}{\mathbf{ #1}{M}}m(u_xv-v_xu)dx= \nonumber \\ \int_{#1}{\mathbf{ #1}{M}}v(2mu_x+um_x)dx&=&\langle(2mu_x+um_x)dx^2,v\partial_x\rangle, \nonumber \end{eqnarray} i.e. $\text{ad}_{u}^*m=2u_xm+um_x$. The ${\rm Diff}(\mathbb{R})$-strand system arises when we choose $G={\rm Diff}(\mathbb{R})$. For a two-parametric group we have two tangent vectors \begin{equation*} \partial_t{g}= u \circ g \quad\hbox{and}\quad \partial_s{g}= v \circ g \,, \end{equation*} where the symbol $\circ$ denotes composition of functions. In this right-invariant case, the $G$-strand PDE system with reduced Lagrangian $\ell(u,v)$ takes the form, \begin{align} \begin{split} \frac{\partial}{\partial t} \frac{\delta\ell}{\delta u} + \frac{\partial}{\partial s} \frac{\delta\ell}{\delta v } &= -\, {\rm ad}^*_{u}\frac{\delta\ell}{\delta u} - {\rm ad}^*_{v }\frac{\delta\ell}{\delta v } \,, \\ \frac{\partial v}{\partial t} - \frac{\partial u}{\partial s} &= {\rm ad}_u v \,. \end{split} \label{Gstrand-eqn1R} \end{align}\medskip Of course, the distinction between the maps $({u},{v}): \mathbb{R}\times \mathbb{R}\to \mathfrak{g}\times\mathfrak{g}$ and their pointwise values $({u}(t,s),{v}(t,s))\in \mathfrak{g}\times\mathfrak{g}$ is clear. Likewise, for the variational derivatives ${\delta\ell}/{\delta {{u}}}$ and ${\delta\ell}/{\delta v}$. \section{The ${\rm Diff}(\mathbb{R})$-strand Hamiltonian structure} Upon setting $m={\delta\ell}/{\delta u }$ and $n={\delta\ell}/{\delta v }$, the right-invariant ${\rm Diff}(\mathbb{R})$-strand equations in (\ref{Gstrand-eqn1R}) for maps $\mathbb{R}\times\mathbb{R}\to G={\rm Diff}(\mathbb{R})$ in one spatial dimension may be expressed as a system of two 1+2 PDEs in $(t,s,x)$, \begin{align} \begin{split} m_t + n_s &= -\, {\rm ad}^*_{u }m - {\rm ad}^*_{v }n = - (u m)_x - m u_x - (v n)_x - nv_x \,, \\ \bigskip v_t - u _s &= -\,{\rm ad}_v u = -u v_x + v u_x \,. \end{split} \label{Gstrand-eqn2R} \end{align} The Hamiltonian structure for these ${\rm Diff}(\mathbb{R})$-strand equations is obtained by Legendre transforming to \[ h(m,v)=\langle m,\, u\rangle - \ell(u,\,v) \,.\] One may then write the equations (\ref{Gstrand-eqn2R}) in Lie-Poisson Hamiltonian form as \begin{equation} \frac{d}{dt} \begin{bmatrix} m \\ v \end{bmatrix} = \begin{bmatrix} -\,{\rm ad}^*(\,\cdot\,) m &\quad \partial_s + {\rm ad}^*_v \\ \partial_s - {\rm ad}_v &\quad 0 \end{bmatrix} \begin{bmatrix} {\delta h}/{\delta m} = u \\ {\delta h}/{\delta v} = -\, n \end{bmatrix}. \label{1stHamForm} \end{equation} \section{Peakon solutions of the ${\rm Diff}(\mathbb{R})$-strand equations} With the following choice of Lagrangian, \begin{equation} \ell (u ,v ) = \frac12 \|u \|^2_{H^1} - \frac12 \|v \|^2_{H^1} \,, \label{Gstrand-pkn-Lag} \end{equation} the corresponding Hamiltonian is positive-definite and the ${\rm Diff}(\mathbb{R})$-strand equations (\ref{Gstrand-eqn2R}) admit peakon solutions in \emph{both} momenta $$m=u-u_{xx} \quad \text{ and} \quad n=-(v-v_{xx}),$$ with continuous velocities $u$ and $v$. This is a two-component generalization of the CH equation. \begin{theorem} The ${\rm Diff}(\mathbb{R})$-strand equations (\ref{Gstrand-eqn2R}) admit singular solutions expressible as linear superpositions summed over $a\in\mathbb{Z}$ \begin{align} \begin{split} m(s,t,x) &= \sum_a M_a(s,t)\delta(x-Q^a(s,t)) \,,\\ n(s,t,x) &= \sum_a N_a(s,t)\delta(x-Q^a(s,t)) \,, \\ u(s,t,x) & =K*m=\sum_a M_a(s,t) K(x,Q^a) \,,\\ v(s,t,x) & = -K*n=-\sum_a N_a(s,t) K(x,Q^a) \,, \end{split} \label{Gstrand-singsolns1} \end{align} that are \emph{peakons} in the case that $K(x,y)= \frac12 e^{-|x-y|}$ is the Green function the inverse Helmholtz operator $(1-\partial_x^2)$: $$(1-\partial_x^2)K(x,0)=\delta(x) $$ \end{theorem} The solution parameters $\{Q^a(s,t), M_a(s,t), N_a(s,t)\}$ with $a\in\mathbb{Z}$ that specify the singular solutions (\ref{Gstrand-singsolns1}) are determined by the following set of evolutionary PDEs in $s$ and $t$, in which we denote $ K^{ab}:=K(Q^a,Q^b) $ with integer summation indices $a,b,c,e\in\mathbb{Z}$: \begin{align} \begin{split} \partial_t Q^a(s,t) &= u(Q^a,s,t) = \sum_b M_b(s,t) K^{ab} \,,\\ \partial_s Q^a(s,t) &= v(Q^a,s,t) =-\sum_b N_b(s,t) K^{ab} \,,\\ \partial_t M_a(s,t) &= -\, \partial_s N_a -\sum_c (M_aM_c-N_aN_c) \frac{\partial K^{ac}}{\partial Q^a} \quad\hbox{(no sum on $a$),} \\ \partial_t N_a(s,t) &=-\partial_s M_a + \sum_{b,c,e} (N_bM_c - M_bN_c) \frac{\partial K^{ec}}{\partial Q^e} (K^{eb}-K^{cb})(K^{-1})_{ae} \,. \end{split} \label{Gstrand-eqns} \end{align} The last pair of equations in (\ref{Gstrand-eqns}) may be solved as a system for the momenta, i.e., Lagrange multipliers $(M_a,N_a)$, then used in the previous pair to update the support set of positions $Q^a(t,s)$. \section{Single-peakon solution of the of the ${\rm Diff}(\mathbb{R})$-strand system} The single-peakon solution of the ${\rm Diff}(\mathbb{R})$-strand equations (\ref{Gstrand-eqn2R}) is straightforward to obtain from (\ref{Gstrand-eqns}). Combining the equations in (\ref{Gstrand-eqns}) for a single peakon shows that $Q^1(s,t)$ satisfies the Laplace equation, \begin{equation*} (\partial_s^2 - \partial _t^2)Q^1(s,t)=0\,. \label{Q-1-peak} \end{equation*} Thus, any function $h(s,t)$ that solves the wave equation provides a solution $Q^1=h(s,t)$. From the first two equations in (\ref{Gstrand-eqns}) \begin{equation*} M_1(s,t)=\frac{1}{K_0}h_t (s,t) \qquad N_1(s,t)=\frac{1}{K_0}h_s (s,t), \label{MN-1-peak} \end{equation*} where $K_0=K(0,0)$. The solutions for the single-peakon parameters $Q^1, M_1$ and $N_1$ depend only on one function $h(s,t)$, which in turn depends on the $(s,t)$ boundary conditions. The shape of the Green's function comes into the corresponding solutions for the peakon profiles \[ u(s,t,x) = M_1(s,t) K(x,Q^1(s,t)) \,,\qquad v(s,t,x) =- N_1(s,t) K(x,Q^1(s,t)) \,. \] \section{Peakon-Antipeakon collisions on a ${\rm Diff}(\mathbb{R})$-strand} Denote the relative spacing $X(s,t)=Q^1-Q^2$ for the peakons at positions $Q^1(t,s)$ and $Q^2(t,s)$ on the real line and the Green's function $K=K(X)$. Then the first two equations in (\ref{Gstrand-eqns}) imply \begin{align} \begin{split} \partial_t X &= (M_1-M_2)(K_0-K(X)) \,,\\ \partial_s X &= - (N_1-N_2)(K_0-K(X)) \,,\end{split} \label{Qdiff-eqns} \end{align} where $K_0=K(0)$. The second pair of equations in (\ref{Gstrand-eqns}) may then be written as \begin{align} \begin{split} \partial_t M_1 &= - \partial_s N_1 - (M_1M_2- N_1N_2)K'(X) \,,\\ \partial_t M_2 &= - \partial_s N_2 + (M_1M_2- N_1N_2)K'(X) \,,\\ \partial_t N_1 &= - \partial_s M_1 + (N_1M_2-M_1N_2) \frac{K_0-K}{K_0+K}K'(X) \,,\\ \partial_t N_2 &= - \partial_s M_2 + (N_1M_2-M_1N_2) \frac{K_0-K}{K_0+K} K'(X) \,. \end{split} \label{Gstrand-pp} \end{align} Asymptotically, when the peakons are far apart, the system (\ref{Gstrand-pp}) simplifies, since $\frac{K_0-K}{K_0+K}\to1$ and $K'(X)\to0$ as $|X|\to\infty$. The system (\ref{Gstrand-pp}) has two immediate conservation laws obtained from their sums and differences, \begin{align} \begin{split} \partial_t (M_1+M_2) &= -\, \partial_s (N_1+N_2) \,,\\ \partial_t (N_1-N_2) &=-\partial_s (M_1-M_2) \,.\end{split} \label{pp-CLs} \end{align} These may be resolved by setting \begin{align} \begin{split} M_1-M_2 &= \frac{\partial_t X}{K_0-K} \,,\qquad N_1-N_2 = - \frac{\partial_s X}{K_0-K} \,,\\ M_1+M_2 &= \partial_s\phi \,,\qquad N_1+N_2 = -\,\partial_t\phi \,, \end{split} \label{pp-Xpotentials} \end{align} and introducing two potential functions, $X$ and $\phi$, for which equality of cross derivatives will now produce the system of equations (\ref{Qdiff-eqns}) and (\ref{Gstrand-pp}). \section{A simplification.} A simplification arises if $\phi=0$, in which case the collision is perfectly antisymmetric, as seen from equation (\ref{pp-Xpotentials}). This is the peakon-antipeakon collision, for which the equation for $X$ reduces to \begin{align} (\partial_t^2 - \partial _s^2) X &+ \frac{K'}{2(K_0-K)} (X_t^2- X_s^2) = 0 \,. \end{align} This equation can be easily rearranged to produce a linear equation: \begin{align} (\partial_t^2 - \partial _s^2) F(X) = 0 \,,\quad\hbox{where}\quad F(X) = \int_{X_0}^X (K_0-K(Y))^{-1/2}\,dY \,. \end{align} When $K(Y)=\frac{1}{2}e^{-|Y|}$, we have \begin{align} F(X) = \sqrt{2}\int_{X_0}^X \frac{1}{\sqrt{1-e^{-|Y|}}}\,dY .\label{F-express} \end{align} We can take for simplicity $X_0=0$, this would change $F(X)$ only by a constant. The computation gives $$F(X)= 2\sqrt{2}\,\text{sign}(X)\cosh^{-1}\left(e^{|X|/2}\right) $$. Hence the solution $X(t,s)$ can be expressed in terms of any solution $h(t,s)$ of the linear wave equation $(\partial_t^2 - \partial _s^2)h(t,s)=0$ as \begin{align} X(t,s) = \pm \ln \left({\,\rm cosh}^2(h(t,s))\right) \,. \label{X-express} \end{align} $h(t,s)$ is any solution of the wave equation. $$M_1=-M_2 = \frac{\partial_t X}{2(K_0-K(X))} \,,\qquad N_1=-N_2 = - \frac{\partial_s X}{2(K_0-K(X))}.$$ \section*{Complex ${\rm Diff}(\mathbb{R})$-strand equations } The ${\rm Diff}(\mathbb{R})$-strands may also be \emph{complexified}. Upon complexifying $(s,t)\in \mathbb{R}^2\to(z,\bar{z})\in \mathbb{C}$ where $\bar{z}$ denotes the complex conjugate of $z$ and setting $\partial_z{g}= u \circ g$ and $\partial_{\bar{z}}{g}= \bar{u} \circ g$ the Euler-Poincar\'e $G$-strand equations in (\ref{Gstrand-eqn2R}) become \begin{align} \begin{split} \frac{\partial}{\partial z} \frac{\delta\ell}{\delta u} + \frac{\partial}{\partial \bar{z} } \frac{\delta\ell}{\delta { \bar{u}} } &= -\, {\rm ad}^*_{u}\frac{\delta\ell}{\delta u} - {\rm ad}^*_{{\bar{u}} }\frac{\delta\ell}{\delta { \bar{u}} } \,, \\ \frac{\partial {\bar{u}}}{\partial z} - \frac{\partial u}{\partial \bar{z}} &= {\rm ad}_u { \bar{u}} \,. \end{split} \label{ComplexGstrand-eqns1} \end{align} Here the Lagrangian $\ell$ is taken to be real: \begin{align} \ell (u, {\bar{u}}) = \frac12\|\nu\|_{H^1}^2 = \frac12\int u\, (1-\partial_x^2) \, {{ \bar{u}}} \,dx . \, \label{ComplexNorm-ell} \end{align} Upon setting $m={\delta\ell}/{\delta u}$, $\bar{m}={\delta\ell}/{\delta {\bar{u}} }$, for the real Lagrangian $\ell$, equations (\ref{ComplexGstrand-eqns1}) may be rewritten as \begin{align} \begin{split} m_{z} + \bar{m}_{\bar{z}} &= -\, {\rm ad}^*_{u }m - {\rm ad}^*_{{\bar{u}} }\,\bar{m} = - (u m)_x - mu_x - ({\bar{u}} \,\bar{m})_x - \bar{m}\,{\bar{u}}_x \,, \\ \\ {\bar{u}}_z- u _{\bar{z}} &= -\,{\rm ad}_{{\bar{u}}} \,u = -u \,{ \bar{u}}_x + { \bar{u}}\,u_x \,, \end{split} \label{ComplexGstrand-eqns2} \end{align}\bigskip where the independent coordinate $x\in\mathbb{R}$ is on the real line, although coordinates $ (z,\bar{z})\in\mathbb{C}$ are complex, as are solutions $u$, and $m=u-u_{xx}$. This is a possible comlexification of the Camassa-Holm equation. These equations are invariant under two involutions, $P$ and $C$, where \[ P: (x,m)\to(-x,-m) \quad \hbox{and} \quad C: \hbox{Complex conjugation.} \] They admit singular solutions just as before, modulo $\mathbb{R}\times\mathbb{R}\to\mathbb{C}$. For real variables $m=\bar{m}$, $u={\bar{u}}$ and real evolution parameter $z=\bar{z}=:t$, they reduce to the CH equation. Their travelling wave solutions and other possible CH complexifications are studied in \cite{Ho-Iv1}. \section*{Conclusions}\label{conclusion-sec} The $G$-strand equations comprise a system of PDEs obtained from the Euler-Poincar\'e (EP) variational equations for a $G$-invariant Lagrangian, coupled to an auxiliary \emph{zero-curvature} equation. Once the $G$-invariant Lagrangian has been specified, the system of $G$-strand equations in (\ref{MatrAlgEq}) follows automatically in the EP framework. For matrix Lie groups, some of the the $G$-strand systems are integrable. The single-peakon and the peakon-antipeakon solution of the ${\rm Diff}(\mathbb{R})$-strand equations (\ref{Gstrand-eqn2R}) depends on a single function of $s,t$. The \emph{complex} ${\rm Diff}(\mathbb{R})$-strand equations and their peakon collision solutions have also been solved by elementary means. The stability of the single-peakon solution under perturbations into the full solution space of equations (\ref{Gstrand-eqn2R}) would be an interesting problem for future work. \section{Acknowledgments} We are grateful for enlightening discussions of this material with F. Gay-Balmaz, T. S. Ratiu and C. Tronci. Work by RII was partially supported by the Science Foundation Ireland (SFI), under Grant No. 09/RFP/MTH2144. Work by DDH was partially supported by Advanced Grant 267382 FCCA from the European Research Council. \section*{References}
\section{Introduction}\label{sec:intro} \subsection{Context and issues} In a comprehensive review of recent publications on the Bayesian Analysis of Inverse Problems it is clear that there is a consistent growth of interest in the uncertainty quantification approach provided by the Bayesian paradigm. However, we also notice that some of the potential strength of the Bayesian approach is currently underexploited, namely, 1) prediction, 2) model selection and (let alone) 3) decision making under uncertainty. Recent reviews on the Bayesian Analysis of inverse problems \citep{MOHAMMAD-D2006, KAIPIO&FOX2011, WATZENIG&FOX2009, KAIPIO&FOX2011, WOODBURY2011} do not discuss these former points in detail. We have now access to important theoretical results on the definition of posterior distributions in infinite dimensions and regularity conditions for correct approximate schemes through numerical, finite dimension posteriors \citep[eg.][]{SCHWAB&STUART2012} that provide a sound theoretical background to the field. There are several applications in the (Bayesian dominated) field of image processing of various kinds \citep[][to mention some recent references]{ZHUetAl2011, CAIetAl2011, FALLetAl2011, CHAMAetAl2012, KOLEHMAINENetAl2007, NISSINENetAl2011, KOZAWAetAl2012}, and also we find a whole range of emerging application areas in the Bayesian Analysis of inverse problems \citep{CALVETTIetAl2006b, KEATSetAl2010, CUIetAl2011, WAN&ZABARAS2011, HAZELTON2010, KAIPIO&FOX2011}. However, only a handful of publications mentions or uses Bayesian predicting tools, that is, the posterior (predictive) distribution of yet to observe variables \citep{VEHTARI&LAMPINEN2000, SOMERSALOetAl2003, KAIPIO&FOX2011, CAPISTRANetAl2012}, and even less consider formally the model selection and model comparison tools developed in Bayesian statistics. We intend to contribute to the formal and more systematic use of the latter in the context of Inverse Problems. \medskip Predictive power is always a desirable property of mathematical models and inference, beyond parameter estimation, for model parameters that may or may not have straightforward physical meaning. We believe that comparing forward models as \textit{statistical} models is the way to proceed when predictive power is of main interest. The Bayesian model comparison and model averaging tools, in particular pairwise model comparison using Bayes Factors, is in such case the main tool to be used in this context, as far as predictive power is concerned \citep{HOETINGetAl1999}. In particular, this idea should be used when analyzing the numerical vs. the theoretical versions of the resulting posterior distribution. That is, the forward map, defined as the solution of a system of ODE's (or PDE's), represents a complex regressor that is only theoretically defined. The actual usage of the model necessarily involves a numerical solver that includes an approximation error depending on the solver step size $h$. Therefore, there is a theoretical posterior distribution $P_{\Theta | Y}( \theta | y)$ and the approximate $P_{\Theta | Y}^h( \theta | y)$ posterior. Recently a series of papers \citep[eg.][]{SCHWAB&STUART2012} discuss regularity conditions under which the latter tends to the former as the approximation error tends to zero, using a suitable metric. A metric comparison (ie. $|| P_{\Theta | Y}( \cdot | y) - P_{\Theta | Y}^h( \cdot | y) ||$) is useful to proving the required convergence theorems, but more practical considerations will be needed when evaluating the relative benefits of a numerical approach with a particular solver step size $h$ (for data $y$). \medskip We believe that both $P_{\Theta | Y}( \cdot | y)$ and $P_{\Theta | Y}^h( \cdot | y)$ (and $P_{\Theta | Y}^{h_i}( \cdot | y)$ for any other $h_i$) should be compared as \textit{models}, being $P_{\Theta | Y}( \cdot | y)$ the reference but computationally very expensive or even only theoretically available model and the approximate, for various solver precisions $h_1, h_2, \ldots$, as alternative, less computationally demanding computing models. Bayes factors may then be used to establish a sound comparison, to balance predictive power on the one hand vs. solver CPU time on the other, to establish a useful solver precision (that may even be far less precise than common usage but achieve comparable results relative to the observations errors, the model characteristics etc. for the problem at hand). One difficulty here is that in most real applications the theoretical model is unavailable. We therefore establish how to approximate the Bayes factors, without having the theoretical reference model, using solely the numerical solver approximation rates. \subsection{Notation} Assume that we observe a process $\mathbf{y} = (y_1,\dots,y_n)$ at the discrete times $t_1, \dots, t_n \in [0,T[ ^n$ such that \begin{equation}\label{eqn.obsmodel} y_i = f(X_{\theta}(t_i)) + \varepsilon_i ,~~ \varepsilon_i \sim_{i.i.d.} \mathcal{N}(0,\sigma^2) \quad \quad \quad \quad (\mathcal{M}) \end{equation} where $X_{\theta}$ is the solution of the following system of ordinary differential equations, namely the Forward Model, \begin{equation}\label{eqn.odemodel} \frac{dX_{\theta}}{dt} = F( X_{\theta}, t, \theta ); ~~ X_{\theta}(t_0) = X_0. \end{equation} $\theta \in \Theta \subset \mathbb{R}^d$ is a vector of unknown parameters. $F: \mathbb{R}^p \times [0, T[ \times \Theta \mapsto \mathbb{R}^p$ is a known function, whose regularity properties ensure the existence of a unique solution of the initial value problem (\ref{eqn.odemodel}) (the regularity conditions of $F$ will be discussed in Section~\ref{sec:solvers}). \medskip $f : \mathbb{R}^p \rightarrow \mathbb{R}^k$ in (\ref{eqn.obsmodel}) is the observation function. Many types of observation functions $f$ can be considered, modeling for instance the observation of a single component of the $p$-vector $X_{\theta}(t)$ or a (linear) combination of the components. In this paper, for the sake of simplicity, we consider a one dimensional observations problem only, that is $k=1$. Generalizations of our results to multivariate observations are possible and will be briefly mentioned in the Section~\ref{sec:discussion}. \medskip Any statistical decision from the data $\mathbf{y}$ --such as estimation, prediction or model selection-- relies on the likelihood function (in a Bayesian and other paradigms) \begin{equation}\label{eqn.lik} P_{ \mathbf{Y} | \phi} (\mathbf{y} | \theta,\sigma) = \sigma^{-n} (2\pi)^{-n/2} \exp\left\{-\frac{1}{2 \sigma^2} \sum_{i=1}^n (y_i -f(X_{\theta}(t_i)))^2 \right\} \end{equation} whose expression involves the computation of $X_{\theta}$, a solution of (\ref{eqn.odemodel}). However, except in very simple cases, an explicit expression of the solution is in general not available (although its existence is ensured by regularity assumptions on $F$). As a consequence, in practice, the system (\ref{eqn.odemodel}) is solved using a numerical solver and the inference is performed, not on the previous ``exact" model but on an approximate model, namely \begin{equation}\label{eqn.obsmodel.approx} y_i = f(X^h_{\theta}(t_i)) + \varepsilon_i ,~~ \varepsilon_i \sim_{i.i.d.} \mathcal{N}(0,\sigma^2)\quad \quad \quad \quad (\mathcal{M}_h) \end{equation}\label{eqn.lik.approx} where $X_{\theta}^h$ denotes the approximate solution of (\ref{eqn.odemodel}) supplied by the numerical solver ($h$ being a precision parameter of the solver). The new likelihood derived from model $\mathcal{M}_h$ is thus \begin{equation} P^h_{ \mathbf{Y} | \phi} (\mathbf{y} | \theta,\sigma) = \sigma^{-n} (2\pi)^{-n/2} \exp\left\{-\frac{1}{2 \sigma^2} \sum_{i=1}^n (y_i -f(X^h_{\theta}(t_i)))^2 \right\} . \end{equation} \medskip Changing from the original statistical model is obviously not without consequences. Since there is no alternative but to use the approximate model above (\ref{eqn.obsmodel.approx}), there exists a real need in understanding and controlling related consequences. In a Bayesian paradigm, any decision is based on the posterior distribution. A first natural choice is to compare the posterior distributions calculated from models $(\mathcal{M})$ and $(\mathcal{M}_h)$. Such a study has been proposed by, for example, \citet{DONNETetAL2007}. However, when comparing models in a Bayesian context, the natural tools are Bayes Factors. In this work, we recall their importance, propose an efficient way to compute them in this context and study some theoretical aspects of their calculation when the exact model is not available. \medskip The paper is organized as follows. In Section \ref{sec:solvers} we discuss the choice of a solver and its required properties from a Bayesian Inverse Problem point of view. Bayesian inference is presented in Section \ref{sec:posterior} and some control results on the posterior distributions are cited in Subsection \ref{subsec:control}. Bayes factors are introduced and our main theoretical results are developed in Section \ref{sec:bayes factor} . Our results are illustrated both with a simulation study and with real data in Section \ref{sec:numerical examples}. Finally, a discussion of the paper is presented in Section \ref{sec:discussion}. \section{ODE solvers from a Bayesian Inverse problem point of view}\label{sec:solvers} Bayesian analysis for inverse problems strongly relies on the numerical approximation of the underlying ODE system. One can choose to use a standard (more or less advanced) implemented solver as a black-box in a sense assuming that no approximation is made on the model. But in our approach, we aim at understanding the influence of this approximation. As a consequence, we are interested in the inherent properties of the numerical solver. When it comes to qualifing a numerical solver, three properties arise, namely its error (local or global), its stability and its stiffness. The three of them are addressed in the following section although global error is surely the most important one as far as the purposes of the paper are concern. \medskip Only a restricted number of nonlinear ODEs has a solution in closed form. Consequently, there is a plethora of numerical methods to solve the initial value problem~(\ref{eqn.odemodel}). Noteworthy are time-stepping methods based on Taylor approximation of the function $F$, multi-step methods and Runge-Kutta methods. These methods span through many orders of local accuracy. Beside the order of accuracy, standard requirements for a numerical method are consistency, convergence, and stability \citep{iserles1996first,quarteroni2007numerical}. However, even when these latter conditions hold, a common concern in the numerical solution of the initial value problem~(\ref{eqn.odemodel}) is error control. \medskip Two types of error can be considered, namely the local and the global errors. In the following, we use the theory developed in \citet{quarteroni2007numerical} to discuss the relationship between local and global errors for one-step Euler and fourth order explicit Runge-Kutta methods. \medskip Let $h$ be the step size of the method. We define a time grid as $ t_{n+1} = t_n + h$, for some fixed $h >0$. Let $X_{\theta,n}$ be the solver approximation of $X_{\theta} (t_n)$. One-step Euler and fourth order explicit Runge-Kutta methods have the form \begin{equation} \label{eq:euler} X_{\theta,n+1} = X_{\theta,n} + hK(t_{n},X_{\theta,n},h,F), \end{equation} where \begin{equation} K(t_{n},X_{\theta,n},h,F) = \left\{ \begin{array}{ll} K_{1} & \mbox{for Euler}\\ \frac{1}{6}(K_{1}+2K_{2}+2K_{3}+K_{4}) & \mbox{for 4th ord. Runge-Kutta,} \end{array}\right. \end{equation} with $$\begin{array}{cclccl} K_{1}&=&F( X_{\theta,n}, t_{n}, \theta) & K_{2}&=&F( X_{\theta} + \frac{1}{2}hK_{1}, t_{n}+\frac{1}{2}h, \theta)\\ K_{3}&=&F( X_{\theta} + \frac{1}{2}hK_{2}, t_{n}+\frac{1}{2}h, \theta)&K_{4}&=&F( X_{\theta} + hK_{1}, t_{n} + h, \theta) . \end{array} $$ (There is also a 2nd order Rungue-Kutta, which will be used in Section~\ref{sec:exa1}, but we avoid presenting its details.) \medskip Local truncation error $e_{n}$ is the error made in one step of the numerical method, that is $$ e_n = e_{h}(t_n,\theta)=||X_{\theta}(t_{n})- X_{\theta}(t_{n-1}) - hK(t_{n},X_{\theta}(t_{n-1}),h,F)||_{2} . $$ Global error $E_{n}$ is the difference between the computed solution and the true solution at any given value of $t$ belonging to the grid $$ E_n = E_{h}(t_n,\theta)=||X_{\theta}(t_{n})-X_{\theta,n}||_{2} . $$ For (order one) Euler and fourth order Runge-Kutta methods $e_{n}$ is $O(h^{2})$ and $O(h^{5})$ respectively (we avoid the details for order two Rungue-Kutta, for which indeed $e_n$ is $O(h^{3})$). Proposition~\ref{prop:error} establishes a relationship between global and local error for one-step numerical methods like Euler and Runge-Kutta. \begin{assum}\label{assum} We assume throughout the paper that the function $F$ in the right hand side of the initial value problem~(\ref{eqn.odemodel}) is continuously differentiable with respect to $X_{\theta}$ on a parallelepiped $\mathcal{R}= \{:t_{0}\leq t\leq t_{0}+a$, $||X_{\theta}-X_{0}||_{2}\leq b\}$ for each $\theta$. Hence a unique solution of the initial value problem exists in a neighborhood of $(t_{0},X_{0})$ for each $\theta$. $F$ as above implies $K$ is Lipschitz continuous on $X_{\theta}$, i.e. there is $L\in\mathbb{R}^{+*}$ such that if $(t,x)$, $(t,y)\in R$, then $$ ||K(t,x,h,F)-K(t,y,h,F)||_{2}\leq L||x-y||_{2} . $$ \end{assum} \begin{prop}\label{prop:error} Assume that \emph{Assumption~\ref{assum}} holds and for the solver at hand and the local truncation error $e_{n}$ is $O(h^{p+1})$. Then the global truncation error is $O(h^p)$. \end{prop} \begin{proof} We have $$ E_{n} = ||X_{\theta}(t_{n}) - X_{\theta,n}||_{2} = ||X_{\theta}(t_{n}) - X_{\theta,n-1} - h K(t_{n},X_{\theta,n-1},h,F)||_{2} . $$ Adding and susbtracting the term $X_{\theta}(t_{n-1}) + h K( t_{n}, X_{\theta}(t_{n-1}), h, F)$ we obtain \begin{align*} E_{n} &\leq e_{n} + || X_{\theta}(t_{n-1}) - X_{\theta,n-1} ||_{2} + h || K( t_{n}, X_{\theta}(t_{n-1}), h, F) - K(t_{n},X_{\theta,n-1},h,F) ||_{2} \\ &\leq e_{n} + E_{n-1} + h L || X_{\theta}(t_{n-1}) - X_{\theta,n-1} ||_{2} \\ &\leq e_{n} + E_{n-1}(1 + hL). \end{align*} Consequently \begin{align*} E_{n} &\leq e_{n} + e_{n-1}(1 + hL) + \ldots + e_{1}(1 + hL)^{n-1}\\ &\leq H h^{p+1} \frac{1-(1 + hL)^n}{1-(1 + hL)} = \frac{H}{L} h^p \{ ( 1 + hL)^n - 1 \} . \end{align*} Therefore, since $h = \frac{l}{n}$ ($l=t_n-t_1$), $$ E_{n} \leq \frac{H}{L} h^p \left\{ \left( 1 + \frac{l L}{n} \right)^n - 1 \right\} \leq \frac{H}{L} e^{l L} h^p. $$ \end{proof} From Proposition \ref{prop:error}, we deduce that, for any explicit one-step method of order $p$ such as Euler ($p=1$) and Runge-Kutta ($p=2$ or $p=4$) schemes, the global error is of order $O(h^p)$ for $h$ small enough. If we set $X_{\theta}^h(t_n) = X_{\theta,n}$, note that we have proved that $$ \max_{t \in \{t_1, t_2, \ldots , t_n\}} || X_{\theta}(t) - X_{\theta}^h(t) || \leq C_{\theta} h^p $$ since $H$ and $L$ above depend on $\theta$. This global error order control will be needed in Sections~\ref{subsec:control} and~\ref{sec:comparing} to prove our main result. \medskip In the numerical community, the control of the error means keeping the error $e_n$ or $E_n$ under a fixed level (beyond the above mentioned asimptotic results), along the application of the scheme. The local error $e_n$ can be controlled directly, using for instance, the ``Milne device'' \citep{iserles1996first}. However, there are not known general methods to control global error $E_n$, although some methods exist to estimate it, for instance, those relying on adjoint state analysis, see~\cite{cao2004posteriori,lang2007global}. In the results that follow we do not require an estimation of this global (or local) error, solely the knowledge of the global error order for the solver at hand. \medskip Another important issue in the numerical solution of problem~(\ref{eqn.odemodel}) is stiffness. According to~\cite{lambert1991numerical}, if a numerical method with finite region of absolute stability applied to system~(\ref{eqn.odemodel}) is forced to use in a certain interval in $t$ a steplength excessively small in relation to the smoothness of the exact solution in that interval, then system~(\ref{eqn.odemodel}) is said to be stiff in that interval. We remark that many systems of ODE modeling real life phenomena are stiff, see~\cite{gutenkunst2007universally}. \medskip Software implementing time-stepping methods mentioned above must address error control, stability and stiffness issues. Actually, local error control mechanisms can cope with stiffness at the expense of taking very small step-sizes. Therefore, advanced features like variable order methods, and variable stepsize methods have been developed and implemented in libraries of common high level programing languages like R, Matlab and Python-Scipy. \medskip The results shown in this paper assume a fixed step method. We only work with the Euler and Rugue-Kutta methods (orders $p=1,2,4$ respectively). \section{Bayesian inference for inverse problems: practical and theoretical aspects}\label{sec:posterior} There are some excellent reviews concerning the Bayesian analysis of Inverse Problems \citep{ KAIPIO&SOMERSALO2005, FOXetAl1999} and for a more detailed description these or other sources should be consulted. In this section we only present the particular aspects of the field relevant to the development of our results. \medskip Any Bayesian statistical decision (such as estimation) is based on the posterior distribution, given by the Bayes formula \begin{equation}\label{eqn.post} P^h_{ \Phi | \mathbf{Y} }( \theta,\sigma | \mathbf{y} ) = \frac{P^h_{ \mathbf{Y} | \Phi }( \mathbf{y} | \theta,\sigma) P_{\Phi}(\theta,\sigma)}{P^h_{\mathbf{Y}} (\mathbf{y})} \end{equation} where $P_{\Phi}(\theta,\sigma)$ is the prior distribution on $(\theta,\sigma)$ and $$ P^h_{\mathbf{Y}} (\mathbf{y}) = \int P^h_{ \mathbf{Y} | \Phi }( \mathbf{y} | \theta,\sigma) P_{\Phi}(\theta,\sigma) d\theta d\sigma $$ is the normalization constant, also called the \emph{marginal likelihood} of data $\mathbf{y}$. \medskip An estimator of $(\theta,\sigma)$ is derived from some characteristics of the posterior distribution. Various approaches can be adopted. Either the estimator is taken as the mode of the posterior distribution, resulting into the MAP estimator $(\hat{\theta}^{MAP},\hat{\sigma}^{MAP}) = \arg \max _{(\theta,\sigma) }P^h_{ \Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} )$ or the estimator derives from the minimization of a loss function $L(\theta,\sigma| \eta)$, that is $ (\hat{\theta},\hat{\sigma}) = \arg \min _{(\theta,\sigma)} \int L(\theta,\sigma| \eta)P^h_{ \Phi | \mathbf{Y} }(\theta,\sigma | \mathbf{y} ) d\theta d\sigma $. If the loss function $ L(\theta,\sigma| \eta)$ is equal to the $L^2$-distance, then the corresponding estimator is the posterior mean $ (\hat{\theta}^{L^2},\hat{\sigma}^{L^2}) = \int (\theta,\sigma ) P^h_{\Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} ) d\theta d\sigma $. If the loss function $ L(\theta| \eta)$ is equal to the $L^1$-distance, then the corresponding estimator is the posterior median, that is $ (\hat{\theta}^{L^1},\hat{\sigma}^{L^1}) = (F^h)^{-1}_{ \Phi| \mathbf{Y} }\left( \frac{1}{2}\right)$, etc. In general, posterior expectations of all types (ie. $\int g(\phi) P^h_{\Phi| \mathbf{Y} }( \phi | \mathbf{y} ) d\phi$, for some measurable function $g$) are used to explore the posterior distribution, and these include all marginal distributions for individual parameters or posterior probabilities of specific regions of interest. \medskip Besides some few simple conjugate models, the normalizing constant $P^h_{\mathbf{Y}} (\mathbf{y})$ has no explicit analytic expression and therefore the above estimators can not be directly calculated. If the dimension of $\Phi$ is 1 or 2 we could rely on numerical integration to obtain the normalizing constant $P_{\mathbf{Y}}^h (\mathbf{y})$. In larger dimensions, the standard solution is to resort to Monte Carlo methods to approximate the estimators. Let $(\theta^{(l)}, \sigma^{(l)})_{i=1\dots L}$ be a sample from the posterior distribution, the $L^2$ estimator for instance, is approximated as $(\hat{\theta}^{L^2},\hat{\sigma}^{L^2}) = \left( \frac{1}{L} \sum_{l=1}^M \theta^{(l)}, \frac{1}{L} \sum_{l=1}^L \sigma^{(l)}\right)$. \medskip Simulation from the posterior distribution is not a direct task and Markov Chain Monte Carlo algorithms \citep[see][for a didactic review]{Robert&Casella2004} are standard tools to sample from the posterior distribution $P^h_{ \Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} )$. However, such algorithms have to be carefully designed when the evaluation of the regression function of the model (and consequently of the likelihood function) is computationally intensive. In the next section, we recall the basics of MCMC algorithms and discuss their application to inverse problems and the calculation of Bayes' Factors. \subsection{Sampling from the posterior distribution in Inverse problems}\label{sec:MCMC} Markov Chain Monte Carlo (MCMC) is specially suited for sampling from complex multidimensional distributions and is ubiquitous in modern Bayesian analyses. We do not intend to present a comprehensive review of MCMC here but merely state the basic principles to consider some aspects of implementing MCMC in the Inverse Problem context; otherwise the reader is referred to \citet{Robert&Casella2004}. The principle of MCMC algorithms is to generate a Markov chain whose invariant distribution is the distribution of interest $P_{ \Phi | \mathbf{Y} }^h( \theta,\sigma | \mathbf{y} )$. Many versions have been proposed in the literature. Among those, the Gibbs algorithm and the Metropolis-Hasting algorithms are the most used and quickly described below. \medskip The Gibbs sampler is a very popular MCMC algorithm. However, only in canonical cases (eg. conditional conjugacy) it make sense to be used. In our Inverse Problem settings, this is not the case and the general Metropolis-Hastings MCMC algorithm needs to be used instead. \medskip The Metropolis-Hastings (MH) algorithm starts by defining a proposal (or instrumental) conditional distribution $q( \phi' | \phi )$, which we are able to simulate from for any $\phi$ in the support of the posterior distribution. At iteration $(\ell)$: \begin{enumerate} \item a proposed value $\phi'$ is simulated given the current state of the Markov Chain $\phi^{(\ell)}$ from the proposal $q( \cdot | \phi^{(\ell)} )$; \item the proposed point $\phi'$ is accepted as the new point in the Markov chain $\phi^{(\ell+1)}$ with probability $ \rho( \phi^{(\ell)}, \phi' )$: $$\ phi^{(\ell +1)} = \left\{ \begin{array}{ccl} \phi' & \mbox{with probability} & \rho( \phi^{(\ell)}, \phi' )\\ \phi'^{(\ell)} & \mbox{with probability} & 1-\rho( \phi^{(\ell)}, \phi' ) \end{array} \right. $$ where the acceptance probability $ \rho( \phi^{(\ell)}, \phi' )$ is equal to \begin{eqnarray*} \rho( \phi^{(\ell)}, \phi' ) &=& \min\left\{1, \frac{P_{ \Phi | \mathbf{Y} }^h( \phi' | \mathbf{y} )}{P_{ \Phi | \mathbf{Y} }^h( \phi^{(\ell)} | \mathbf{y} )} \frac{q( \phi^{(\ell)} | \phi' )}{q( \phi' | \phi^{(\ell)} )}\right\}\\ & =& \min\left\{1, \frac{P_{\mathbf{Y} |\Phi}^h( \mathbf{y} | \phi' )P_{\Phi}(\phi')}{P_{\mathbf{Y} |\Phi}^h( \mathbf{y} | \phi^{(\ell)} )P_{\Phi}( \phi^{(\ell)})} \frac{q( \phi^{(\ell)} | \phi' )}{q( \phi' | \phi^{(\ell)} )}\right\} . \end{eqnarray*} \end{enumerate} Commonly a series of proposal distributions $q_1, q_2, \ldots , q_m$ are entertained, leading to a series of $m$ transition kernels that are systematically or randomly scanned (the latter leads to the desirable reversibility property) to form an easily provable convergent chain to the posterior $P_{ \Phi | \mathbf{Y} }^h( \cdot | \mathbf{y} )$. \medskip In any case, at each iteration of the MH MCMC we need to evaluate the (unnormalized) posterior (It is therefore crucial to minimize the number of iterations in the MCMC and the number of likelihood evaluations). \medskip Optimizing MCMC algorithms (that is to say minimizing the number of iterations) has been a very active research topic in the last decade. There are adaptive algorithms \citep{HaarioEtAl1998, Atchade&Rosenthal2005} that try to learn from previous steps of the chain to adapt the proposals $q_1, \ldots , q_m$. These methods require additional regularity conditions on the adaptive scheme, model and prior that might limit their applicability. \citet{Christen&Fox2010} also propose the t-walk, which self adjusts keeping two points in the parameter space, and that commonly samples with reasonable efficiency. However, robust, multipurpose, automatic and optimal methods are still far away in the MCMC horizon (to make an optimistic metaphor). \medskip When it comes out to inverse problem specifically, various savings can be considered. \medskip $\bullet$ First, a naive but straightforward computational economy is to save $U^{(\ell)} = -\log P_{ \mathbf{Y} | \Phi }^h( \mathbf{y} | \phi^{(\ell)} ) - \log P_{\Phi}( \phi^{(\ell)} )$. Indeed, this quantity will be used for any new simulated value $\phi'$ until the chain reaches a new point $\phi^{(\ell+1)}$. We will see in Section~\ref{subset:comp Bayes factor} that these quantities can also be recycled for model comparison purposes. Note that these $U^{(\ell)}$ are very useful for convergence analysis. \medskip $\bullet$ Second, in this Inverse Problem setting we may exploit the possibility that a coarser numerical solver with a higher error rate (or equivalently a larger step size $h_1$) might lead to a far less CPU demanding approximate calculation of the likelihood $P_{ \mathbf{Y} | \phi }^{h_1}( \mathbf{y} | \phi^{(i)} )$. More precisely, \cite{Christen&Fox2005} suggest to propose a candidate $\phi'$ and test its promising acceptance probability using an approximate and cheap calculation of the likelihood (using the coarser numerical solver). If this $\phi'$ has a ``good" probability to being accepted, the full blow and expensive calculation of the likelihood is used to firmly accept or reject $\phi'$. This two-step or ``delay acceptance MH'' algorithm may save substantial CPU time during the MCMC \citep[see][]{Christen&Fox2005}. \subsection{Theoretical results on the error control of the approximate posterior distribution}\label{subsec:control} After having examined how the standard algorithmic tools for Bayesian inference can be adapted or designed specifically to the Inverse problem context, we may want to understand and control the consequences implied by the use of an approximate model from a theoretical point of view. Such a result can be found in \citet{DONNETetAL2007} who compare the posterior distributions of the exact and approximate models respectively --namely $P_{ \Phi | \mathbf{Y} }$ and $P^h_{ \Phi | \mathbf{Y} }$-- through the total variation distance. \begin{prop} Assume that $\phi = (\theta, \sigma)$ remains in a compact set and that the numerical scheme of step size $h$ is such that $\{t_1,\dots,t_n\}\subset h \mathbb{N}$ and \begin{equation}\label{eqn:globerr} \max_{t \in \{t_1,\dots,t_n\}} \|X_{\theta}(t)-X^h_{\theta}(t)\|_{\mathbb{R}^p} \leq C_{\theta} h^p. \end{equation} Also assume that the observation function $f$ is differentiable with a bounded derivative. Then there exists a constant $C_{\mathbf{y}}$ such that for every $(\theta, \sigma)$ and $h$ small enough \begin{equation}\label{eq:ineq1} |P_{ \Phi | \mathbf{Y} }(\theta,\sigma;\mathbf{y})- P^h_{ \Phi | \mathbf{Y} }(\theta,\sigma;\mathbf{y})| \leq C_{\mathbf{y}} \pi(\theta,\sigma) h^p . \end{equation} As a consequence, \begin{equation}\label{eq:ineq2} D_{T.V}(P_{ \Phi | \mathbf{Y} }, P^h_{ \Phi | \mathbf{Y} }) \leq C_{\mathbf{y}} h^p \end{equation} where $D_{T.V}$ is the total variation distance. Moreover, \begin{equation}\label{eq:ineq3} \| (\hat{\theta}^{L^2},\hat{\sigma}^{L^2}) - (\hat{\theta}^{h,L^2},\hat{\sigma}^{h,L^2}) \| \leq h^p C'_{\mathbf{y}} \end{equation} \end{prop} \begin{proof} \cite{DONNETetAL2007}'s results are developed for mixed effects models in a maximum likelihood context but can be adapted to models~(\ref{eqn.obsmodel}) and ~(\ref{eqn.obsmodel.approx}). Inequality (\ref{eq:ineq1}) is derived from Theorem 4 of \cite{DONNETetAL2007}. The control on the total variation distance is derived directly. The inequality (\ref{eq:ineq3}) is obtained as follows $$ \begin{array}{rcl} && \| (\hat{\theta}^{L^2},\hat{\sigma}^{L^2}) - (\hat{\theta}^{h,L^2},\hat{\sigma}^{h,L^2}) \| \\ &&= \|\int (\theta,\sigma ) P_{\Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} ) d\theta - \int (\theta,\sigma ) P^h_{\Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} ) d\theta d\sigma\| \\ &&= \|\int (\theta,\sigma ) \left[P_{\Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} ) - P^h_{\Phi| \mathbf{Y} }( \theta,\sigma | \mathbf{y} ) \right]d\theta d\sigma\|\\ &&\leq \int \|(\theta;\sigma)\| C_{\mathbf{y}} P_{\Phi}(\theta,\sigma) h^p d\theta d\sigma\\ &&\leq h^p C_{\mathbf{y}} \int \|(\theta;\sigma)\| P_{\Phi}(\theta,\sigma) d\theta d\sigma \end{array} $$ \end{proof} Note that obtaining the same type of control on the MAP estimator requires additional regularity assumptions on the posterior distributions, such as unicity of the extremum and properties on the second derivatives. This leads to encourage the use of $L^2$ estimates over the MAP estimates in this context. \medskip \cite{COTTERetAl2010} also proposes control results of the same type. Even if these results provide interesting theoretical error controls, they rely on unknown constants and so can not be used as such in practice. In the next section, we adopt a Bayes factor point of view and highlight that such an approach leads to results of more practical interest. \section{Model selection and Bayes Factor for inverse problems}\label{sec:bayes factor} In Bayesian inference, model selection is performed using the Bayes factors whose principle is recalled here in a general context. Let $\mathbf{y}$ be the observations and let $\mathcal{M}_1$ and $\mathcal{M}_2$ be two models in competition. Each model $\mathcal{M}_i$ is defined through a likelihood depending on a set of parameters and a prior distribution on the parameters. More precisely, $$ (\mathcal{M}_1) \left\{ \begin{array}{ccl} \mathbf{y} &\sim& P^1_{\mathbf{Y} | \Phi_1}(\mathbf{y}|\phi_1)\\ \phi_1 & \sim & P_{\Phi_1}^1(\phi_1) \end{array}\right. \quad \quad (\mathcal{M}_2) \left\{ \begin{array}{ccl} \mathbf{y} &\sim& P^2_{\mathbf{Y} | \Phi_2}(\mathbf{y}|\phi_2)\\ \phi_2 & \sim & P_{\Phi_2}^2(\phi_2) \end{array}\right. . $$ Consider a prior distribution on the set of the models $\{\mathcal{M}_1, \mathcal{M}_2\}$, the decision between the competing models $\mathcal{M}_1$ and $\mathcal{M}_2$ is based on the ratio of their respective posterior probabilities $$ \frac{P( \mathcal{M}_2 | \mathbf{y} )}{P( \mathcal{M}_1 | \mathbf{y} )} = \frac{P_{\mathbf{Y}}^2(\mathbf{y})}{P_{\mathbf{Y}}^1(\mathbf{y})} \frac{ P(\mathcal{M}_2)}{ P(\mathcal{M}_1)} $$ where $P_{\mathbf{Y}}^i(\mathbf{y})$ is the `integrated likelihood' or the marginal distribution of $\mathbf{Y}$ of model $\mathcal{M}_i$, namely $$ P_{\mathbf{Y}}^i(\mathbf{y}) = \int P^i_{\mathbf{Y} | \Phi_i}(\mathbf{y}|\phi_i) P_{\Phi_i}(\phi_i)d\phi_i . $$ Finally, the comparison of models relies on the computation of the marginal likelihoods which has been the object of a rich literature. Two approaches may be cited. One consist in running a specific MCMC to approximate the quantity of interest \citep[see][for a review]{HanetCarlin2001}. The second relies on Monte Carlo approximations of the marginal likelihood $P_{\mathbf{Y}}^i(\mathbf{y})$ \citep[see for instance][]{ChenAndShao1997}. \subsection{Computation of Bayes Factors in an Inverse Problem context}\label{subset:comp Bayes factor} In our inverse problem context --where each iteration of the MCMC requires the computationally intensive approximation of an ODE -- we would like to avoid increasing the computational burden by using a specific MCMC and would prefer recycling the outputs of the MCMC algorithm into a Monte Carlo strategy. An answer can be found in the Gelfand and Dey's estimator \citep{GelfandetDey1994}. \medskip More precisely, a standard solution to compute the marginal likelihood is to propose an estimation based on a Monte Carlo approximation. Let $\{\phi_i^{(l)}\}_{l=1\dots L}$ be an i.i.d. sample from a proposal distribution $\pi_{IS}$ then the following estimator $$ \widehat{p}_i = \frac{1}{L} \sum_{l=1}^L P^i_{\mathbf{Y} | \Phi^{(l)}_i}(\mathbf{y}|\phi^{(l)}_i) \frac{P_{\Phi_i}(\phi^{(l)}_i)}{\pi_{IS}(\phi^{(l)}_i)} $$ supplies a convergent and non biased estimator of $P_{\mathbf{Y}}^i(\mathbf{y}) $. However, in the inverse problems context (when one or both models $\mathcal{M}_i$ are defined through ODE without explicit solution) such a strategy requires the evaluation of the likelihood and thus the evaluation of an approximate solution of the dynamical system for each newly generated value of parameters $\phi^{(l)}_i$, which can be burdensome from a computational time point of view. \medskip The Gelfand and Dey's estimator is an alternative solution to this situation. Assume that (as it is in standard situations), $P_{ \Phi | \mathbf{Y} }^h( \theta, \sigma | \mathbf{y} )$ is sampled using an intensive Monte Carlo procedure, typically a Metropolis-Hasting MCMC. (Note that for ease of presentation we avoid the exponent $i$ or $h$ and, notationally, consider the posterior distribution $P_{ \Phi | \mathbf{Y} }( \theta, \sigma | \mathbf{y} )$ and estimation of the normalization constant $P_\mathbf{Y}(\mathbf{y})$) \medskip Assume that the prior distribution $P_\Phi$ is absolutely continuos w.r.t the Lebesgue measure, and thus $P_\Phi( \theta, \sigma )$ and $P_{\Phi | \mathbf{Y}}( \theta, \sigma | \mathbf{y})$ are densities in the usual sense. The Gelfand and Dey's estimator relies on the obvious following expression $$ \left[ P_\mathbf{Y}(\mathbf{y}) \right]^{-1} = \int \frac{\alpha( \theta, \sigma)}{P_{\mathbf{Y} | \Phi}( \mathbf{y} | \theta, \sigma) P_{\Phi}( \theta, \sigma)} P_{ \Phi | \mathbf{Y} }( \theta, \sigma | \mathbf{y} ) d\theta d\sigma , $$ where $\alpha$ is any density ($\int \alpha( \theta, \sigma) d\theta d\sigma = 1$) with support containing the support of the posterior. Now, let $\theta^{(1)}, \sigma^{(1)}, \theta^{(2)}, \sigma^{(2)}, \ldots ,\theta^{(L)}, \sigma^{(L)}$ be a MCMC sample of the posterior $P_{ \Phi | \mathbf{Y} }( \theta, \sigma | \mathbf{y} )$. Considering the above expression, the desired marginal may be approximated by \begin{equation} \label{eqn.estMarg} \widehat{P}_\mathbf{Y}(\mathbf{y}) = \left[ \frac{1}{L} \sum_{l=1}^L \frac{\alpha( \theta^{(l)}, \sigma^{(l)} )}{P_{\mathbf{Y} | \Phi}( \mathbf{y} | \theta^{(l)}, \sigma^{(l)} ) P_{\Phi}( \theta^{(l)}, \sigma^{(l)} )} \right]^{-1} . \end{equation} The choice of $\alpha$ conditions the quality of the estimator (its variance). If $\alpha(\theta,\sigma) = \pi(\theta,\sigma)$, the estimator $\widehat{P}_\mathbf{Y}(\mathbf{y})$ is the harmonic mean which is known to have a dramatic unstable behaviour (infinite variance) in some cases. Best strategies are those that use a weighting $\alpha$ density that stabilizes this estimators, for instance using somehow an $\alpha$ that resembles $P_{\mathbf{Y} | \Phi}( \mathbf{y} | \theta, \sigma) P_{\Phi}( \theta, \sigma)$. A simple calculation leads to the fact that using a \textit{thinner} tailed $\alpha$ (as oppose to the result in importance sampling) is better suited to obtaing a finite variance for the estimator above. \medskip Moreover, in an inverse problem context, it is critical to avoid recalculating the likelihood $P_{ \mathbf{Y} | \Phi }( \mathbf{y} | \theta, \sigma) $ since it involves numerically solving the ODE system in (\ref{eqn.odemodel}). As a consequence we propose to proceed as follows. \medskip $\bullet$ At each iteration of a typical MH MCMC, the computation of the probability of acceptation requires to evaluate $P_{ \Phi | \mathbf{Y} }( \mathbf{y} | \theta^{(l)}, \sigma^{(l)} ) P_{\Phi}( \theta^{(l)}, \sigma^{(l)} )$. After the burn in period, we save these values, letting $U_l = -\log P_{ \Phi | \mathbf{Y} }( \mathbf{y} | \theta^{(l)}, \sigma^{(l)} ) - \log P_{\Phi}( \theta^{(l)}, \sigma^{(l)} )$. \medskip $\bullet$ A small subsample of $ \theta^{(l)}, \sigma^{(l)} $, typically of size less than 1,000, is then used to create a Kernel Density Estimate (KDE), which we will use as our weighting density $\alpha$. This KDE is (typically) a mixture of Gaussians, with support in the whole space, and will roughly resemble the posterior $P_{\Phi | \mathbf{Y}}$. \medskip $\bullet$ Let $A_i = - \log \alpha( \theta^{(l)}, \sigma^{(l)} )$, then our estimate becomes $$ P_\mathbf{Y}(\mathbf{y}) \approx \left[ \frac{1}{L} \sum_{l=1}^L \exp( U_l - A_l ) \right]^{-1} . $$ This procedure is fast and basically is a byproduct of the MCMC sample, with little CPU burden added. There are robust and fast KDE's routines available in popular programming languages like R and Python-Scipy. \begin{Rem} Note that $P_{\mathbf{Y} | \Phi}( \mathbf{y} | \theta, \sigma) P_{\Phi}( \theta, \sigma)$ needs to be known exactly, and be coded accordingly which is not the case in some situations, for example, when the prior is not normalized and only implicitly defined etc. \end{Rem} \subsection{Comparing the exact and approximate models through Bayes factors}\label{sec:comparing} In section \ref{subsec:control}, we compared the true and approximate models (models \ref{eqn.obsmodel} and \ref{eqn.obsmodel.approx}) through a derived quantity, that is to say their corresponding posterior distribution or the estimators they supplied. However, a more natural way to compare models is to use the Bayes factors. Let $(\mathcal{M})$ and $(\mathcal{M}_h)$ be the two following models $$ (\mathcal{M}) \left\{ \begin{array}{ccl} y_i &=& f(X_{\theta}(t_i)) + \varepsilon_i \\ \varepsilon_i &\sim&_{i.i.d.} \mathcal{N}(0,\sigma^2)\\ (\theta, \sigma) &\sim& P_{\Phi}(\theta,\sigma) \end{array}\right. \quad \quad (\mathcal{M}_h) \left\{ \begin{array}{ccl} y_i &=& f(X^h_{\theta}(t_i)) + \varepsilon_i \\ \varepsilon_i &\sim&_{i.i.d.} \mathcal{N}(0,\sigma^2)\\ (\theta, \sigma) &\sim& P_{\Phi}(\theta,\sigma) . \end{array}\right. $$ Comparing $\mathcal{M}$ and $\mathcal{M}_h$ through a Bayes factor requires to compute the following quantity $$ B_{\mathcal{M},\mathcal{M}_h} = \frac{P( \mathcal{M} | \mathbf{y} )}{P( \mathcal{M}_h | \mathbf{y} )} = \frac{\int P_{\mathbf{Y}|\Phi}(\mathbf{y} | \theta,\sigma) P_{\Phi}(\theta,\sigma)d\theta d\sigma}{\int P^h_{\mathbf{Y}|\Phi}(\mathbf{y} | \theta,\sigma) P_{\Phi}(\theta,\sigma)d\theta d\sigma}\frac{P(\mathcal{M}) }{ P(\mathcal{M}_h)} . $$ However, $\int P_{\mathbf{Y}|\Phi}(\mathbf{y} | \theta,\sigma) P_{\Phi}(\theta,\sigma)d\theta d\sigma$ is not known in general, since it involves the theoretical model. In order to understand the behavior of the Bayes Factor, we study $B_{\mathcal{M},\mathcal{M}_h}$ for small $h$ and obtain the following result. \begin{prop}\label{prop:2} Assume that the numerical solver is such that the global error truncation can be written as $$ E_h(t,\theta) = X_{\theta}^h(t) -X_{\theta}(t) = O(h^p) , $$ where $h$ is the stepsize of the method. In addition, assume that the observation function $f$ is differentiable on $\{X_{\theta}(t), \theta \in \Theta, t\in [0,T]\}$. Then, there exists a constant $B(\mathbf{y}) \in \mathbb{R}$ (which does not depend on $h$) such that $$ \frac{P_\mathbf{Y}(\mathbf{y})}{P^h_\mathbf{Y}(\mathbf{y})} \simeq 1+ B(\mathbf{y}) h^p . $$ \end{prop} \begin{proof} Using the asymptotic behavior of the global error truncation and assuming that $f$ is differentiable, we can write \begin{equation}\label{eqn:Dh} D_h(t,\theta) = f(X_{\theta}^h(t)) - f(X_{\theta}(t)) = \nabla f \left( X_{\theta}(t)\right)(X_{\theta}^h(t) -X_{\theta}(t)) + O(h^p) \end{equation} This approximation allows us to obtain a development of the marginal likelihood. Let $R_h(\phi) = \frac{P^h_{\mathbf{Y} | \Phi}(\mathbf{y}|\theta,\sigma)}{P_{\mathbf{Y} | \Phi}(\mathbf{y}|\theta,\sigma)}$, then \begin{eqnarray*} P^h_\mathbf{Y}(\mathbf{y}) &=& \int P_{\mathbf{Y} | \Phi}^h( \mathbf{y} | \phi) P_{\Phi}(\phi) d\phi = \int P_{\mathbf{Y} | \Phi}(\mathbf{y}| \phi)R_h(\phi) P_{\Phi}(\phi) d\phi \\ &=& P_\mathbf{Y}(\mathbf{y}) + \int P_{\mathbf{Y} | \Phi}(\mathbf{y}| \phi)(R_h(\phi)-1) P_{\Phi}(\phi) d\phi . \end{eqnarray*} We see that \begin{eqnarray} \label{eqn.rtheta} R_h(\phi) -1 &=& \exp \left\{ -\frac{1}{2\sigma^2}\sum_{i=1}^n \left[f(X^h_{\theta}(t_i))-f(X_{\theta}(t_i))\right]^2 + \right.\nonumber \\ && \left. 2\left[y_i-f(X_{\theta}(t_i))\right]\left[f(X_{\theta}(t_i))-f(X^h_{\theta}(t_i))\right] \right\} -1 \nonumber \\ &=& - \frac{1}{2\sigma^2}\sum_{i=1}^n D_h(t_i,\theta) ^2 + 2 (y_i-f(X_{\theta}(t_i)))D_h(t_i,\theta) \nonumber \\ && + O(D_h(t_i,\theta)^2) , \end{eqnarray} since $e^x -1 = x + O(x^2)$ for $x$ small enough. Using the expression in (\ref{eqn:Dh}) for $D_h$ and the solver global error control assumption in (\ref{eqn:globerr}) we must have that $R_h(\phi)-1 = O(h^p)$. From this we get $$ P^h_\mathbf{Y}(\mathbf{y}) = P_\mathbf{Y}(\mathbf{y}) + O(h^p) $$ implying that for $h$ small enough, $$ \frac{P_\mathbf{Y}(\mathbf{y})}{P^h_\mathbf{Y}(\mathbf{y})} \simeq 1+ B(\mathbf{y})h^p . $$ \end{proof} \medskip \begin{corollaire}\label{prop:2} If $\hat{g} = \int g(\phi) P_{ \Phi | \mathbf{Y} }( \phi | \mathbf{y} ) d\phi$ and $\hat{g}^h = \int g(\phi) P^h_{ \Phi | \mathbf{Y} }( \phi | \mathbf{y} ) d\phi$ exists, then $| \hat{g}^h - \hat{g} | = \frac{P_\mathbf{Y}(\mathbf{y})}{P^h_\mathbf{Y}(\mathbf{y})} B_g(\mathbf{y})h^p = O(h^p)$. \end{corollaire} \begin{proof} Note that $| \hat{g}^h - \hat{g} | = \left| \int g(\phi) R_h(\phi) \frac{P_\mathbf{Y}(\mathbf{y})}{P^h_\mathbf{Y}(\mathbf{y})} P_{ \Phi | \mathbf{Y} }( \phi | \mathbf{y} ) d\phi - \hat{g} \right|$ and therefore $| \hat{g}^h - \hat{g} | = \frac{P_\mathbf{Y}(\mathbf{y})}{P^h_\mathbf{Y}(\mathbf{y})} \left| \int g(\phi) (R_h(\phi) -1) P_{ \Phi | \mathbf{Y} }( \phi | \mathbf{y} ) d\phi - \left( \frac{P^h_\mathbf{Y}(\mathbf{y})}{P_\mathbf{Y}(\mathbf{y})} - 1 \right) \hat{g} \right|$. Combining (\ref{eqn.rtheta}) and the above theorem one reaches the result. \end{proof} Regarding the above result, we make the following comments and remarks. \medskip $\bullet$ From (\ref{eqn.rtheta}), we note that the error in the regression term $D_h(t,\theta)$ is not important per se but with respect to the observation noise standard error $\sigma$. This obvious remark has consequences. It means that, when working on a statistical model involving the numerical approximation of a differential system, there is no need in choosing a step size the smaller as possible but adapting it such that the global error is small with respect to $\sigma$. This can allow computational time savings as illustrated on the following numerical examples. \medskip $\bullet$ Note that $B( \mathbf{y})$ only depends on the numerical method and on the data, but not on the step size $h$. \medskip $\bullet$ The constant $B( \mathbf{y})$ can be estimated. Indeed, an obvious method is to compute $P_\mathbf{Y}^{h_k}(\mathbf{y})$ for various values of step size $\{h_k, k=1\dots K\}$. A simple linear regression of $P_\mathbf{Y}^{h_k}(\mathbf{y})$ against $h_k^p$ gives an estimation of $B(\mathbf{y})$. This means that using a multi-resolution computation of the $P_\mathbf{Y}^{h_k}(\mathbf{y})$ on various approximate models, we are able to estimate the marginal likelihood of the true model. \medskip $\bullet$ This last point is of major importance. Assume that, two models $\mathcal{M}_1$ and $\mathcal{M}_2$ have to be compared, one or both of them being defined by a differential system without explicit solution. These two models can be compared per se. Indeed, whereas one could have thought that only the approximate models $\mathcal{M}_1^h$ and $\mathcal{M}_2^{h}$ where comparable, we establish that the both true or exact models $\mathcal{M}_1$ and $\mathcal{M}_2$ may be directly compared. \medskip In the next section, we develop two examples where we estimate the Bayes factor of the theoretically exact model with approximate models. As a consequence the step size may be coarsened, obtaining basically the same posterior distributions, but at far lower CPU costs. \section{Numerical examples}\label{sec:numerical examples} \subsection{Logistic Growth Models}\label{sec:exa1} We base our first numerical study on the logistic growth model which is a common model of population growth in ecology. Recently it has also been used to model tumors growth in medicine, among many other applications. Let $X(t)$ be the size of the tumor to time $t$. The dynamics are governed by the following differential equation \begin{equation}\label{eq:logistic} \frac{dX}{dt} =\lambda X(t) (K-X(t)), \quad X(0)=X_0 \end{equation} with $r = \lambda K$ being the growth rate and $K$ the carrying capacity e.g. $\lim_{t\rightarrow \infty} X(t) = K$. Equation (\ref{eq:logistic}) has an explicit solution equal to $$ X(t) = \frac{KX_0e^{\lambda Kt}}{K + X_0(e^{\lambda K t}-1)}. $$ We simulate two synthetic data sets with the error model $ y_i =X(t_i) + \varepsilon_i$, where $\varepsilon_i \sim \mathcal{N}(0,\sigma^2)$, and the following parameters $X(0)=100, \quad \lambda= 1, \quad K=1000, \quad \sigma = 1 \mbox{ or } 30.$ The datasets are plotted on Figure \ref{fig.data} for the two chosen values of $\sigma$. We consider $26$ observations at times $t_i$ regularly spaced between $0$ and $10$. \begin{figure} \centering \begin{tabular}{c l} \includegraphics[width=6cm]{Data_s1.png} & \includegraphics[width=6cm]{Data_s30.png} \end{tabular} \caption{Synthetic data for the Logistic growth with $\lambda=1$, $K=1000$ and $\sigma=1$ (left) or $\sigma=30$ (right).} \label{fig.data} \end{figure} \medskip For this first toy example, $K$ is taken as known and inference is concentrated on the single parameter $\lambda$; we consider a Gamma distribution for the prior on $\lambda$. \medskip To highlight our result presented in Section \ref{sec:posterior}, we consider the following strategy. For $\sigma=1$, we first compute what we call the ``true" marginal likelihood $P_\mathbf{Y}(\mathbf{y})$ (red line in Figure \ref{fig.ExaLog1}), using the explicit solution of (\ref{eq:logistic}) and numerical integration. In a second step, we approximate the solution of (\ref{eq:logistic}) by the Euler scheme (equivalent to the Runge-Kutta solver of order 1), for various step sizes $h_k$. The marginal likelihood $P_\mathbf{Y}^{h_k(1)}(\mathbf{y})$ is computed using both numerical integration and the Monte Carlo strategy presented in Section \ref{subset:comp Bayes factor}. These results are plotted in black in Figure \ref{fig.ExaLog1}, with solid thin lines for the numerical integration and triangles for the Monte Carlo computation. In a third step, the same strategy is applied but using a Runge-Kutta solver of order $2$ (in green in Figure \ref{fig.ExaLog1}). Finally, we use a classical Runge-Kutta solver of order $4$ (in green in Figure \ref{fig.ExaLog1}). In the end, for each order $p$, the estimated values $\hat{P}_\mathbf{Y}^{h_k( p)}(\mathbf{y})$ are used to compute the regression functions $\hat{P}_\mathbf{Y}^{h_k( p)}(\mathbf{y}) = a + bh^p$ (thick lines in black, green and blue on Figure \ref{fig.ExaLog1}). These results are presented in Figures \ref{fig.ExaLog1} and \ref{fig.ExaLog2}. \medskip The same is done for $\sigma=30$ but only the RK solver of order 4 is considered. The equivalent results are presented in Figure \ref{fig.ExaLog4}. The samples from the posterior distributions are obtained using a t-walk MCMC algorithm \citep{Christen&Fox2010}. Next we discuss some aspects of this numerical experiment. \begin{figure} \centering \includegraphics[height=10cm, width=12cm]{BFs_s1} \caption{\emph{Study on synthetic data for the Logistic growth with $\sigma=1$}. Marginal $P_\mathbf{Y}^h(\mathbf{y})$ for various step sizes, computed by numerical integration (solid thin lines) or estimated using the MCMC sample (triangles). In black, Runge-Kutta solver (RK) of order 1 (Euler), in green RK of order $p=2$, in blue RK of order $p=4$. Red line: true marginal $P_\mathbf{Y}(\mathbf{y})$ calculated using numerical integration on the analytic solution. Thick lines indicate the regression for estimated values for $\hat{P}_\mathbf{Y}^h(\mathbf{y}) = a + b h^p$ for the orders $p=1,2,4$.} \label{fig.ExaLog1} \end{figure} \medskip $\bullet$ We would like to highlight that the thick lines and the triangles are quite similar, meaning that the Monte Carlo strategy (derived as a by-product of the MCMC implementation) is an efficient solution to estimating the marginal likelihood in this context. This approximation has the great advantage of not to involving any new ODE solver runs and thus has a minimal computational cost. \medskip $\bullet$ As predicted, the Euler scheme has a linear approximation regime to the correct marginal. To have any substantial save in CPU without compromising posterior inference precision we would need to have a very small step size, i.e. there is no `flat part' in order to take considerable larger step sizes. On the contrary, the Runge-Kutta solvers of order 2 and specially the classical RK of order 4, they indeed have a clear flat section where a nearly perfect estimation has been reached. This allows for choosing a much larger step size, meaning a far coarser ODE numerical solver which still has basically no difference in the resulting inference and may be seen in the resulting posterior distributions in Figures~\ref{fig.ExaLog2}(b) and~\ref{fig.ExaLog4}(c). \medskip $\bullet$ For the RK or order $4$, we perform a linear regression using the estimated $\hat{P}_\mathbf{Y}^{h_k}(\mathbf{y})$ with $h_k= 0.2$, $0.1$, $0.05$, $0.025$ for $\sigma=1$ and $h_k= 0.8$, $0.6$, $0.4$, $0.2$, $0.1$, $0.05$ for $\sigma=30$. Using the formula given in Proposition \ref{prop:2}, we deduce an estimation (projection) of the exact marginal likelihood $\hat{P}_\mathbf{Y}(\mathbf{y})$, which has to be compared to the true value $P_\mathbf{Y}(\mathbf{y})$ (obtained using the exact solution of the ODE and a numerical integration). The results are given in the following Table~\ref{tab.comp}. \begin{table} \begin{center} \begin{tabular}{ccc} \hline $\sigma$ & $P_\mathbf{Y}(\mathbf{y})$ & $\hat{P}_\mathbf{Y}(\mathbf{y})$\\ \hline $1$ & $1.854\, 10^{-18}$ & $1.862\, 10^{-18}$\\ $30$ & $1.638\, 10^{-60}$ & $1.699\, 10^{-60}$\\ \hline \end{tabular} \caption{\label{tab.comp}Comparison of exact an estimated marginals for the Ringue-Kutta method of order 4.} \end{center} \end{table} $\bullet$ We believe that the most important message is that, for the RK of order $4$, as soon as $h$ is lower than some threshold (we took 0.05 for $\sigma=1$ and 0.1 for $\sigma=30$), the Bayes factor ratio $P_\mathbf{Y}^{h_k(2)}(\mathbf{y}) / P_\mathbf{Y}(\mathbf{y})$ is greater than 0.99, making the models indistinguishable on the Jeffrey's Bayesian scale and leading to nearly identical posterior distributions (see Figures~\ref{fig.ExaLog2}(b) and~\ref{fig.ExaLog4}(c)) for $\lambda$. However, the computational time required to estimate the parameters using the smallest $h$ explodes (see Figures~\ref{fig.ExaLog2}(a) and~\ref{fig.ExaLog4}(b)) from $2$ min for $h=0.05$ to $17$ min for $h=0.01$, for $\sigma=1$, and from $2.5$ min for $h=0.1$ to $36$ min for $h=0.000625$, for $\sigma=30$. \begin{figure} \begin{center} \begin{tabular}{c l} (a) \includegraphics[height=5cm, width=5cm]{CPU_s1} & (b) \includegraphics[height=5cm, width=5cm]{Posts1vs2_s1} \\ \end{tabular} \\ \caption{(a) CPU time for various step values $h_k$ and $p=1,2,4$, relative to 10,000 iterations of the MCMC. (b) Posterior distribution of $\lambda$ the for RK4 solver, $p=4$, for step sizes $h=0.01$ and $h=0.06$ (histograms) and exact posterior (black density). 10,000 iterations of the MCMC took 17 min for $h=0.01$ and 2 min for $h=0.05$; a 90\% reduction in CPU time with no noticeable difference in the resulting posterior distribution.} \label{fig.ExaLog2} \end{center} \end{figure} \smallskip $\bullet$ Note that the ranges of considered values for $h_k$ are different for $\sigma=1$ and $\sigma=30$. This has to be linked to the remark we have made above: the error induced by the numerical integration of the ODE can not be considered per-se but with respect to the observation noise. When $\sigma=30$, the step size $h^*$ such that for any $h \leq h^*$ all the models $\mathcal{M}^h$ are equivalent on the Jeffrey's scale is much more higher involving even larger computational time savings. \begin{figure} \begin{center} \begin{tabular}{c l} \multicolumn{2}{c}{(a)\includegraphics[height=7cm, width=10cm]{BFs_s30}}\\ (b)\includegraphics[height=5cm, width=5cm]{CPU_s30}& (c)\includegraphics[height=5cm, width=5cm]{Posts1vs2_s30} \\ \end{tabular} \\ \caption{\label{fig.ExaLog4} \emph{Study on synthetic data for the Logistic growth with $\sigma=30$}. (a) Marginal $P_\mathbf{Y}^h(\mathbf{y})$ for various step sizes, both exact (solid thin lines, using numerical integration) and estimated using the MCMC sample (triangles). We use a Rungue-Kutta solver of order 4 (classical RK4, blue), only. Red line: true marginal $P_\mathbf{Y}(\mathbf{y})$ calculated using numerical integration on the analytic solution. Thick lines indicate the regression for estimated values for $\hat{P}_\mathbf{Y}^h(\mathbf{y}) = a + b h^p$ for the order $p=4$. (b) Corresponding CPU time, relative to 10,000 iterations of the MCMC. (c) Posterior distribution of $\lambda$ the for RK4 solver, $p=4$, for step sizes $h=0.00625$ and $h=0.1$ (histograms; and exact posterior, black density). The former takes 36 min and the latter 2.5 min.} \end{center} \end{figure} \subsection{A Diabetes minimal model}\label{sec:exa2} We know illustrate our results on a real dataset and a more complex model for an Oral Oral Glucose Tolerance Test (OGTT). After having briefly described the experiment and the model, we present our results. \medskip An Oral Glucose Tolerance Test (OGTT) is performed for diagnosis of diabetes, methabolic syndrome and other conditions. After a night sleep, fasting patients are measured for blood glucose and asked to drink a sugar concentrate. Blood glucose is then measured at the hour, two hours and sometime at three hours, depending on local practices. We are developing a minimal model for blood glucose-insulin interaction base on a two compartment model. One simple transfer compartment of glucose in the digestive system and one more complex compartment for blood glucose and interactions with Insulin and other glucose substitution mechanisms. Here we present this model to show our methodology estimating one parameter only (namely, the insulin sensitivity). Although there is no analytical solution we are able to find the marginal $P_\mathbf{Y}^h(\mathbf{Y})$ by numerical integration (since there is only one parameter involved) for comparison purposes. \medskip Let $G(t)$ be blood glucose level at time $t$, in mg/dL. Let $I(t)$ be blood insulin level at time $t$ and $L(t)$ ``glucagon'' levels, to promote liver Glycogen glucose production, in arbitrary units. Let $D(t)$ be the digestive system `glucose level'; we take it as a compartment in which glucose is first stored (eg. stomach and digestive tract) and in turn delivered into the blood stream (we state $D(t)$ in the same units as for $G(t)$ and therefore the mean life parameter $\theta_2$ in (\ref{eqn.minmod.D}) is the same as the one used in (\ref{eqn.minmod.G}) below). Let also $G_b$ be the glucose base line, (=80 mg/dL, fixed). Our model is described by the following system of differential equations \begin{eqnarray} \label{eqn.minmod.G} \frac{dG}{dt} & = & \left(L - I \right) G + \frac{D}{\theta_2} , \\ \label{eqn.minmod.I} \frac{dI}{dt} & = & \theta_0 \left( \frac{G}{G_b} - 1 \right)^+ - \frac{I}{a} , \\ \label{eqn.minmod.L} \frac{dL}{dt} & = & \theta_1 \left( 1 - \frac{G}{G_b} \right)^+ - \frac{L}{b} , \\ \label{eqn.minmod.D} \frac{dD}{dt} & = & - \frac{D}{\theta_2} . \end{eqnarray} \medskip A brief explanation of the model goes as follows. When glucose goes above the normal threshold $G_b$, Insulin is produced, ie. its derivative increases, see (\ref{eqn.minmod.I}). This, in turn, acts on blood glucose to decrease its concentration; a mass-action type term is introduced in (\ref{eqn.minmod.G}) to decrease the derivative of $G$. $L$ is an abstract term related to the glucose recovery system. When Glucose $G(t)$ goes below the normal threshold ($G_b$) $L$ increases, see (\ref{eqn.minmod.L}), to increase the derivative of $G(t)$ (thus eventually increasing the glucose), see (\ref{eqn.minmod.G}). Finally, $D(t)$ represents the glucose in the digestive compartment that will be transferred to the blood stream, see (\ref{eqn.minmod.D}) and (\ref{eqn.minmod.G}). We analyze data from an OGTT conducted in an obese male adult patient with a suspected methabolic syndrome condition; the corresponding data are plotted on Figure~\ref{fig.ExaDiabMinMod.data}. All parameters are positive and will be set to $\theta_1 = 26.6, \theta_2 = 0.2, a=1, b=2$, while $\theta_0$ is taken as unknown and will be estimated from data. This is an unusual experimental data set in which Glucose was measured every 30 min up to 2 hr. \begin{figure} \begin{center} \includegraphics[height=8cm, width=12cm]{Data_DiabMinMod} \caption{\label{fig.ExaDiabMinMod.data} OGTT test performed in an obese male adult, with glucose measurents taken every 30 min up to 2 hr. Note the oscilating nature of the data, typically of a not well control Insulin-Glucose system. Both $\theta_0$ and $\theta_1$ have large values in comparison to normal subjects. The MAP model is shown in red, along with draws from the posterior predictive distribution shown in the shaded areas.} \end{center} \end{figure} \medskip Our Bayesian inference is performed as follows. We have observations $d_1, d_2, \ldots, d_n$ for Glucose, thus we let $$ d_i = G(t_i) + e_i ~~\text{where}~~e_i \sim N(0, \sigma ), $$ and $G(0) = d_0$ the initial condition; we fix the measurement error to $\sigma = 5$ (the observation functional is therefore $f(X) = X_1$). From this a likelihood is constructed. A Gamma prior distribution is assumed for parameter $\theta_0$, with shape parameter $=5$ and rate $=2/5$, thus with mean $=2$, apparently a value for a normal person. Using an order 4 Rugue-Kutta solver with varying step size we perform a MCMC for these parameters using the t-walk (Christen and Fox, 2010). \begin{figure} \begin{center} \begin{tabular}{c l} \multicolumn{2}{c}{(a)\includegraphics[height=7cm, width=10cm]{BFs_DiabMinMod}}\\ (b)\includegraphics[height=5cm, width=5cm]{CPU_DiabMinMod}& (c)\includegraphics[height=5cm, width=5cm]{Posts1vs2_DiabMinMod} \\ \end{tabular} \\ \caption{\label{fig.ExaDiabMinMod} Bayes Factors study for the Diabetes minimal model. An order 4 Runge-Kutta solver was used to produce marginal values $P_\mathbf{Y}^h(\mathbf{Y})$ for step sizes as show in (a) and their corresponding CPU times are depicted in (b). The red line in (a) is the numerical integration approximation of $P_\mathbf{Y}^h(\mathbf{Y})$ using step size $0.25 \cdot 2^{-7}$ (smallest step size used) while the triangles are Monte Carlo estimates performed as in (\ref{eqn.estMarg}); these seem to slightly underestimate the former. The solid blue line is a regression model $a + b h^4$ estimate using step sizes marginal estimates from $0.25 \cdot 2^{-1}$ to $0.25 \cdot 2^{-4}$ only. In (c) we compare the resulting posterior with step size $0.25 \cdot 2^{-3}$ and $0.25 \cdot 2^{-7}$ showing basically no difference and resulting in a near 90\% reduction in CPU evaluation time.} \end{center} \end{figure} To make the solver evaluation time steps (in hr) include the observation times we take a rough time step of 15min (=0.25hr) and divide it into finer time steps defined as $h^{(k)} = 0.25 \cdot 2^{-k}$. Our experiments included $k=0, \ldots , 7$, as seen in in Figure~\ref{fig.ExaDiabMinMod}(a). We only use an order 4 Runge-Kutta solver, resulting in the 4th grade polynomial regression and a flat section already at $h^{(3)} = 0.25 \cdot 2^{-3} = 1.875$min. If compared with our minimum time step of $h^{(7)} = 0.25 \cdot 2^{-7} = 7$sec, the resulting CPU time of the MCMC is more than 90\% larger. However, the resulting posterior distributions for $h^{(3)}$ and $h^{(7)}$ are basically identical (see Figure~\ref{fig.ExaDiabMinMod}(c)). The estimated marginal is $P_{\mathbf{Y}}^{h^{(3)}}(\mathbf{y}) \approx 3\cdot 10^{-14}$, calculated using the MCMC samples and (\ref{eqn.estMarg}), while $P_{\mathbf{Y}}^{h^{(7)}}(\mathbf{y}) \approx 3.2\cdot 10^{-14}$ calculated using numerical integration (red line in Figure~\ref{fig.ExaDiabMinMod}(a)). \section{Discussion}\label{sec:discussion} We advance on some theoretical aspects of the Bayesian analysis of ODE systems. As opposed to more standard (Bayesian) statistical analyses, Inverse Problems present the added difficulty that the regressor function is not analytically tractable and numerical approximations need to be used. In general, the replacement of the theoretical (non-available) solution of the differential system by a numerical approximation is ignored and the solver being used as a black box. However, recently, research has been directed at trying to quantify the consequences of such an approximation, commonly by comparing expected values of the resulting posterior distributions, like the exact vs the numerical Posterior means. In this paper we adopt a different approach, basing our comparison on the use of Bayes Factors, which is the natural tool to comparing models in a Bayesian context. There are still some particular issues to be solve when applying our results to more realistic inverse problems like estimating the marginals in a multidimensional parameter problem and analyzing stiff problems were a multistep method would need to be used. However, we may highlight the following remarks. First, we contribute to the intuitive idea that the ODE solver approximation error should be put in the perspective of the observational error. Bayes Factors, and the Bayesian model comparison machinery, can be used as an appropriate measure of the solver accuracy, precisely in the perspective of the observational error considered in the model. Result~\ref{prop:2} establishes a consistency in order accuracy for the solver and for the posterior distribution, considering BF's. As a consequence, numerical solver precision may be viewed in this perspective and not solely as a black box regressor. As far as the main aim is to make inference on parameters, there is no need to use to highest precision if the data are contaminated by a non-neglectable quantity of noise. In a domain where the computational time is important, we have proved that considerable time savings can be done, only by using a reasonable step size in the solver. Secondly, we show how the BF may be approximated even in this scenario where the exact model is not available. This result is of particular interest, since it allows to compare the accuaracy of our approximate posterior without being able to work on the theoretical model directly. The computation of marginal likelihoods is an important topic in the Bayesian literature. In this paper, we propose the use of the Gelfand and Dey's estimator, which has the great advantage of not requiring any additional numerical evaluation of the differential system, after the MCMC was performed. However, we are aware that the Gelfand and Dey's estimator may be highly unstable when the dimension of the parameters increases. The use of a Kernel Density Estimate weighting function in (\ref{eqn.estMarg}) can help to stabilize the estimate but is not a universal solution. If the dimension of the parameters increases, other strategies should be considered, still keeping in mind that any additional numerical evaluation of the ODE system may have a considerable computational cost. Our results would also need to be stated for multiple dimension observation functions $f$; we leave these considerations for future research. \section{Acknowledgments} We thank Dr Silvia Quintana for kindly providing the OGTT data for example in~\ref{sec:exa2}. MAC and JAC would like to acknowledge financial support from Fondo Mixto de Fomento a la Investigaci\'on Cient\'{\i}fica y Tecnol\'ogica, CONACYT-Gobierno del Estado de Guajanuato, GTO-2011-C04-168776. \bibliographystyle{gSCS.bst}
\section{Introduction} The first, second and fourth Painlev\'{e} equations in Hamiltonian forms are given by \begin{eqnarray} (\text{P}_\text{I}) \left\{ \begin{array}{l} \displaystyle \frac{dx}{dz} = 6y^2 + z \\[0.2cm] \displaystyle \frac{dy}{dz} = x, \\ \end{array} \right. \label{P1} \end{eqnarray} \begin{eqnarray} (\text{P}_\text{II}) \left\{ \begin{array}{l} \displaystyle \frac{dx}{dz} = 2y^3 + yz +\alpha \\[0.2cm] \displaystyle \frac{dy}{dz} = x, \\ \end{array} \right. \label{P2} \end{eqnarray} \begin{eqnarray} (\text{P}_\text{IV}) \left\{ \begin{array}{l} \displaystyle \frac{dx}{dz} = -x^2 + 2xy + 2xz - 2\theta _{\infty} \\[0.2cm] \displaystyle \frac{dy}{dz} = -y^2+2xy-2yz-2\kappa_0, \\ \end{array} \right. \label{P4} \end{eqnarray} with Hamiltonian functions \begin{eqnarray*} & & H_\text{I} = \frac{1}{2}x^2 - 2y^3 - zy, \\ & & H_\text{II} = \frac{1}{2}x^2 - \frac{1}{2}y^4 - \frac{1}{2}zy^2 - \alpha y, \\ & & H_\text{IV} = -xy^2 + x^2y - 2xyz - 2\kappa_0 x + 2\theta _\infty y, \end{eqnarray*} where $\alpha , \theta _\infty$ and $\kappa_0 \in \C$ are parameters. These equations are investigated by means of the weighted projective spaces $\C P^3(p,q,r,s)$ with natural numbers $p,q,r,s$ given by \begin{eqnarray*} (\text{P}_\text{I}) & & (p,q,r,s) = (3,2,4,5), \\ (\text{P}_\text{II}) & & (p,q,r,s) = (2,1,2,3), \\ (\text{P}_\text{IV}) & & (p,q,r,s) = (1,1,1,2). \end{eqnarray*} These numbers will be determined by the Newton diagrams of the equations or the versal deformations of a certain class of dynamical systems. The weighted projective space $\C P^3(p,q,r,s)$ is a three dimensional compact orbifold (toric variety) with singularities, see Sec.2 for the definition. ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are given as differential equations on $\C P^3(p,q,r,s)$, which is regarded as a compactification of the original phase space $\C^3_{(x,y,z)}$ of the Painlev\'{e} equations. The Painlev\'{e} equations are invariant under the $\Z_s$ action of the form \begin{equation} (x,y,z) \mapsto (\omega ^px, \omega ^qy, \omega ^rz), \quad \omega := e^{2\pi i/s}, \label{1-4} \end{equation} with $p,q,r,s$ as above. As a result, it turns out that ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are well defined as meromorphic differential equations on $\C P^3(p,q,r,s)$. The space $\C P^3(p,q,r,s)$ is decomposed as \begin{equation} \C P^3(p,q,r,s) = \C^3/\Z_s \,\, \cup \,\, \C P^2(p,q,r), \quad (\text{disjoint}). \label{1-5} \end{equation} This means that $\C P^3(p,q,r,s)$ is a compactification of $\C^3/\Z_s$ obtained by attaching a 2-dim weighted projective space $\C P^2(p,q,r)$ at infinity. The Painlev\'{e} equations ($\text{P}_\text{J}$),\, ($\text{J} = \text{I}, \text{II},\text{IV}$) divided by the $\Z_s$ action are given on $\C^3/\Z_s$, and the 2-dim space $\C P^2(p,q,r)$ describes the behavior of ($\text{P}_\text{J}$) near infinity (i.e. $x=\infty$ or $y=\infty$ or $z=\infty$). On the ``infinity set" $\C P^2(p,q,r)$, there exist several singularities of the foliation defined by solutions of the equation. Some of them correspond to movable poles of ($\text{P}_\text{J}$), and the others correspond to the irregular singular point $z=\infty$. Local properties of these singularities of the foliation will be investigated by means of dynamical systems theory. Our main results include \begin{itemize} \item the fact that the Painlev\'{e} equations are locally transformed into integrable systems near movable singularities, \item a simple proof of the fact that any solutions of ($\text{P}_\text{J}$) are meromorphic on $\C$, \item a simple construction of the symplectic atlas of Okamoto's space of initial conditions, \item for ($\text{P}_\text{I}$), a geometric interpretation of Painlev\'{e}'s coordinates defined by \begin{equation} \left\{ \begin{array}{l} x = uw^3 - 2w^{-3} - \frac{1}{2}z w - \frac{1}{2} w^2 \\ y = w^{-2}, \\ \end{array} \right. \label{1-6} \end{equation} which was introduced in his original work \cite{Pai} to prove the Painlev\'{e} property of ($\text{P}_\text{I}$), \item a geometric interpretation of Boutroux's coordinates introduced in \cite{Bou} to investigate the irregular singular point of ($\text{P}_\text{I}$) and ($\text{P}_\text{II}$). \end{itemize} In Sec.2, the Newton diagram of the Painlev\'{e} equation will be introduced to find a suitable weight of the weighted projective space $\C P^3(p,q,r,s)$. Furthermore, it is shown that the Painlev\'{e} equations are obtained from certain problems of dynamical systems theory. Such a relationship between the Painlev\'{e} equations and dynamical systems proposes normal forms of the Painlev\'{e} equations because for dynamical systems (germs of vector fields), the normal form theory have been well developed. In Sec.3, with the aid of the orbifold structure of $\C P^3(p,q,r,s)$ and the Poincar\'{e} linearization theorem, it will be shown that ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are locally transformed into integrable systems near each movable singularities. For example, ($\text{P}_\text{I}$) and ($\text{P}_\text{II}$) can be transformed into the equations $y'' = 6y^2$ and $y'' = 2y^3$, respectively. See Sec.3 for the result for ($\text{P}_\text{IV}$). This fact was first obtained by \cite{CosCos} for ($\text{P}_\text{I}$), in which the transformed equation $y'' = 6y^2$ is called the singular normal form. Our proof is based on the Poincar\'{e} linearization theorem and it is easily applied to other Painlev\'{e} equations, including ($\text{P}_\text{III}$), ($\text{P}_\text{V}$) ($\text{P}_\text{VI}$) and higher order Painlev\'{e} equations. By using this result, a simple proof of the Painlev\'{e} property is proposed; that is, a new proof of the fact that any solutions of ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are meromorphic on $\C$ will be given. In Sec.4, the weighted blow-up will be introduced to construct the spaces of initial conditions. For a polynomial system, a manifold $E(z)$ parameterized by $z\in \C$ is called the space of initial conditions if any solutions give global holomorphic sections on the fiber bundle $\mathcal{P} = \{ (x,z) \, | \, x\in E(z) , z\in \C\}$ over $\C$. It is remarkable that only one, two and three times blow-ups are sufficient to obtain the spaces of initial conditions for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$), respectively, if we use suitable weights, while Okamoto performed blow-ups (without weights) eight times to obtain the space of initial conditions \cite{Oka}. Further, our method easily provides a symplectic atlas of the space of initial conditions. Then, each Painlev\'{e} equation is characterized as a unique Hamiltonian system on the space of initial conditions admitting a holomorphic symplectic form. Symplectic atlases of the spaces of initial conditions were first obtained by Takano \textit{et al.} \cite{Tak1, Tak2, Tak3} only for ($\text{P}_\text{II}$) to ($\text{P}_\text{VI}$), while left open for ($\text{P}_\text{I}$). In the present paper, the orbifold structure plays an important role to obtain a symplectic atlas for ($\text{P}_\text{I}$), see also Iwasaki and Okada \cite{Iwa} for the orbifold setting of ($\text{P}_\text{I}$). By the weighted blow-up of $\C P^3(3,2,4,5)$ for ($\text{P}_\text{I}$), we will recover the famous Painlev\'{e}'s coordinates (\ref{1-6}) in a purely geometric manner. Painlev\'{e} found the coordinate transformation (\ref{1-6}) in an analytic way to prove the Painlev\'{e} property of ($\text{P}_\text{I}$) (see \cite{Gro}). From our approach based on the weighted projective space, Painlev\'{e}'s coordinates prove to be nothing but the Darboux coordinates of the nonsingular algebraic surface $M(z)$ defined by \begin{eqnarray*} V^2 = UW^4 + 2z W^3 + 4W, \end{eqnarray*} which admits a holomorphic symplectic form, where $z \in \C$ is an independent variable of ($\text{P}_\text{I}$) and it is a parameter of the surface. Our space of initial conditions is obtained by glueing $\C^2_{(x,y)}$ (the original space for dependent variables) and the surface $M(z)$ by a symplectic mapping. Then, ($\text{P}_\text{I}$) is a Hamiltonian system with respect to the symplectic form. Since (\ref{1-6}) is a one-to-two transformation, an orbifold setting is essential to give a geometric meaning to Painlev\'{e}'s coordinates; the orbifold $\C P^3 (3,2,4,5)$ provides a natural $\Z_2$-action which makes (\ref{1-6}) a one-to-one transformation. In Sec.5, the characteristic indices for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) will be defined. A few simple properties such as a relation with the Kovalevskaya exponents and the weights of $\C P^3 (p,q,r,s)$ will be given. In Sec.6, the Boutroux coordinates will be introduced. It is shown that the weighted blow-ups of $\C P^3(p,q,r,s)$ constructed in Sec.4 also includes the space of initial conditions written in the Boutroux coordinates. Further, we will show that autonomous Hamiltonian systems are embedded in the Boutroux coordinates. In Sec.7, the extended affine Weyl group for ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) will be considered. The action of the group on the original chart $\C^3_{(x,y,z)}$ is extended to a birational transformation on $\C P^3(p,q,r,s)$. It is proved that on the ``infinity set", $\C P^2(p,q,r)$, the foliation defined by an autonomous Hamiltonian system is invariant under the automorphism group $\mathrm{Aut} (X)$, where $X = A^{(1)}_1$ for ($\text{P}_\text{II}$) and $X = A^{(1)}_2$ for ($\text{P}_\text{IV}$). In Sec.8, a cellular decomposition of the weighted blow-ups of $\C P^3(p,q,r,s)$ will be given. We will show that the weighted blow-ups of $\C P^3(p,q,r,s)$ is naturally decomposed into the fiber space for ($\text{P}_\text{J}$) (a fiber bundle over $\C$ whose fiber is the space of initial conditions), a certain elliptic fibration over the moduli space of complex tori, and the projective curve $\C P^1$. We also show that the extended Dynkin diagrams of type $\tilde{E}_8, \tilde{E}_7$ and $\tilde{E}_6$ are hidden in the weighted blow-ups of $\C P^3(p,q,r,s)$. An approach using toric varieties is also applicable to the third, fifth, sixth Painlev\'{e} equations and higher order Painlev\'{e} equations, which will appear in a forthcoming paper. \section{Weighted projective spaces} In this section, a weighted projective space $\C P^3(p,q,r,s)$ is defined and the first, second and fourth Painlev\'{e} equations are given as meromorphic equations on $\C P^3(p,q,r,s)$ for suitable integers $p,q,r,s$. Such integers $p,q,r,s$ will be found via the Newton diagrams of the equations. We also give a relationship between the Painlev\'{e} equations and the normal form theory of dynamical systems, which proposes normal forms of the Painlev\'{e} equations. \subsection{Newton diagram} Let us consider the system of polynomial differential equations \begin{eqnarray} \frac{dx}{dz} = f_1(x,y,z), \quad \frac{dy}{dz} = f_2(x,y,z). \label{2-1} \end{eqnarray} The exponent of a monomial $x^iy^jz^k$ included in $f_1$ is defined by $(i-1,j,k+1)$, and by $(i,j-1,k+1)$ for one in $f_2$. Each exponent specifies a point of the integer lattice in $\R^3$. The Newton polyhedron of the system (\ref{2-1}) is the convex hull of the union of the positive quadrants $\R_+^3$ with vertices at the exponents of the monomials which appear in the system. The Newton diagram of the system is the union of the compact faces of its Newton polyhedron. Suppose that the Newton diagram consists of only one compact face. Then, there is a tuple of positive integers $(p_1, p_2, r, s)$ such that the compact face lies on the plane $p_1 x + p_2 y + rz = s$ in $\R^3$. In this case, the function $f_i\, (i=1,2)$ satisfies \begin{eqnarray*} f_i(\lambda ^{p_1}x, \lambda ^{p_2}y , \lambda ^r z) = \lambda ^{s-r+p_i} f_i (x_1, \cdots ,x_m,z), \end{eqnarray*} for any $\lambda \in \C$. We also consider the perturbation of the system (\ref{2-1}) of the form \begin{eqnarray} \frac{dx}{dz} = f_1(x,y,z)+g_1(x,y,z), \quad \frac{dy}{dz} = f_2(x,y,z)+g_2(x,y,z). \label{2-1b} \end{eqnarray} Suppose that $g_i(\lambda ^{p_1}x , \lambda ^{p_2}y , \lambda ^{r}z) \sim o(\lambda ^{s-r+p_i})$ for $i=1,2$ as $\lambda \to \infty$. This implies that exponents of any monomials included in $g_i$ lie on the lower side of the plane $p_1 x + p_2 y + rz = s$. The Newton polyhedron of the first Painlev\'{e} equation (\ref{P1}) is defined by three points $(-1,2,1), (-1,0,2)$ and $(1,-1,1)$. Hence, the Newton diagram consists of the unique face which lies on the plane $3x+2y + 4z = 5$. One of the normal vector to the plane is given by $\mathbf{e}_0 = (-3/5, -2/5, -4/5)$. Put $\mathbf{e}_1 = (1,0,0), \mathbf{e}_2 = (0,1,0)$ and $\mathbf{e}_3 = (0,0,1)$. Then, the toric variety defined by the fan made up of the cones generated by all proper subsets of $\{ \mathbf{e}_0, \mathbf{e}_1,\mathbf{e}_2,\mathbf{e}_3\}$ is the weighted projective space $\C P^3(3,2,4,5)$ \cite{Cox}. Next, let us consider the second Painlev\'{e} equation (\ref{P2}) with $f = (2y^3 + yz, x)$ and $g= (\alpha ,0)$. The Newton polyhedron of $f = (f_1, f_2)$ is defined by three points $(-1,3,1), (-1,1,2), (1,-1,1)$, and the Newton diagram is given by the unique face on the plane $2x + y+ 2z = 3$. The associated toric variety is $\C P^3(2,1,2,3)$. For the fourth Painlev\'{e} equation (\ref{P4}), put $f = (-x^2 + 2xy + 2xz, -y^2+2xy-2yz)$ and $g= (- 2\theta _{\infty} ,-2\kappa_0)$. The Newton diagram of $f = (f_1, f_2)$ is given by the unique face on the plane $x + y+ z = 2$ passing through the exponents $(1,0,1), (0,1,1)$ and $(0,0,2)$. The associated toric variety is $\C P^3(1,1,1,2)$. In what follows, the weights $(p,q,r,s)$ denote $(3,2,4,5), (2,1,2,3)$ and $(1,1,1,2)$, respectively, for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$). The weighted degree of a monomial $x^iy^jz^k$ with respect to the weight $(p,q,r)$ is defined by $\mathrm{deg} (x^iy^jz^k) = pi+qj+rk$. The weighted degree of a polynomial $f = \sum a_{ijk} x^iy^jz^k$ is defined by \begin{eqnarray*} \mathrm{deg} (f) = \max_{i,j,k} \{ \mathrm{deg} (x^iy^jz^k) \, | \, a_{ijk} \neq 0\}. \end{eqnarray*} For ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) with the weights $(p,q,r) = (3,2,4), (2,1,2)$ and $(1,1,1)$, respectively, the weighted degrees of the Hamiltonian functions are \begin{eqnarray*} \mathrm{deg}(H_\text{I}) = 6, \quad \mathrm{deg}(H_\text{II}) = 4, \quad \mathrm{deg}(H_\text{IV}) = 3. \end{eqnarray*} They satisfy $\mathrm{deg}(H_\text{J}) = s + 1\, (\text{J} = \text{I},\text{II},\text{IV})$. Further, it will be shown that they coincide with the Kovalevskaya exponents (Sec.2.3) and the characteristic index $\lambda _1$ (Sec.5). \subsection{Weighted projective space} Let $\widetilde{U}$ be a complex manifold and $\Gamma $ a finite group acting analytically and effectively on $\widetilde{U}$. In general, the quotient space $\widetilde{U}/\Gamma $ is not a smooth manifold if the action has fixed points. Roughly speaking, a (complex) orbifold $M$ is defined by glueing a family of such spaces $\widetilde{U}_\alpha /\Gamma _\alpha $; a Hausdorff space $M$ is called an orbifold if there exist an open covering $\{ U_\alpha \}$ of $M$ and homeomorphisms $\varphi _\alpha : U_\alpha \simeq \widetilde{U}_\alpha /\Gamma _\alpha $. See \cite{Thu} for more details. In this article, only quotient spaces of the form $\C^n/ \Z_p$ will be used. Consider the weighted $\C^*$-action on $\C^{4}$ defined by \begin{equation} (x,y,z,\varepsilon ) \mapsto (\lambda ^{p}x , \lambda ^{q} y, \lambda ^rz, \lambda ^s \varepsilon ), \quad \lambda \in \C^* := \C\backslash \{ 0\}, \end{equation} where the weights $p,q,r,s$ are positive integers. We assume $1\leq p,q,r \leq s$ without loss of generality. Further, we suppose that any three numbers among them have no common divisors. The quotient space \begin{eqnarray*} \C P^3(p,q,r,s) := \C^{4}\backslash \{ 0\} /\C^* \end{eqnarray*} gives a three dimensional orbifold called the weighted projective space. The inhomogeneous coordinates of $\C P^3(p,q,r,s)$, which give an orbifold structure of $\C P^3(p,q,r,s)$, are defined as follows. The space $\C P^3(p,q,r,s)$ is defined by the equivalence relation on $\C^4\backslash \{ 0\}$ \begin{eqnarray*} (x,y,z, \varepsilon ) \sim (\lambda^p x, \lambda ^qy, \lambda ^rz, \lambda ^s \varepsilon ). \end{eqnarray*} (i) When $x\neq 0$, \begin{eqnarray*} (x,y,z,\varepsilon ) \sim (1,\,\, x^{-q/p}y, \,\, x^{-r/p}z,\,\, x^{-s/p}\varepsilon ) =:(1, Y_1, Z_1, \varepsilon _1). \end{eqnarray*} Due to the choice of the branch of $x^{1/p}$, we also obtain \begin{eqnarray*} (Y_1, Z_1, \varepsilon _1) \sim (e^{-2q\pi i /p}Y_1, e^{-2r\pi i /p}Z_1, e^{-2s\pi i /p}\varepsilon _1), \end{eqnarray*} by putting $x \mapsto e^{2\pi i }x$. This implies that the subset of $\C P^3(p,q,r,s)$ such that $x\neq 0$ is homeomorphic to $\C^3 / \Z_p$, where the $\Z_p$-action is defined as above. (ii) When $y\neq 0$, \begin{eqnarray*} (x,y,z,\varepsilon ) \sim (y^{-p/q}x,\,\, 1 ,\,\, y^{-r/q}z,\,\, y^{-s/q}\varepsilon ) =:(X_2, 1, Z_2, \varepsilon _2). \end{eqnarray*} Because of the choice of the branch of $y^{1/q}$, we obtain \begin{eqnarray*} (X_2, Z_2, \varepsilon _2) \sim (e^{-2p\pi i/q}X_2, e^{-2r\pi i/q}Z_2, e^{-2s\pi i/q} \varepsilon _2). \end{eqnarray*} Hence, the subset of $\C P^3(p,q,r,s)$ with $y\neq 0$ is homeomorphic to $\C^3 / \Z_q$. (iii) When $z\neq 0$, \begin{eqnarray*} (x,y,z,\varepsilon ) \sim (z^{-p/r}x,\,\, z^{-q/r}y ,\,\, 1,\,\, z^{-s/r}\varepsilon ) =: (X_3, Y_3, 1, \varepsilon _3). \end{eqnarray*} Similarly, the subset $\{ z \neq 0\} \subset \C P^3(p,q,r,s)$ is homeomorphic to $\C^3 / \Z_r$. (iv) When $\varepsilon \neq 0$, \begin{eqnarray*} (x,y,z,\varepsilon ) \sim (\varepsilon ^{-p/s}x,\,\, \varepsilon ^{-q/s}y ,\,\, \varepsilon ^{-r/s}z ,\,\, 1) =:(X_4, Y_4, Z_4, 1). \end{eqnarray*} The subset $\{ \varepsilon \neq 0\} \subset \C P^3(p,q,r,s)$ is homeomorphic to $\C^3 / \Z_s$. This proves that the orbifold structure of $\C P^3(p,q,r,s)$ is given by \begin{eqnarray*} \C P^3(p,q,r,s) = \C^3/\Z_p \,\, \cup \,\, \C^3/\Z_q \,\, \cup \,\, \C^3/\Z_r \,\, \cup \,\, \C^3/\Z_s. \end{eqnarray*} The local charts $(Y_1, Z_1, \varepsilon _1)$, $(X_2, Z_2, \varepsilon _2)$, $(X_3, Y_3, \varepsilon _3)$ and $(X_4, Y_4, Z_4)$ defined above are called inhomogeneous coordinates as the usual projective space. Note that they give coordinates on the lift $\C^3$, not on the quotient $\C^3 / \Z_i\,\, (i=p,q,r,s)$. Therefore, any equations written in these inhomogeneous coordinates should be invariant under the corresponding $\Z_i$ actions. In what follows, we use the notation $(x,y,z)$ for the fourth local chart instead of $(X_4, Y_4, Z_4)$ because the Painlev\'{e} equation will be given on this chart. The transformations between inhomogeneous coordinates are give by \begin{eqnarray} \left\{ \begin{array}{rrrr} x = & \varepsilon _1^{-p/s} = & X_2\varepsilon _2^{-p/s} =& X_3\varepsilon _3^{-p/s} \\ y = & Y_1\varepsilon _1^{-q/s} = & \varepsilon _2^{-q/s}=& Y_3\varepsilon _3^{-q/s}\\ z = & Z_1\varepsilon_1 ^{-r/s} = & Z_2\varepsilon _2^{-r/s}=& \varepsilon _3^{-r/s}. \end{array} \right. \label{2-3} \end{eqnarray} We give the differential equation defined on the $(x, y, z)$-coordinates as \begin{eqnarray} \frac{dx}{dz} =f(x, y, z), \quad \frac{dy}{dz} = g(x, y, z), \label{2-4} \end{eqnarray} where $f$ and $g$ are rational functions. By the transformation (\ref{2-3}), the above equation is rewritten as equations on the other inhomogeneous coordinates $(Y_1, Z_1, \varepsilon _1)$, $(X_2, Z_2, \varepsilon _2)$ and $(X_3, Y_3, \varepsilon _3)$. It is easy to verify that the equations written in the other inhomogeneous coordinates are rational if and only if Eq.(\ref{2-4}) is invariant under the $\Z_s$-action \begin{eqnarray} (x, y, z) \mapsto (\omega ^px, \omega ^qy, \omega ^r z), \quad \omega = e^{2\pi i/s}. \label{2-5} \end{eqnarray} In this case, the equations written in $(Y_1, Z_1, \varepsilon _1)$, $(X_2, Z_2, \varepsilon _2)$ and $(X_3, Y_3, \varepsilon _3)$ are invariant under the $\Z_p, \Z_q$ and $\Z_r$-actions, respectively. Hence, a tuple of these four equations gives a well-defined rational differential equation on $\C P^3(p,q,r,s)$. When $x = \infty$ or $y = \infty$ or $z = \infty$, we have $\varepsilon _1=0$ or $\varepsilon _2 = 0$ or $\varepsilon _3 = 0$. In this case, the transformation (\ref{2-3}) results in \begin{eqnarray} \left\{ \begin{array}{ll} Y_1 = X_2^{-q/p} & = Y_3 X_3^{-q/p}, \\ Z_1 = Z_2 X_2^{-r/p} & = X_3^{-r/p}. \\ \end{array} \right. \label{2-6} \end{eqnarray} The space obtained by glueing three copies of $\C^2$ by the above relations gives the 2-dim weighted projective space $\C P^2(p,q,r)$. Thus, we have obtained the decomposition \begin{equation} \C P^3(p,q,r,s) = \C^3/\Z_s \,\, \cup \,\, \C P^2(p,q,r), \quad (\text{disjoint}). \label{2-7} \end{equation} On the covering space $\C^3$ of $\C^3 /\Z_s$, the coordinates $(x, y, z)$ is assigned and Eq.(\ref{2-4}) is given. The equation on $\C P^2(p,q,r)$ is obtained by putting $\varepsilon _1 = 0$ or $\varepsilon _2 =0$ or $\varepsilon _3 = 0$, which describes the behavior of Eq.(\ref{2-4}) near infinity; \begin{equation} \C P^2 (p,q,r) = \{ \varepsilon _1 = 0\} \cup \{ \varepsilon _2 = 0\} \cup \{ \varepsilon _3 = 0\}. \label{2-8} \end{equation} Now we give the first Painlev\'{e} equation (\ref{P1}) on the fourth local chart of $\C P^3(3,2,4,5)$. By (\ref{2-4}), ($\text{P}_\text{I}$) is transformed into the following equations \begin{eqnarray} & & \frac{dY_1}{d\varepsilon _1} = \frac{3 - 12Y_1^3 - 2Y_1Z_1}{\varepsilon _1 (-30 Y_1^2 - 5Z_1)}, \quad \frac{dZ_1}{d\varepsilon _1} = \frac{3\varepsilon _1 - 24Y_1^2 Z_1 - 4 Z_1^2}{\varepsilon _1 (-30 Y_1^2 - 5Z_1)}, \label{P11} \\[0.2cm] & & \frac{dX_2}{d\varepsilon _2} = \frac{-12 - 2Z_2 + 3X_2^2}{5X_2\varepsilon _2}, \quad \frac{dZ_2}{d\varepsilon _2} = \frac{-2\varepsilon _2 + 4 X_2Z_2}{5X_2\varepsilon _2}, \label{P12} \\[0.2cm] & & \frac{dX_3}{d\varepsilon _3} = \frac{24Y_3^2 + 4 - 3X_3\varepsilon _3}{-5\varepsilon _3^2}, \quad \frac{dY_3}{d\varepsilon _3} = \frac{4X_3 - 2Y_3\varepsilon _3}{-5\varepsilon _3^2}, \label{P13} \end{eqnarray} on the other inhomogeneous coordinates. Although the transformations (\ref{2-3}) have branches, the above equations are rational due to the symmetry (\ref{2-5}) of ($\text{P}_\text{I}$). Hence, they define a rational ODE on $\C P^3(3,2,4,5)$ in the sense of an orbifold. Next, we give the second Painlev\'{e} equation (\ref{P2}) on the fourth local chart of $\C P^3(2,1,2,3)$. By (\ref{2-3}), ($\text{P}_\text{II}$) is transformed into the following equations \begin{eqnarray} & & \frac{dY_1}{d\varepsilon _1} = \frac{-2 + Y_1(2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1)}{3\varepsilon _1 (2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1)}, \frac{dZ_1}{d\varepsilon _1} = \frac{-2\varepsilon _1 + 2Z_1(2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1)}{3\varepsilon _1 (2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1)}, \quad \quad \quad \label{P21} \\[0.2cm] & & \frac{dX_2}{d\varepsilon _2} = \frac{2X_2^2 - (2+Z_2 + \alpha \varepsilon _2)}{3X_2\varepsilon _2}, \quad \frac{dZ_2}{d\varepsilon _2} = \frac{2X_2Z_2 - \varepsilon _2}{3X_2\varepsilon _2}, \label{P22} \\[0.2cm] & & \frac{dX_3}{d\varepsilon _3} = \frac{4Y_3^3 + 2Y_3 + 2\alpha \varepsilon _3 - 2X_3\varepsilon _3}{-3\varepsilon _3^2}, \quad \frac{dY_3}{d\varepsilon _3} = \frac{2X_3 - Y_3\varepsilon _3}{-3\varepsilon _3^2} \label{P23}, \end{eqnarray} on the other local charts. They define a rational ODE on $\C P^3(2,1,2,3)$ because of the symmetry (\ref{2-5}) of ($\text{P}_\text{II}$). Similarly, we give the fourth Painlev\'{e} equation (\ref{P4}) on the fourth local chart of $\C P^3(1,1,1,2)$. The equations written in the other inhomogeneous coordinates are given by \begin{eqnarray} \left\{ \begin{array}{ll} \displaystyle \frac{dY_1}{d\varepsilon _1} = \frac{-Y_1^2+2Y_1-2Y_1Z_1-2\kappa_0\varepsilon_1+Y_1(1-2Y_1-2Z_1+2\theta_\infty\varepsilon_1)}{2\varepsilon_1 (1-2Y_1-2Z_1+2\theta_\infty\varepsilon_1)},\\[0.3cm] \displaystyle \frac{dZ_1}{d\varepsilon _1} = \frac{\varepsilon _1+Z_1(1-2Y_1-2Z_1+2\theta _\infty\varepsilon _1)}{2\varepsilon _1 (1-2Y_1-2Z_1+2\theta _\infty\varepsilon _1)}, \end{array} \right. \label{P41} \end{eqnarray} \begin{eqnarray} \left\{ \begin{array}{ll} \displaystyle \frac{dX_2}{d\varepsilon _2} = \frac{ -X_2^2 + 2X_2 + 2X_2Z_2-2\theta _\infty \varepsilon _2+X_2(1-2X_2+2Z_2+2\kappa_0\varepsilon _2)}{2\varepsilon _2 (1-2X_2+2Z_2+2\kappa_0\varepsilon _2)},\\[0.3cm] \displaystyle \frac{dZ_2}{d\varepsilon _2} = \frac{\varepsilon _2 + Z_2 (1-2X_2+2Z_2+2\kappa_0\varepsilon _2)}{2\varepsilon _2 (1-2X_2+2Z_2+2\kappa_0\varepsilon _2)}, \end{array} \right. \label{P42} \end{eqnarray} \begin{eqnarray} \left\{ \begin{array}{ll} \displaystyle \frac{dX_3}{d\varepsilon _3} = \frac{-X_3^2+2X_3Y_3+2X_3-2\theta _\infty \varepsilon _3-X_3\varepsilon _3}{-2\varepsilon _3^2},\\[0.3cm] \displaystyle \frac{dY_3}{d\varepsilon _3} = \frac{-Y_3^2+2X_3Y_3-2Y_3-2\kappa_0 \varepsilon _3-Y_3\varepsilon _3}{-2\varepsilon _3^2}. \end{array} \right. \label{P43} \end{eqnarray} They define a rational ODE on $\C P^3(1,1,1,2)$. \subsection{Laurent series of solutions} Before starting the analysis of the Painlev\'{e} equations by using the weighted projective spaces, it is convenient to write down Laurent series of solutions. Since any solutions of ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are meromorphic, a general solution admits the Laurent series with respect to $T:=z-z_0$, where $z_0$ is a movable pole. For the first Painlev\'{e} equation ($\text{P}_\text{I}$), the Laurent series of a general solution is given by \begin{equation} \left( \begin{array}{@{\,}c@{\,}} x\\ y \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} \! -2 \! \\ 0 \end{array} \right) T^{-3} + \left( \begin{array}{@{\,}c@{\,}} 0 \\ 1 \end{array} \right) T^{-2} - \left( \begin{array}{@{\,}c@{\,}} \! z_0/5 \!\\ 0 \end{array} \right) T - \left( \begin{array}{@{\,}c@{\,}} 1/2 \\ \! z_0/10 \! \end{array} \right) T^2 + \left( \begin{array}{@{\,}c@{\,}} A_6 \\ -1/6 \! \end{array} \right) T^3 + \cdots , \end{equation} where $A_6$ is an arbitrary constant. For the second Painlev\'{e} equation ($\text{P}_\text{II}$), the Laurent series are expressed in two ways as \begin{eqnarray*} & (i) & \!\! \left( \begin{array}{@{\,}c@{\,}} x\\ y \end{array} \right)\! =\! \left( \begin{array}{@{\,}c@{\,}} 1 \\ 0 \end{array} \right) T^{-2} - \left( \begin{array}{@{\,}c@{\,}} 0 \\ 1 \end{array} \right) T^{-1} + \left( \begin{array}{@{\,}c@{\,}} \! z_0/6 \!\\ 0 \end{array} \right) + \left( \begin{array}{@{\,}c@{\,}} (1\! -\!\alpha )/2 \\ z_0/6 \end{array} \right) T + \left( \begin{array}{@{\,}c@{\,}} A_4 \\ B_3 \end{array} \right) T^2 + \cdots , \\ & (ii) &\!\! \left( \begin{array}{@{\,}c@{\,}} x\\ y \end{array} \right)\! =\! -\left( \begin{array}{@{\,}c@{\,}} 1 \\ 0 \end{array} \right) T^{-2} + \left( \begin{array}{@{\,}c@{\,}} 0 \\ 1 \end{array} \right) T^{-1} - \left( \begin{array}{@{\,}c@{\,}} \! z_0/6 \!\\ 0 \end{array} \right) - \left( \begin{array}{@{\,}c@{\,}} \! (1 \! + \! \alpha )/2 \! \\ z_0/6 \end{array} \right) T - \left( \begin{array}{@{\,}c@{\,}} A_4 \\ B_3 \end{array} \right) T^2 + \cdots, \end{eqnarray*} where $B_3 = (1 - \alpha )/4$ for the first line, $B_3 = (1 + \alpha )/4$ for the second line and $A_4$ is an arbitrary constant. For the fourth Painlev\'{e} equation ($\text{P}_\text{IV}$), there are three types of the Laurent series given by \begin{eqnarray*} & (i) & \!\! \left( \begin{array}{@{\,}c@{\,}} x\\ y \end{array} \right)\! =\! \left( \begin{array}{@{\,}c@{\,}} 1 \\ 0 \end{array} \right) T^{-1} + \left( \begin{array}{@{\,}c@{\,}} z_0 \\ 0 \end{array} \right) + \left( \begin{array}{@{\,}c@{\,}} \! (2+z_0^2-2\theta _\infty + 4\kappa_0)/3 \!\\ 2\kappa_0 \end{array} \right) T + \left( \begin{array}{@{\,}c@{\,}} A_3 \\ B_3 \end{array} \right) T^2 + \cdots , \\ & (ii) & \!\! \left( \begin{array}{@{\,}c@{\,}} x\\ y \end{array} \right)\! =\! \left( \begin{array}{@{\,}c@{\,}} \! -1 \! \\ \! -1 \! \end{array} \right) T^{-1}\! + \! \left( \begin{array}{@{\,}c@{\,}} z_0 \\ \! -z_0 \! \end{array} \right) \!+\! \frac{1}{3} \left( \begin{array}{@{\,}c@{\,}} \! 6-z_0^2 + 2\theta _\infty - 4\kappa_0 \!\\ \! -6-z_0^2-4\theta _\infty + 2\kappa_0 \! \end{array} \right) T \!+ \! \left( \begin{array}{@{\,}c@{\,}} A_3 \\ B_3 \end{array} \right) T^2 + \cdots , \\ & (iii) & \!\! \left( \begin{array}{@{\,}c@{\,}} x\\ y \end{array} \right)\! =\! \left( \begin{array}{@{\,}c@{\,}} 0 \\ 1 \end{array} \right) T^{-1} - \left( \begin{array}{@{\,}c@{\,}} 0 \\ z_0 \end{array} \right) - \left( \begin{array}{@{\,}c@{\,}} 2\theta _\infty\\ \! (2-z_0^2-4\theta _\infty +2\kappa_0)/3\! \end{array} \right) T + \left( \begin{array}{@{\,}c@{\,}} A_3 \\ B_3 \end{array} \right) T^2 + \cdots . \end{eqnarray*} $A_3$ is an arbitrary constant and $B_3$ is a certain constant depending on $A_3$. Let us consider a general system (\ref{2-1b}) satisfying the assumptions given in Sec.2.1; the Newton diagram consists of one compact face that lies on the plane $p_1 x + p_2 y + rz = s$, and $g_i$ satisfies $g_i(\lambda ^{p_1}x , \lambda ^{p_2}y , \lambda ^{r}z) \sim o(\lambda ^{s-r+p_i})$. In this case, the system has a formal series solution of the form \begin{equation} \left\{ \begin{array}{l} \displaystyle x(z) = \sum^\infty_{n=0} A_n(z-z_0)^{-p_1 + n}, \\ \displaystyle y(z) = \sum^\infty_{n=0}B_n(z-z_0)^{-p_2 + n}. \end{array} \right. \end{equation} The coefficients $A_n$ and $B_n$ are determined by substituting the series into the equation. If the series solution represents a general solution, it includes an arbitrary parameter other than $z_0$. The Kovalevskaya exponent $\kappa$ is defined to be the least integer $n$ such that the coefficient $(A_n, B_n)$ includes an arbitrary parameter. For the Laurent series solution of ($\text{P}_\text{I}$), $\kappa = 6$. For ($\text{P}_\text{II}$), $\kappa = 4$ for both series, and for ($\text{P}_\text{IV}$), $\kappa = 3$ for all Laurent series solutions. Note that the Kovalevskaya exponents of them coincide with the weighted degrees of Hamiltonian functions given in Sec.2.1. In Sec.5, it is shown that the Kovalevskaya exponent coincides with an eigenvalue of a Jacobi matrix of a certain vector field, and the exponent is invariant under the automorphism of $\C P^3(p,q,r,s)$. See \cite{Abl, Bor, Chi4, HuYa, Yos1, Yos2} for more general definition and properties of the Kovalevskaya exponent. \subsection{Relation with dynamical systems theory} In this section, a relationship between the Painlev\'{e} equations and the normal form theory of dynamical systems is shown. The Painlev\'{e} equations will be obtained from certain singular perturbed problems of vector fields. Let us consider a singular perturbation problem of the form \begin{equation} \left\{ \begin{array}{l} \dot{\bm{x}} = \bm{f}(\bm{x}, \bm{z}, \varepsilon ), \\ \dot{\bm{z}} = \varepsilon \bm{g}(\bm{x}, \bm{z}, \varepsilon ), \\ \end{array} \right. \label{2-22} \end{equation} where $\bm{x}\in \R^{m}, \bm{z}\in \R^n$, and $(\bm{f}, \bm{g})$ is a smooth vector field on $\R^{m+n}$ parameterized by a small parameter $\varepsilon \in \R$. The dot $(\,\dot{\,\,}\,)$ denotes the derivative with respect to time $t\in \R$. Such a system is called a fast-slow system because it is characterized by two different time scales; fast motion $\bm{x}$ and slow motion $\bm{z}$. This structure yields nonlinear phenomena such as a relaxation oscillation, which is observed in many physical, chemical and biological problems. See Grasman~\cite{Gra}, Hoppensteadt and Izhikevich~\cite{Hop} and references therein for applications of fast-slow systems. The unperturbed system is defined by putting $\varepsilon =0$ as \begin{equation} \dot{\bm{x}} = \bm{f}(\bm{x}, \bm{z}, 0 ), \quad \dot{\bm{z}} = 0. \end{equation} Since $\bm{z}$ is a constant for the unperturbed system, it is regarded as a parameter of the fast system $\dot{\bm{x}} = \bm{f}(\bm{x}, \bm{z}, 0 )$. It is known that when $\bm{f} \sim O(1)$ as $\varepsilon \to 0$ in some region of $\R^{m+n}$, the dynamics of (\ref{2-22}) is approximately governed by the first system $\dot{\bm{x}} = \bm{f}(\bm{x}, \bm{z}, 0 )$, and when $\bm{f} \sim 0$ while $D\bm{f} \sim O(1)$, the dynamics of (\ref{2-22}) is approximately governed by the slow system $\dot{\bm{z}} = \varepsilon \bm{g}(\varphi(\bm{z}), \bm{z}, 0 )$, where $D\bm{f}$ is the derivative of $\bm{f}$ with respect to $\bm{x}$ and $\varphi$ is a function satisfying $\bm{f}(\varphi(\bm{z}), \bm{z}, 0 ) = 0$. However, if both of $\bm{f}$ and $D\bm{f}$ are nearly equal to zero, both of the fast and slow motion should be taken into account and a nontrivial dynamics may occur. The condition \begin{eqnarray*} \bm{f}(\bm{x}_0, \bm{z}_0, 0 )=0, \quad D\bm{f}(\bm{x}_0, \bm{z}_0, 0 )=0 \end{eqnarray*} implies that the first system $\dot{\bm{x}} = \bm{f}(\bm{x}, \bm{z}, 0 )$ undergoes a bifurcation at $\bm{x}=\bm{x}_0$ with a bifurcation parameter $\bm{z}=\bm{z}_0$. A type of a bifurcation almost determines the local dynamics of (\ref{2-22}) around $(\bm{x}, \bm{z})=(\bm{x}_0, \bm{z}_0 )$ For the most generic case, in which the fast system undergoes a saddle-node bifurcation, it is well known that a local behavior of (\ref{2-22}) is governed by the Airy equation $d^2u/dz^2 = zu$. In particular, the asymptotic analysis of the Airy function plays an important role, see \cite{Kru1, Gil}. Chiba \cite{Chi2} found that when the fast system $\dot{\bm{x}} = \bm{f}(\bm{x}, \bm{z}, 0 )$ undergoes a Bogdanov-Takens bifurcation, then a local behavior of (\ref{2-22}) is determined by the asymptotic analysis of Boutroux's tritronqu\'{e}e solution of the first Painlev\'{e} equation. This result was applied to prove the existence of a periodic orbit and chaos in a certain biological system \cite{Chi3}. Here, we will demonstrate how the first, second and fourth Painlev\'{e} equations are obtained from fast-slow systems. For a normal form and versal deformation of germs of vector fields, the readers can refer to \cite{Cho}. Suppose that a one-dimensional dynamical system $\dot{x} = f(x)$ lies on a codimension 1 bifurcation at $x=0$. This means that $f$ satisfies \begin{equation} f(0) = f'(0) = 0, \quad f''(0) \neq 0. \end{equation} The normal form is given by $f(x) = x^2$, and its versal deformation is $\dot{x} = x^2 + z$ or $\dot{x} = x^2 + zx$ with a deformation parameter $z\in \R$. The former is called a saddle-node bifurcation and the latter is a transcritical bifurcation. The fast-slow system having the saddle-node as an unperturbed fast system is written by \begin{equation} \dot{x} = x^2 + z + \varepsilon f(x,z,\varepsilon ), \quad \dot{z} = \varepsilon g(x,z,\varepsilon ). \label{2-25} \end{equation} We also assume the generic condition $g(0,0,0) \neq 0$ so that we can write $g(x,z,\varepsilon ) = 1 + O(x,z,\varepsilon )$ without loss of generality. In order to investigate the local behavior of the system near $(x,z) = (0,0)$ for a small $\varepsilon $, we rewrite Eq.(\ref{2-25}) as a three dimensional system \begin{equation} \left\{ \begin{array}{l} \dot{x} = x^2 + z + \varepsilon f(x,z,\varepsilon ) \\ \dot{z} = \varepsilon + \varepsilon \cdot O(x,z,\varepsilon ) \\ \dot{\varepsilon } = 0, \end{array} \right. \label{2-26} \end{equation} by adding the trivial equation $\dot{\varepsilon } = 0$. For this system, we perform the weighted blow-up at the origin defined by \begin{equation} \left( \begin{array}{@{\,}c@{\,}} x \\ z \\ \varepsilon \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_1 \\ r_1^2 Z_1 \\ r_1^3 \varepsilon _1 \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_2 X_2 \\ r_2^2 \\ r_2^3 \varepsilon _2 \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_3 X_3 \\ r_3^2 Z_3 \\ r_3^3 \end{array} \right). \end{equation} The weight (exponents of $r_i$'s) $(1,2,3)$ can be found by the Newton diagram of Eq.(\ref{2-26}) as in Sec.2.1. The exceptional divisor of the blow-up is $\C P^2 (1,2,3)$ given by the set $\{ r_1 = 0\} \cup \{ r_2 = 0\} \cup \{ r_3 = 0\}$. On the $(X_3, Z_3, \varepsilon _3)$ chart, Eq.(\ref{2-26}) is written as \begin{equation} \dot{X}_3 = r_3X_3^2 + r_3Z_3 + O(r_3^2), \quad \dot{Z}_3 = r_3 + O(r_3^2), \quad \dot{r}_3 = 0. \end{equation} This is equivalent to \begin{eqnarray*} \frac{dX_3}{dZ_3} = \frac{X_3^2 + Z_3 + O(r_3)}{1 + O(r_3)}. \end{eqnarray*} In particular, on the exceptional divisor $\C P^2 (1,2,3)$, it is reduced to the Riccati equation \begin{equation} \frac{dX_3}{dZ_3} = X_3^2 + Z_3, \end{equation} which is equivalent to the Airy equation $u'' = -Z_3u$ by $X_3 = -u'/u$. Similarly, if we use the transcritical bifurcation as the fast system and apply the weighted blow-up with the weight $(1,1,2)$, we obtain the Hermite equation $u'' - Z_3u' - \alpha u = 0$ on the exceptional divisor. See \cite{Chi} for the analysis of the Airy equation based on the geometry of $\C P^2 (1,2,3)$. The Painlev\'{e} equations are obtained from codimension 2 bifurcations in a similar manner. Suppose that a 2-dim system of $(x,y)$ undergoes a generic codimension 2 bifurcation called the Bogdanov-Takens bifurcation. The normal form is given by $\dot{x} = y^2 + xy,\, \dot{y} = x$; that is, the linear part has two zero eigenvalues. Its versal deformation is \begin{equation} \left\{ \begin{array}{l} \dot{x} = y^2 + xy + z, \\ \dot{y} = x, \\ \end{array} \right. \end{equation} where $z$ is a deformation parameter. The fast-slow system having it as an unperturbed fast system is written by \begin{equation} \left\{ \begin{array}{l} \dot{x} = y^2 + xy + z + \varepsilon f_1(x,y,z,\varepsilon ) \\ \dot{y} = x+ \varepsilon f_2(x,y,z,\varepsilon ), \\ \dot{z} = \varepsilon + \varepsilon \cdot O(x,y,z,\varepsilon ), \\ \dot{\varepsilon } = 0, \end{array} \right. \label{2-31} \end{equation} where the trivial equation $\dot{\varepsilon } = 0$ is added as before. For this system, we introduce the weighted blow-up with the weight $(3,2,4,5)$ defined by \begin{equation} \left( \begin{array}{@{\,}c@{\,}} x \\ y \\ z \\ \varepsilon \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_1^3 \\ r_1^2 Y_1 \\ r_1^4 Z_1\\ r_1^5 \varepsilon _1 \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_2^3 X_2\\ r_2^2 \\ r_2^4 Z_2\\ r_2^5 \varepsilon _2 \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_3^3 X_3 \\ r_3^2 Y_3 \\ r_3^4 \\ r_3^5 \varepsilon _3 \end{array} \right) = \left( \begin{array}{@{\,}l@{\,}} r_4^3 X_4 \\ r_4^2 Y_4 \\ r_4^4 Z_4\\ r_4^5 \end{array} \right). \end{equation} The weight $(3,2,4,5)$ can be obtained via the Newton diagram of the system (\ref{2-31}). The exceptional divisor of the blow-up is the weighted projective space $\C P^3(3,2,4,5)$ given as the set $\{ r_1 =0 \} \cup \{ r_2 =0 \} \cup\{ r_3 =0 \} \cup\{ r_4 =0 \}$. Transforming the system (\ref{2-31}) to the $(X_4, Y_4, Z_4, r_4)$ chart, we obtain \begin{eqnarray*} \left\{ \begin{array}{l} \dot{X}_4 = r_4Y_4^2+r_4Z_4+ O(r_4^2), \\ \dot{Y}_4 = r_4X_4 + O(r_4^2), \\ \dot{Z}_4 = r_4 + O(r_4^2). \end{array} \right. \end{eqnarray*} As $r_4 \to 0$, it is reduced to the first Painlev\'{e} equation $dX_4/dZ_4 = Y_4^2 + Z_4,\, dY_4/dZ_4 = X_4$. This implies that the dynamics on the divisor $\C P^3(3,2,4,5)$ is governed by the compactified first Painlev\'{e} equation, and a local behavior of the system (\ref{2-31}) near $(x,y,z) = (0,0,0)$ can be investigated by a global analysis of the first Painlev\'{e} equation. Next, we consider a 2-dim system that undergoes a Bogdanov-Takens bifurcation with $\Z_2$-symmetry $(x,y) \mapsto (-x,-y)$. The versal deformation of the normal form is given by \begin{equation} \left\{ \begin{array}{l} \dot{x} = y^3 - xy^3 + zy, \\ \dot{y} = x, \\ \end{array} \right. \end{equation} with a deformation parameter $z$. The fast-slow system having it as an unperturbed fast system is written by \begin{equation} \left\{ \begin{array}{l} \dot{x} = y^3 - xy^3 + zy + \varepsilon f_1(x,y,z,\varepsilon ) \\ \dot{y} = x+ \varepsilon f_2(x,y,z,\varepsilon ), \\ \dot{z} = \varepsilon + \varepsilon \cdot O(x,y,z,\varepsilon ), \\ \dot{\varepsilon } = 0. \end{array} \right. \label{2-34} \end{equation} For this system, we introduce the weighted blow-up with the weight $(2,1,2,3)$, which is found by the Newton diagram of the system (\ref{2-34}). On the $(X_4, Y_4, Z_4, r_4)$ chart, it provides \begin{eqnarray*} \left\{ \begin{array}{l} \dot{X}_4 = r_4Y_4^3 + r_4Z_4Y_4 - r_4^3 X_4Y_4^3 + r_4 f_1(r_4^2X_4, r_4Y_4, r_4^2Z_4, r_4^3), \\ \dot{Y}_4 = r_4X_4 + O(r_4^2), \\ \dot{Z}_4 = r_4 + O(r_4^2). \end{array} \right. \end{eqnarray*} Note that $r_4 f_1(r_4^2X_4, r_4Y_4, r_4^2Z_4, r_4^3) = \alpha r_4 + O(r_4^2)$, where $\alpha := f_1(0,0,0,0)$. As $r_4 \to 0$, this system is reduced to the second Painlev\'{e} equation $X_4' = Y_4^3 + Y_4Z_4 + \alpha ,\, Y_4' = X_4$ with a parameter $\alpha $. Finally, we consider a 2-dim system that undergoes a Bogdanov-Takens bifurcation with $\Z_3$-symmetry. Using the complex variable $\eta \in \C$, the normal form of such a bifurcation is given by $\dot{\eta} = \eta |\eta |^2 + \overline{\eta}^2$. Note that this system is invariant under the $\Z_3$ action $\eta \mapsto e^{2\pi i/ 3} \eta$. The versal deformation of the normal form is \begin{equation} \dot{\eta} = \eta |\eta |^2 + \overline{\eta}^2 +\eta z. \end{equation} where $z\in \C$ is a deformation parameter. Putting $\eta = x + iy, z = z_1 + iz_2$ yields \begin{eqnarray*} \left\{ \begin{array}{l} \dot{x} = x^2 - y^2 + z_1x - z_2y + O(\eta^3), \\ \dot{y} = -2xy + z_1y + z_2 x + O(\eta^3). \\ \end{array} \right. \end{eqnarray*} We assume $z_1 = 0$ so that the above system may become a Hamiltonian system, and change the notation $z_2 \mapsto z$ to obtain \begin{eqnarray*} \left\{ \begin{array}{l} \dot{x} = x^2 - y^2 - zy + O(\eta^3), \\ \dot{y} = -2xy + z x + O(\eta^3). \\ \end{array} \right. \end{eqnarray*} The fast-slow system having it as an unperturbed fast system is written by \begin{equation} \left\{ \begin{array}{l} \dot{x} = x^2 - y^2 - zy + O(\eta^3) + \varepsilon f_1(x,y,z,\varepsilon ) \\ \dot{y} = -2xy + z x + O(\eta^3) + \varepsilon f_2(x,y,z,\varepsilon ), \\ \dot{z} = \varepsilon + \varepsilon \cdot O(x,y,z,\varepsilon ), \\ \dot{\varepsilon } = 0. \end{array} \right. \label{2-36} \end{equation} For this system, we introduce the weighted blow-up with the weight $(1,1,1,2)$. Moving to the $(X_4,Y_4,Z_4)$ chart and putting $r_4 = 0$ as before, it turns out that the system (\ref{2-36}) is reduced the system \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{dX_4}{dZ_4} = X_4^2 - Y_4^2 - Z_4Y_4 + \alpha _1 \\ \displaystyle \frac{dY_4}{dZ_4} = -2X_4Y_4 + Z_4Y_4 + \alpha _2, \\ \end{array} \right. \label{2-37} \end{equation} where $\alpha_i := f_i(0,0,0,0)$ is a constant ($i=1,2$). This is equivalent to the fourth Painlev\'{e} equation (\ref{P4}) through an affine transformation of $(X_4, Y_4)$. \begin{table}[h] \begin{center} \begin{tabular}{|c|c|c|} \hline Bifurcation type & Exceptional divisor & Equation on the divisor \\ \hline \hline saddle-node & $\C P^2(1,2,3)$ & Airy \\ \hline transcritical & $\C P^2(1,1,2)$ & Hermite \\ \hline Bogdanov-Takens (BT) & $\C P^3(3,2,4,5)$ & ($\text{P}_\text{I}$) \\ \hline BT with $\Z_2$ symmetry & $\C P^3(2,1,2,3)$ &($\text{P}_\text{II}$) \\ \hline BT with $\Z_3$ symmetry & $\C P^3(1,1,1,2)$ & ($\text{P}_\text{IV}$) \\ \hline \end{tabular} \end{center} \caption{Differential equations obtained by the weighted blow-up of the fast-slow systems.} \end{table} The results are summarized in Table 1. It is remarkable that the weights $(p,q,r,s)$ derived in Sec.2.1 via the Newton diagrams are determined only by the versal deformations of the codimension 2 bifurcations. Further, the Painlev\'{e} equations are obtained in a compactified manner on $\C P^3(p,q,r,s)$, which is the exceptional divisor of the weighted blow-up. \section{Singular normal forms and the Painlev\'{e} property} Recall that a singularity $z = z_*$ of a solution of a differential equation is called movable if the position of $z_*$ depends on the choice of an initial condition. In this section, we give local analysis near such movable singularities based on the normal form theory. Our purpose is to show that near movable singularities, the Painlev\'{e} equations are locally transformed into integrable systems called the singular normal form. Further, we will give a new proof of the Painlev\'{e} property; any solutions of ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are meromorphic on $\C$ (for this purpose, we do not use the Laurent series given in Sec.2.3). \subsection{The first Painlev\'{e} equation} ($\text{P}_\text{I}$) is given on the weighted projective space $\C P^3(3,2,4,5)$ as a tuple of equations (\ref{P1}), (\ref{P11}), (\ref{P12}) and (\ref{P13}). Coordinate transformations between inhomogeneous coordinates are given by \begin{eqnarray} \left\{ \begin{array}{rrrr} x = & \varepsilon _1^{-3/5} = & X_2\varepsilon _2^{-3/5} =& X_3\varepsilon _3^{-3/5} \\ y = & Y_1\varepsilon _1^{-2/5} = & \varepsilon _2^{-2/5}=& Y_3\varepsilon _3^{-2/5} \\ z = & Z_1\varepsilon_1 ^{-4/5} = & Z_2\varepsilon _2^{-4/5}=& \varepsilon _3^{-4/5} . \end{array} \right. \label{3-1} \end{eqnarray} Due to the orbifold structure of $\C P^3(3,2,4,5)$, local charts $(Y_1, Z_1, \varepsilon _1)$, $(X_2, Z_2,\varepsilon _2)$ and $(X_3, Y_3, \varepsilon _3)$ should be divided by the $\Z_3, \Z_2$ and $\Z_4$ actions, respectively, defined by \begin{eqnarray} & & (Y_1, Z_1, \varepsilon _1) \mapsto (\omega Y_1, \omega ^2 Z_1, \omega \varepsilon _1), \quad \omega := e^{2\pi i/3}, \label{3-2} \\ & & (X_2, Z_2,\varepsilon _2) \mapsto (-X_2, Z_2, -\varepsilon _2), \label{3-3} \\ & & (X_3, Y_3, \varepsilon _3) \mapsto (i X_3, -Y_3, -i \varepsilon _3). \label{3-4} \end{eqnarray} Indeed, Eqs.(\ref{P11}),(\ref{P12}),(\ref{P13}) are invariant under these actions. For our purposes, it is convenient to rewrite Eqs.(\ref{P11}), (\ref{P12}) and (\ref{P13}) as 3-dim vector fields (autonomous ODEs) given by \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{Y}_1 = 3 - 12Y_1^3 - 2Y_1Z_1, \\ \displaystyle \dot{Z}_1 = 3\varepsilon _1 - 24Y_1^2 Z_1 - 4 Z_1^2, \\ \displaystyle \dot{\varepsilon }_1 = \varepsilon _1 (-30 Y_1^2 - 5Z_1), \end{array} \right. \label{3-5} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_2 = (-12 - 2Z_2 + 3X_2^2)/X_2, \\ \displaystyle \dot{Z}_2 = (-2\varepsilon _2 + 4 X_2Z_2)/X_2, \\ \displaystyle \dot{\varepsilon }_2 = 5\varepsilon _2, \end{array} \right. \label{3-6} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_3 =24Y_3^2 + 4 - 3X_3\varepsilon _3, \\ \displaystyle \dot{Y}_3 = 4X_3 - 2Y_3\varepsilon _3, \\ \displaystyle \dot{\varepsilon }_3 = -5\varepsilon _3^2, \end{array} \right. \label{3-7} \end{equation} where $(\dot{\,\,}) = d/dt$ and $t$ is an additional parameter. Recall the decomposition (\ref{2-7}) with (\ref{2-8}). According to Eqs.(\ref{3-5}),(\ref{3-6}),(\ref{3-7}), the set $\C P^2 (3,2,4)$ is an invariant manifold of the vector fields; that is, $\varepsilon _i(t) \equiv 0$ when $\varepsilon _i (0) = 0$ at an initial time. The dynamics on the invariant manifold describes the behavior of ($\text{P}_\text{I}$) near infinity. In particular, fixed points of the vector fields play an important role. Vector fields (\ref{3-5}),(\ref{3-6}),(\ref{3-7}) have exactly two fixed points on $\C P^2 (3,2,4)$; \\[0.2cm] \textbf{(i).} $(X_2, Z_2, \varepsilon _2) = (\pm 2,0,0)$. Due to the $\Z_2$ action on the $(X_2,Z_2, \varepsilon _2)$-coordinates, two points $(2,0,0)$ and $(-2,0,0)$ represent the same point on $\C P^3(3,2,4,5)$ and it is sufficient to consider one of them. We will show that this fixed point corresponds to movable singularities of ($\text{P}_\text{I}$). By applying the normal form theory of vector fields to this point, we will construct the singular normal form of ($\text{P}_\text{I}$). In Sec.4, the space of initial conditions for ($\text{P}_\text{I}$) is constructed by applying the weighted blow-up at this point. \\[0.2cm] \textbf{(ii).} $(X_3, Y_3, \varepsilon _3) = (0, (-6)^{-1/2} ,0)$ Again, two points should be identified due to the $\Z_4$ action on $(X_3, Y_3, \varepsilon _3)$. This fixed point corresponds to the irregular singular point of ($\text{P}_\text{I}$) because $\varepsilon _3 = 0$ provides $z = \infty$. Note that fixed points obtained from the $(Y_1, Z_1, \varepsilon _1)$-coordinates are the same as one of the above. For example, the fixed point $(Y_1, Z_1, \varepsilon _1) = ((1/4)^{1/3},0,0)$ is the same as $(X_2, Z_2, \varepsilon _2) = (\pm 2,0,0)$ due to (\ref{3-1}). At first, let us show that ($\text{P}_\text{I}$) is locally transformed into a linear system around $(X_2,Z_2,\varepsilon _2) = (2,0,0)$. By putting $\hat{X}_2 = X_2 - 2$, Eq.(\ref{3-6}) is rewritten as \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{\hat{X}}_2 = 6\hat{X}_2 - Z_2 + \frac{-3\hat{X}_2^2 + \hat{X}_2Z_2}{2+\hat{X}_2}, \\ \displaystyle \dot{Z}_2 = 4Z_2 - \varepsilon _2 + \frac{\hat{X}_2\varepsilon _2}{2+\hat{X}_2}, \\ \displaystyle \dot{\varepsilon }_2 = 5\varepsilon _2. \end{array} \right. \label{3-8} \end{equation} The origin is a fixed point with the Jacobi matrix \begin{eqnarray} J = \left( \begin{array}{@{\,}ccc@{\,}} 6 & - 1 & 0 \\ 0 & 4 & - 1 \\ 0 & 0 & 5 \end{array} \right). \label{3-9} \end{eqnarray} The eigenvalues $\lambda =6,4,5$ satisfy the conditions on the Poincar\'{e} linearization theorem (for the convenience of the reader, we give the statement of this theorem in the end of this subsection). Hence, there exists a neighborhood $U$ of the origin and a local analytic transformation defined on $U$ of the form \begin{eqnarray*} \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 \\ Z_2 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} \hat{u}_2 \\ v_2 \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 + \varphi _1(\hat{X}_2,Z_2,\varepsilon _2) \\ Z_2 + \hat{\varphi} _2 (\hat{X}_2, Z_2, \varepsilon _2) \end{array} \right), \end{eqnarray*} such that Eq.(\ref{3-8}) is linearized as \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{\hat{u}}_2 = 6\hat{u}_2 - v_2, \\ \displaystyle \dot{v}_2 = 4v_2 - \varepsilon _2, \\ \displaystyle \dot{\varepsilon }_2 = 5\varepsilon _2, \end{array} \right. \label{3-10} \end{equation} where local analytic functions $\varphi_1$ and $\hat{\varphi}_2$ satisfy $\varphi_1, \hat{\varphi}_2 \sim O(|| \bm{X} ||^2)$, $\bm{X}:= (\hat{X}_2, Z_2, \varepsilon _2)$. Note that we need not change $\varepsilon _2$ because the equation of $\varepsilon _2$ is already linear. Furthermore, we have $\hat{\varphi}_2 (\hat{X}_2, Z_2, 0) = 0$ because the equation of $Z_2$ is linear when $\varepsilon _2 =0$. Thus, we can set $\hat{\varphi}_2 = \varepsilon _2 \varphi_2$ and the above transformation takes the form \begin{eqnarray} \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 \\ Z_2 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} \hat{u}_2 \\ v_2 \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 + \varphi _1(\hat{X}_2,Z_2,\varepsilon _2) \\ Z_2 + \varepsilon _2 \varphi _2 (\hat{X}_2, Z_2, \varepsilon _2) \end{array} \right), \label{3-11} \end{eqnarray} with $\varphi _1 \sim O(|| \bm{X} ||^2)$ and $\varphi _2 \sim O(X_2, Z_2, \varepsilon _2)$. For the linear system Eq.(\ref{3-10}), let us change coordinates as $\hat{u}_2 = u_2 - 2$, and $\tilde{x} = u_2\varepsilon _2^{-3/5}, \tilde{y} = \varepsilon _2^{-2/5}, \tilde{z} = v_2 \varepsilon _2^{-4/5}$; that is, we move to the original chart for ($\text{P}_\text{I}$). We can verify that $(\tilde{x}, \tilde{y}, \tilde{z})$ satisfies the equation $d^2 \tilde{y}/d\tilde{z}^2 = 6\tilde{y}^2$, whose solution can be expressed by the Weierstrass's elliptic function. The relation between $(x,y,z)$ and $(\tilde{x}, \tilde{y}, \tilde{z})$ is \begin{eqnarray*} \left( \begin{array}{@{\,}c@{\,}} \tilde{x} \\ \tilde{y} \\ \tilde{z} \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} x + y^{3/2} \varphi_1 (xy^{-3/2}-2, zy^{-2}, y^{-5/2}) \\ y \\ z + y^{-1/2}\varphi_2 (xy^{-3/2}-2, zy^{-2}, y^{-5/2}) \end{array} \right). \end{eqnarray*} In particular, $\tilde{y} = y$ is not changed. Now we have obtained \\[0.2cm] \textbf{Proposition \thedef.} There is a local analytic transformation $(x,z) \mapsto (\tilde{x},\tilde{z})$ defined near $(\hat{X}_2, Z_2, \varepsilon _2) = (0,0,0)$ such that ($\text{P}_\text{I}$) $y'' = 6y^2 + z$ is transformed into the integrable system $y'' = 6y^2$. This fact was first obtained by \cite{CosCos} for ($\text{P}_\text{I}$). Our proof using the weighted projective space and the Poincar\'{e} theorem is also applicable to the second Painlev\'{e} to sixth Painlev\'{e} equations to prove that they are locally transformed to solvable systems. Since Eq.(\ref{3-10}) is linear, we can construct two integrals explicitly as \begin{eqnarray*} C_1 = \varepsilon _2^{-4/5}v_2 + \varepsilon _2^{1/5}, \quad C_2 = \frac{1}{2}\varepsilon _2^{-1/5} + \varepsilon _2^{-6/5}\hat{u}_2 - \frac{1}{2}\varepsilon _2^{-6/5}v_2. \end{eqnarray*} By applying the transformations (\ref{3-11}), $\hat{X}_2 = X_2 - 2$ and (\ref{3-1}), we obtain the local integrals of ($\text{P}_\text{I}$) of the form \begin{eqnarray} \left\{ \begin{array}{ll} \displaystyle C_1 = z+ y^{-1/2}+y^{-1/2} \varphi _2(xy^{-3/2}-2, zy^{-2}, y^{-5/2}), \\[0.2cm] \displaystyle C_2 = \frac{1}{2}y^{1/2}-2y^3 + xy^{3/2} - \frac{1}{2}yz + y^3 \varphi _1(\cdots ) - \frac{1}{2}y^{1/2} \varphi _2(\cdots ), \\ \end{array} \right. \label{3-12} \end{eqnarray} Arguments of $\varphi _1$ and $\varphi _2$ in the second line are the same as that of the first line. Now we give a new proof of the well known theorem: \\[0.2cm] \textbf{Theorem.\thedef.} Any solutions of ($\text{P}_\text{I}$) are meromorphic on $\C$. \\ A well known proof of this result is essentially based on Painlev\'{e}'s argument modified by Hukuhara (\cite{OkaTak}, \cite{Hin}, see also \cite{Gro}). Here, we will prove the theorem by applying the implicit function theorem to the above integrals. \\[0.2cm] \textbf{Proof.} Fix a solution $(x(z), y(z))$ of ($\text{P}_\text{I}$) with an initial condition $(x(z_0), y(z_0)) = (x_0, y_0)$. The existence theorem of solutions shows that the solution is holomorphic near $z_0$. Let $B(z_0, R)$ be the largest disk of radius $R$ centered at $z_0$ such that the solution is holomorphic inside the disk. Let $z_* \, (\neq \infty)$ be a singularity on the boundary of the disk (if $R=\infty$, there remains nothing to prove). The next lemma implies that the fixed point $(X_2, Z_2, \varepsilon _2) = (\pm 2,0,0)$ corresponds to the singularity $z_*$. \\[0.2cm] \textbf{Lemma.\thedef.} $(X_2, Z_2, \varepsilon _2) \to (\pm 2,0,0)$ as $z \to z_*$ along a curve $\Gamma $ inside the disk $B(z_0, R)$. \\[0.2cm] \textbf{Proof.} Suppose that there exists a sequence $\{z_n\}_{n=1}^\infty$ converging to $z_*$ on the curve $\Gamma $ such that both of $x(z_n)$ and $y(z_n)$ are bounded as $n\to \infty$. Taking a subsequence if necessary, we can assume that $(x, y)$ converges to some point $(x_*, y_*)$. Because of the existence theorem of solutions, a solution of ($\text{P}_\text{I}$) satisfying the initial condition $(x_*, y_*, z_*)$ is holomorphic around this point, which contradicts with the definition of $z_*$. Hence, either $x$ or $y$ diverges as $z \to z_*$. (i) Suppose that $y \to \infty$ as $z \to z_*$. We move to the $(X_2, Z_2, \varepsilon _2)$-coordinates. Eq.(\ref{3-1}) provides \begin{equation} X_2 = xy^{-3/2}, \quad Z_2 = zy^{-2}, \quad \varepsilon _2 = y^{-5/2}. \end{equation} This immediately yields $Z_2 \to 0,\, \varepsilon _2 \to 0$ as $z \to z_*$. Let us show $X_2 \to \pm 2$. ($\text{P}_\text{I}$) is a Hamiltonian system with the Hamiltonian function $H = x^2/2 - 2y^3 - zy$. Thus, the equality $H = -\int\! y(z)dz $ holds along a solution. In the $(X_2, Z_2, \varepsilon _2)$-coordinates, this is written as \begin{eqnarray*} X_2(Z_2)^2 = 4 + \frac{2}{3}Z_2 - \frac{2}{3}\varepsilon _2(Z_2)^{6/5} \int^{Z_2}_{\xi}\! \varepsilon _2(z)^{-6/5}dz, \end{eqnarray*} where $(X_2(Z_2), \varepsilon _2(Z_2))$ is a solution of the ODE solved as a function of $Z_2$, and $\xi$ is a certain nonzero number determined by the initial condition. Since $Z_2 \to 0,\, \varepsilon _2 \to 0$ as $z \to z_*$, we obtain $X^2_2 \to 4$. (ii) Suppose that $x \to \infty$ as $z \to z_*$. In this case, we use the $(Y_1, Z_1, \varepsilon _1)$-coordinates given by \begin{equation} Y_1 = yx^{-2/3}, \quad Z_1 = zx^{-4/3}, \quad \varepsilon _1 = x^{-5/3}. \end{equation} By the assumption, we have $Z_1 \to 0,\, \varepsilon _1 \to 0$ as $z \to z_*$. Then, we can show that $Y_1^3 \to 1/4$ as $z \to z_*$ by the same way as above. This means that $(Y_1, Z_1, \varepsilon _1)$ converges to the fixed point $((1/4)^{1/3},0,0)$ of the vector field (\ref{3-5}). It is easy to verify that this fixed point is the same point as $(X_2, Z_2, \varepsilon _2) = (\pm 2,0,0)$ if written in the $(X_2, Z_2, \varepsilon _2)$-coordinates. $\Box$ The sign of $X_2 = xy^{-3/2}$ depends on the choice of the branch of $y^{1/2}$ and two points $(2,0,0)$ and $(-2,0,0)$ are the essentially the same. In what follows, we assume that $(X_2, Z_2, \varepsilon _2) \to (2,0,0)$ as $z \to z_*$. Due to the above lemma, when $z$ is sufficiently close to $z_*$, the solution is included in the neighborhood $U$, on which local holomorphic functions $\varphi _1$ and $\varphi _2$ are well defined. Then, the integrals (\ref{3-12}) are available. To apply the implicit function theorem, put \begin{eqnarray} \left( \begin{array}{@{\,}c@{\,}} w \\ u \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} y^{-1/2} \\ \frac{1}{2}y^{1/2}-2y^3 + xy^{3/2} - \frac{1}{2}yz \end{array} \right), \end{eqnarray} Then, (\ref{3-12}) takes the form \begin{eqnarray*} \left\{ \begin{array}{l} \displaystyle C_1 = z + w + w \varphi _2(uw^6+\frac{1}{2}zw^4-\frac{1}{2}w^5, z w^4, w^5)\\ \displaystyle C_2 = u + w^{-6}\varphi _1 (\cdots ) - \frac{1}{2}w^{-1}\varphi _2 (\cdots ). \\ \end{array} \right. \end{eqnarray*} Note that $C_1 = z+ w + O(w^5) $ and $C_2 = u + O(w^2)$ as $w \to 0$. Since $w\to 0$ as $z \to z_*$, the constant $C_1 = z_*$ is just the position of the singularity. If we set \begin{eqnarray*} & & f_1(w,u, z) = z + w + w\varphi _2 (\cdots ) - z_* \\ & & f_2(w,u, z) = u + w^{-6}\varphi _1 (\cdots )- \frac{1}{2}w^{-1}\varphi _2 (\cdots ) -C_2, \end{eqnarray*} then $f_i(0,C_2,z_*) = 0$. The Jacobi matrix of $(f_1, f_2)$ with respect to $(w,u)$ at $(w,u,z) = (0, C_2, z_*)$ is the identity matrix. Hence, the implicit function theorem proves that there exists a local holomorphic function $g(z)$ such that $f_1 = f_2 = 0$ is solved as $w = g(z) \sim O(z - z_*)$. Since $y = w^{-2}$, $z = z_*$ is a pole of second order of $y$. This completes the proof of Thm.3.2. $\Box$ \\ The Poincar\'{e} linearization theorem used in this subsection is stated as follows: Let $Ax + f(x)$ be a holomorphic vector field on $\C^n$ with a fixed point $x=0$, where $A$ is an $n\times n$ constant matrix and $f(x) \sim O(|x|^2)$ is a nonlinearity. Let $\lambda _1 , \cdots , \lambda _n$ be eigenvalues of $A$. We consider the following two conditions: \\[0.2cm] \textbf{(Nonresonance)} There are no $j\in \{ 1, \cdots ,n\}$ and non-negative integers $m_1, \cdots ,m_n$ satisfying the resonant condition \begin{equation} m_1 \lambda _1 + \cdots + m_n \lambda _n = \lambda _j, \quad (m_1 + \cdots + m_n \geq 2). \end{equation} \textbf{(Poincar\'{e} domain)} The convex hull of $\{ \lambda _1, \cdots , \lambda _n\}$ in $\C$ does not include the origin. \\[0.2cm] Suppose that $A$ is diagonal and eigenvalues satisfy the above two conditions. Then, there exists a local analytic transformation $y = x + \varphi (x),\,\, \varphi (x) \sim O(|x|^2)$ defined near the origin such that the equation $dx/dt = Ax + f(x)$ is transformed into the linear system $dy/dt = Ay$. See \cite{Cho} for the proof. \subsection{The second Painlev\'{e} equation} ($\text{P}_\text{II}$) is given on the weighted projective space $\C P^3(2,1,2,3)$ as a tuple of equations (\ref{P2}), (\ref{P21}), (\ref{P22}) and (\ref{P23}). Coordinate transformations between inhomogeneous coordinates are given by \begin{eqnarray} \left\{ \begin{array}{rrrr} x = & \varepsilon _1^{-2/3} = & X_2\varepsilon _2^{-2/3} =& X_3\varepsilon _3^{-2/3} \\ y = & Y_1\varepsilon _1^{-1/3} = & \varepsilon _2^{-1/3}=& Y_3\varepsilon _3^{-1/3}\\ z = & Z_1\varepsilon_1 ^{-2/3} = & Z_2\varepsilon _2^{-2/3}=& \varepsilon _3^{-2/3}. \end{array} \right. \label{3-17} \end{eqnarray} Due to the orbifold structure of $\C P^3(2,1,2,3)$, local charts $(Y_1, Z_1, \varepsilon _1)$ and $(X_3, Y_3, \varepsilon _3)$ should be divided by the $\Z_2$ actions defined by \begin{eqnarray} & & (Y_1, Z_1, \varepsilon _1) \mapsto (-Y_1, Z_1, - \varepsilon _1), \\ & & (X_3, Y_3, \varepsilon _3) \mapsto (X_3, -Y_3, -\varepsilon _3). \label{3-19} \end{eqnarray} For our purposes, it is convenient to rewrite Eqs.(\ref{P21}), (\ref{P22}) and (\ref{P23}) as 3-dim vector fields (autonomous ODEs) given by \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{Y}_1 = -2 + Y_1(2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1), \\ \displaystyle \dot{Z}_1 = -2\varepsilon _1 + 2Z_1(2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1), \\ \displaystyle \dot{\varepsilon }_1 =3\varepsilon _1 (2Y_1^3 + Y_1Z_1 + \alpha \varepsilon _1), \end{array} \right. \label{3-20} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_2 = 2X_2 - (2+Z_2 + \alpha \varepsilon _2)/X_2, \\ \displaystyle \dot{Z}_2 = 2Z_2 - \varepsilon _2/X_2, \\ \displaystyle \dot{\varepsilon }_2 = 3\varepsilon _2, \end{array} \right. \label{3-21} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_3 =4Y_3^3 + 2Y_3 + 2\alpha \varepsilon _3 - 2X_3\varepsilon _3, \\ \displaystyle \dot{Y}_3 = 2X_3 - Y_3\varepsilon _3, \\ \displaystyle \dot{\varepsilon }_3 = -3\varepsilon _3^2. \end{array} \right. \label{3-22} \end{equation} The set $\C P^2 (2,1,2)$ expressed as $\{ \varepsilon _1 = 0\} \cup \{ \varepsilon _2 =0 \} \cup \{ \varepsilon _3 = 0\}$ is an invariant manifold of the vector fields. The dynamics on the invariant manifold describes the behavior of ($\text{P}_\text{II}$) near infinity. Vector fields (\ref{3-20}),(\ref{3-21}),(\ref{3-22}) have four fixed points on the ``infinity set" $\C P^2 (2,1,2)$; \\[0.2cm] \textbf{(I).} $(X_2, Z_2, \varepsilon _2) = (\pm 1,0,0)$. We will show later that these fixed points correspond to movable singularities of ($\text{P}_\text{II}$). \\[0.2cm] \textbf{(II).} $(X_2, Z_2, \varepsilon _2) = (0,-2,0)$ and $(X_3, Y_3, \varepsilon _3) = (0, 0 ,0)$. In this case, it is easy to see from (\ref{3-17}) that $z=\infty$. Thus, these fixed points correspond to the irregular singular point of ($\text{P}_\text{II}$). Note that other fixed points represent the same point as one of the above. For example, the fixed point $(Y_1, Z_1, \varepsilon _1) = (1^{1/4},0,0)$ is the same as $(X_2, Z_2, \varepsilon _2) = (\pm 1,0,0)$ due to the transformation (\ref{3-17}). At first, we show that ($\text{P}_\text{II}$) is locally transformed into a linear system. Putting $\hat{X}_2 = X_2 \pm 1$ for Eq.(\ref{3-21}) yields \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{\hat{X}}_2 = 4\hat{X}_2 \pm Z_2 \pm \alpha \varepsilon _2 - \frac{2\hat{X}_2^2\pm \hat{X}_2Z_2 \pm \alpha \hat{X}_2\varepsilon _2}{\hat{X}_2 \mp 1}, \\ \displaystyle \dot{Z}_2 = 2Z_2 \pm \varepsilon _2 \mp \frac{\hat{X}_2\varepsilon _2}{\hat{X}_2 \mp 1}, \\ \displaystyle \dot{\varepsilon }_2 = 3\varepsilon _2. \end{array} \right. \label{3-23} \end{equation} The origin is a fixed point with the Jacobi matrix \begin{eqnarray} J = \left( \begin{array}{@{\,}ccc@{\,}} 4 & \pm 1 & \pm \alpha \\ 0 & 2 & \pm 1 \\ 0 & 0 & 3 \end{array} \right). \label{3-24} \end{eqnarray} Now we apply the normal form theory \cite{Cho} to the fixed point. The eigenvalues $\lambda _1 = 4, \lambda _2 = 2, \lambda _3 = 3$ satisfy the resonance relation $2\lambda _2 = \lambda _1$. However, Eq.(\ref{3-23}) does not include the corresponding resonance term. Hence, Poincar\'{e}'s theorem on normal forms proves that there exists a neighborhood $U$ of the origin and a local analytic transformation defined on $U$ of the form \begin{eqnarray*} \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 \\ Z_2 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} \hat{u}_2 \\ v_2 \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 + \varphi _1(\hat{X}_2,Z_2,\varepsilon _2) \\ Z_2 + \hat{\varphi} _2 (\hat{X}_2, Z_2, \varepsilon _2) \end{array} \right), \end{eqnarray*} such that Eq.(\ref{3-23}) is linearized as \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{\hat{u}}_2 = 4\hat{u}_2 \pm v_2 \pm \alpha \varepsilon _2, \\ \displaystyle \dot{v}_2 = 2v_2 \pm \varepsilon _2, \\ \displaystyle \dot{\varepsilon }_2 = 3\varepsilon _2, \end{array} \right. \label{3-25} \end{equation} where local analytic functions $\varphi_1$ and $\hat{\varphi}_2$ satisfy $\varphi_1, \hat{\varphi}_2 \sim O(|| \bm{X} ||^2)$, $\bm{X}:= (\hat{X}_2, Z_2, \varepsilon _2)$. Note that we need not change $\varepsilon _2$ because the equation of $\varepsilon _2$ is already linear. Furthermore, we have $\hat{\varphi}_2 (\hat{X}_2, Z_2, 0) = 0$ because the equation of $Z_2$ is linear when $\varepsilon _2 =0$. Thus, we can set $\hat{\varphi}_2 = \varepsilon _2 \varphi_2$ and the above transformation takes the form \begin{eqnarray} \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 \\ Z_2 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} \hat{u}_2 \\ v_2 \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} \hat{X}_2 + \varphi _1(\hat{X}_2,Z_2,\varepsilon _2) \\ Z_2 + \varepsilon _2 \varphi _2 (\hat{X}_2, Z_2, \varepsilon _2) \end{array} \right), \label{3-26} \end{eqnarray} with $\varphi _1 \sim O(|| \bm{X} ||^2)$ and $\varphi _2 \sim O(X_2, Z_2, \varepsilon _2)$. Although $\varphi _1, \varphi _2$ and the neighborhood $U$ depend on the choice of the sign of $\hat{X}_2 = X_2 \pm 1$, we need not distinguish them in this subsection. For the linear system (\ref{3-25}), let us move to the original chart for ($\text{P}_\text{II}$); that is, change coordinates as $\hat{u}_2 = u_2 \pm 1$ and $\tilde{x} = u_2\varepsilon _2^{-2/3}, \tilde{y} = \varepsilon _2^{-1/3}, \tilde{z} = v_2 \varepsilon _2^{-2/3}$. Then, we obtain the solvable system \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \frac{d\tilde{x}}{d\tilde{z}} = \pm 2\tilde{x}\tilde{y} + 4\tilde{y}^3 + \tilde{z}\tilde{y} + \alpha, \\[0.2cm] \displaystyle \frac{d\tilde{y}}{d\tilde{z}} = \mp \tilde{y}^2. \\ \end{array} \right. \label{3-27} \end{eqnarray} The coordinate transformation is given by \begin{eqnarray*} \left( \begin{array}{@{\,}c@{\,}} \tilde{x} \\ \tilde{y} \\ \tilde{z} \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} x + y^2\varphi_1 (xy^{-2} \pm 1, zy^{-2}, y^{-3}) \\ y \\ z + y^{-1} \varphi_2 (xy^{-2} \pm 1, zy^{-2}, y^{-3}) \end{array} \right). \end{eqnarray*} Hence, we have obtained the next proposition. \\[0.2cm] \textbf{Proposition.\thedef.} There exists a local analytic transformation $(x, z) \mapsto (\tilde{x},\tilde{z})$ defined near $(X_2, Z_2, \varepsilon _2) = (\pm 1, 0,0)$ such that ($\text{P}_\text{II}$) is transformed into the solvable system (\ref{3-27}). For ($\text{P}_\text{II}$) $d^2y/dz^2 = 2y^3 + yz + \alpha $, if $y$ becomes sufficiently large for a finite $z$, we expect that the equation is well approximated by $d^2y/dz^2 \sim 2y^3$ . Indeed, the second equation of (\ref{3-27}) provides $d^2\tilde{y}/d\tilde{z}^2 = 2\tilde{y}^3$. \\ Since Eq.(\ref{3-25}) is linear, we can construct two integrals explicitly as \begin{eqnarray*} C_1 = \varepsilon _2^{-2/3}v_2 \mp \varepsilon _2^{1/3}, \quad C_2 =\varepsilon _2^{-4/3}\hat{u}_2 \pm \frac{1}{2}\varepsilon _2^{-4/3}v_2 + \frac{1}{2}\varepsilon _2^{-1/3} \pm \alpha \varepsilon _2^{-1/3}. \end{eqnarray*} By applying the transformations (\ref{3-26}), $\hat{X}_2 = X_2 \pm 1$ and (\ref{3-17}), we obtain the local integrals of ($\text{P}_\text{II}$) of the form \begin{eqnarray} \left\{ \begin{array}{ll} \displaystyle C_1 = z \mp y^{-1} + y^{-1} \varphi _2(xy^{-2}\pm 1, zy^{-2}, y^{-3}), \\[0.2cm] \displaystyle C_2 = \left( \frac{1}{2} \pm \alpha \right) y + xy^{2} \pm \frac{1}{2}y^2 z \pm y^4 + y^4 \varphi _1(\cdots ) \pm \frac{1}{2}y \varphi _2(\cdots ). \\ \end{array} \right. \label{3-28} \end{eqnarray} Arguments of $\varphi _1$ and $\varphi _2$ in the second line are the same as that of the first line. The next theorem is proved by the same way as Thm.3.2 for ($\text{P}_\text{I}$). \\[0.2cm] \textbf{Theorem.\thedef.} Any solutions of ($\text{P}_\text{II}$) are meromorphic on $\C$. \\[0.2cm] \textbf{Proof.} Fix a solution $(x(z), y(z))$ of ($\text{P}_\text{II}$) and take a disk $B(z_0,R)$ as in the proof of Thm.3.2. Let $z_* \, (\neq \infty)$ be a singularity on the boundary of the disk. The next lemma implies that the fixed point $(X_2, Z_2, \varepsilon _2) = (\pm 1,0,0)$ corresponds to the singularity $z_*$. \\[0.2cm] \textbf{Lemma.\thedef.} $(X_2, Z_2, \varepsilon _2) \to (1,0,0)$ or $\to (-1,0,0)$ as $z \to z_*$ along a curve $\Gamma $ inside the disk $B(z_0, R)$. \\[0.2cm] \textbf{Proof.} The same argument as the proof of Lemma 3.3 proves that either $x$ or $y$ diverges as $z \to z_*$. (i) Suppose that $y \to \infty$ as $z \to z_*$. We move to the $(X_2, Z_2, \varepsilon _2)$-coordinates. Eq.(\ref{3-17}) provides \begin{equation} X_2 = xy^{-2}, \quad Z_2 = zy^{-2}, \quad \varepsilon _2 = y^{-3}. \end{equation} This immediately yields $Z_2 \to 0,\, \varepsilon _2 \to 0$ as $z \to z_*$. Let us show $X_2 \to \pm 1$. ($\text{P}_\text{II}$) is a Hamiltonian system with the Hamiltonian function $H = x^2/2 - y^4/2 - zy^2/2 - \alpha y$. Thus, the equality $H = -\int\! y(z)^2dz/2$ holds along a solution. In the $(X_2, Z_2, \varepsilon _2)$-coordinates, this equality is written as \begin{eqnarray*} X_2(Z_2)^2 -1 = Z_2 + 2\alpha \varepsilon_2 (Z_2)- \varepsilon _2(Z_2)^{4/3} \int^{Z_2}_{\xi}\! \varepsilon _2(z)^{-2/3}dz, \end{eqnarray*} where $(X_2(Z_2), \varepsilon _2(Z_2))$ is a solution of the ODE solved as a function of $Z_2$, and $\xi$ is a certain nonzero number determined by the initial condition. Since $Z_2 \to 0$ and $\varepsilon _2 \to 0$ as $z \to z_*$, we obtain $X^2_2 \to 1$. (ii) Suppose that $x \to \infty$ as $z \to z_*$. In this case, we use the $(Y_1, Z_1, \varepsilon _1)$-coordinates given by \begin{equation} Y_1 = yx^{-1/2}, \quad Z_1 = zx^{-1}, \quad \varepsilon _1 = x^{-3/2}. \end{equation} By the assumption, we have $Z_1 \to 0,\, \varepsilon _1 \to 0$ as $z \to z_*$. Then, we can show that $Y_1^4 \to 1$ as $z \to z_*$ by the same way as above. This means that $(Y_1, Z_1, \varepsilon _1)$ converges to the fixed point $(1^{1/4},0,0)$ of the vector field (\ref{3-20}). It is easy to verify that this fixed point is the same point as $(X_2, Z_2, \varepsilon _2) = (\pm 1,0,0)$ if written in the $(X_2, Z_2, \varepsilon _2)$-coordinates. $\Box$ Due to the above lemma, when $z$ is sufficiently close to $z_*$, the solution is included in the neighborhood $U$, on which local holomorphic functions $\varphi _1$ and $\varphi _2$ are well defined. Then, the integrals (\ref{3-28}) are available. To apply the implicit function theorem, put \begin{eqnarray} \left( \begin{array}{@{\,}c@{\,}} w \\ u \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} \mp y^{-1} \\ \left( \frac{1}{2} \pm \alpha \right) y + xy^{2} \pm \frac{1}{2}zy^2 \pm y^4 \end{array} \right), \end{eqnarray} Then, (\ref{3-28}) takes the form \begin{eqnarray*} \left\{ \begin{array}{l} \displaystyle C_1 = z + w \mp w \varphi _2(uw^4 \pm ( \frac{1}{2} \pm \alpha )w^3 \mp \frac{1}{2}zw^2, z w^2, \mp w^3),\\ \displaystyle C_2 = u + w^{-4}\varphi _1 (\cdots ) - \frac{1}{2}w^{-1}\varphi _2 (\cdots ). \\ \end{array} \right. \end{eqnarray*} Note that $C_1 = z+ w + O(w^3) $ and $C_2 = u + O(1)$ as $w \to 0$ because $\varphi_1 \sim O(|| \bm{X} ||^2)$ and $\varphi_2 \sim O(\bm{X})$. Let $a$ be a constant such that \begin{eqnarray*} w^{-4}\varphi _1 (uw^4 \pm ( \frac{1}{2} \pm \alpha )w^3 \mp \frac{1}{2}zw^2, z w^2, \mp w^3 ) = az^2 + O(w), \end{eqnarray*} so that $C_2 = u + az^2 + O(w)$ as $w \to 0$. Since $w\to 0$ as $z \to z_*$, the constant $C_1 = z_*$ is just the position of the singularity. If we set \begin{eqnarray*} & & f_1(w,u, z) = z + w \mp w\varphi _2 (\cdots ) - z_* \\ & & f_2(w,u, z) = u + w^{-4}\varphi _1 (\cdots )- \frac{1}{2}w^{-1}\varphi _2 (\cdots ) -C_2, \end{eqnarray*} then $f_i(0,C_2-az_*^2, z_*) = 0$. The Jacobi matrix of $(f_1, f_2)$ with respect to $(w,u)$ at $(w,u,z) = (0, C_2-az_*^2, z_*)$ is \begin{eqnarray*} \left( \begin{array}{@{\,}cc@{\,}} 1 & 0 \\ * & 1 \end{array} \right). \end{eqnarray*} Hence, the implicit function theorem proves that there exists a local holomorphic function $g(z)$ such that $f_1 = f_2 = 0$ is solved as $w = g(z) \sim O(z - z_*)$. Since $y = \mp w^{-1}$, $z = z_*$ is a pole of first order of $y$. This completes the proof of Thm.3.5. $\Box$ \subsection{The fourth Painlev\'{e} equation} ($\text{P}_\text{IV}$) is given on the weighted projective space $\C P^3(1,1,1,2)$ as a tuple of equations (\ref{P4}), (\ref{P41}), (\ref{P42}) and (\ref{P43}). Coordinate transformations between inhomogeneous coordinates are given by \begin{eqnarray} \left\{ \begin{array}{rrrr} x = & \varepsilon _1^{-1/2} = & X_2\varepsilon _2^{-1/2} =& X_3\varepsilon _3^{-1/2} \\ y = & Y_1\varepsilon _1^{-1/2} = & \varepsilon _2^{-1/2}=& Y_3\varepsilon _3^{-1/2}\\ z = & Z_1\varepsilon_1 ^{-1/2} = & Z_2\varepsilon _2^{-1/2}=& \varepsilon _3^{-1/2}. \end{array} \right. \label{3-32} \end{eqnarray} We rewrite Eqs.(\ref{P41}), (\ref{P42}) and (\ref{P43}) as three-dimensional polynomial vector fields as before, which results in \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{Y}_1 = 3Y_1 - 2\kappa_0\varepsilon _1 - Y_1^2 - 2Y_1Z_1 - Y_1(2Y_1 + 2Z_1 - 2\theta _\infty\varepsilon _1), \\ \displaystyle \dot{Z}_1 = Z_1 + \varepsilon _1 - Z_1(2Y_1 + 2Z_1 - 2\theta _\infty\varepsilon _1), \\ \displaystyle \dot{\varepsilon }_1 =2\varepsilon _1 - 2\varepsilon _1 (2Y_1 + 2Z_1 - 2\theta _\infty\varepsilon _1), \end{array} \right. \label{3-33} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_2 = 3X_2 -2\theta _\infty \varepsilon _2 -X_2^2 + 2X_2Z_2 - X_2( 2X_2 - 2Z_2 - 2\kappa_0\varepsilon _2), \\ \displaystyle \dot{Z}_2 = Z_2 + \varepsilon _2 - Z_2 (2X_2 - 2Z_2 - 2\kappa_0\varepsilon _2), \\ \displaystyle \dot{\varepsilon }_2 = 2\varepsilon _2- 2\varepsilon _2 (2X_2 - 2Z_2 - 2\kappa_0\varepsilon _2) , \end{array} \right. \label{3-34} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_3 = -X_3^2+2X_3Y_3+2X_3-2\theta _\infty \varepsilon _3-X_3\varepsilon _3, \\ \displaystyle \dot{Y}_3 = -Y_3^2+2X_3Y_3-2Y_3-2\kappa_0 \varepsilon _3-Y_3\varepsilon _3, \\ \displaystyle \dot{\varepsilon }_3 = -2\varepsilon _3^2 . \end{array} \right. \label{3-35} \end{equation} These vector fields have seven fixed points on the ``infinity set" $\C P^2 (1,1,1)$; \\[0.2cm] \textbf{(I).} $(Y_1, Z_1, \varepsilon _1) = (0,0,0), (1,0,0)$ and $(X_2, Z_2, \varepsilon _2) = (0,0,0)$. \\ \textbf{(II).} $(X_3, Y_3, \varepsilon _3) = (0, 0 ,0), (0,-2,0), (2,0,0)$ and $(2/3, -2/3, 0)$. As in the case of ($\text{P}_\text{II}$), three fixed points in type (I) correspond to movable singularities of ($\text{P}_\text{IV}$), and four fixed points in type (II) correspond to the irregular singular point $z=\infty$ of ($\text{P}_\text{IV}$). Note that other fixed points represent the same point as one of the above. For example, the fixed point $(X_2, Z_2, \varepsilon _2) = (1,0,0)$ is the same as $(Y_1, Z_1, \varepsilon _1) = (1,0,0)$. By the same way as for ($\text{P}_\text{I}$) and ($\text{P}_\text{II}$), we can show that ($\text{P}_\text{IV}$) is locally transformed into a solvable system, and that any solutions of ($\text{P}_\text{IV}$) are meromorphic. Suppose that a solution $(x(z), y(z))$ has a singularity $z = z_*$. As in the proof of Lemma.3.6, we can show that as $z \to z_*$, a solution converges to one of the fixed points listed in type (I) above (the proof is the same as Lemma 3.6 and omitted). Indeed, the Laurent series (i),(ii),(iii) given in Sec.2.3 correspond to the fixed points $(Y_1, Z_1, \varepsilon _1) = (0,0,0), (1,0,0)$ and $(X_2, Z_2, \varepsilon _2) = (0,0,0)$, respectively. Local analysis around these fixed points using the normal form theory is done in the same way as before. Let us consider the fixed point $(Y_1, Z_1, \varepsilon _1) = (0,0,0)$. The Jacobi matrix of the vector field at the origin is given by \begin{equation} J = \left( \begin{array}{@{\,}ccc@{\,}} 3 & 0 & -2\kappa_0 \\ 0& 1& 1 \\ 0 & 0& 2 \end{array} \right). \end{equation} We can confirm that the vector field does not have resonance terms. Hence, due to Poincar\'{e}'s theorem, there exists a neighborhood $U$ of the origin and a local analytic transformation defined on $U$ of the form \begin{eqnarray} \left( \begin{array}{@{\,}c@{\,}} Y_1 \\ Z_1 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} u_1 \\ v_1 \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} Y_1 + \varphi _1(Y_1,Z_1,\varepsilon _1) \\ Z_1 + \varepsilon _1 \varphi _2 (Y_1, Z_1, \varepsilon _1) \end{array} \right), \label{3-38} \end{eqnarray} such that Eq.(\ref{3-33}) is linearized as \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{u}_1 = 3u_1 -2\kappa_0 \varepsilon _1, \\ \displaystyle \dot{v}_1 = v_1 + \varepsilon _1, \\ \displaystyle \dot{\varepsilon }_1 = 2\varepsilon _1, \end{array} \right. \label{3-39} \end{equation} with $\varphi _1 \sim O(|| \bm{Y} ||^2)$ and $\varphi _2 \sim O(\bm{Y} )$, where $\bm{Y} = (Y_1, Z_1, \varepsilon _1)$. For this linear system, we move to the original chart for ($\text{P}_\text{IV}$) by $\tilde{x} = \varepsilon _1^{-1/2}, \tilde{y} = u_1 \varepsilon _1^{-1/2}, \tilde{z} = v_1 \varepsilon _2^{-1/2}$. Then, we obtain the solvable system \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \frac{d\tilde{x}}{d\tilde{z}} = -\tilde{x}^2, \\[0.2cm] \displaystyle \frac{d\tilde{y}}{d\tilde{z}} = 2\tilde{x}\tilde{y} - 2\kappa_0. \\ \end{array} \right. \label{3-40} \end{eqnarray} The coordinate transformation is given by \begin{eqnarray*} \left( \begin{array}{@{\,}c@{\,}} \tilde{x} \\ \tilde{y} \\ \tilde{z} \end{array} \right) = \left( \begin{array}{@{\,}c@{\,}} x \\ y + x \varphi_1(x^{-1}y, x^{-1}z, x^{-2}) \\ z + x^{-1} \varphi_2(x^{-1}y, x^{-1}z, x^{-2}) \end{array} \right). \end{eqnarray*} Hence, we have obtained the next proposition. \\[0.2cm] \textbf{Proposition.\thedef.} There exists a local analytic transformation $(y, z) \mapsto (\tilde{y},\tilde{z})$ defined near $(Y_1, Z_1, \varepsilon _1) = (0, 0,0)$ such that ($\text{P}_\text{IV}$) is transformed into the integrable system (\ref{3-40}). \\ Since Eq.(\ref{3-39}) or (\ref{3-40}) is solvable, we can construct two local analytic integrals of ($\text{P}_\text{IV}$). By applying the implicit function theorem for them, it is proved that if a solution satisfies $(Y_1, Z_1, \varepsilon _1) \to (0,0,0)$ as $z \to z_*$, then $z_*$ is a pole of first order. The same procedure can be done for the other fixed points $(Y_1, Z_1, \varepsilon _1) = (1,0,0)$ and $(X_2, Z_2, \varepsilon _2) = (0,0,0)$ to prove that \\[0.2cm] \textbf{Theorem.\thedef.} Any solutions of ($\text{P}_\text{IV}$) are meromorphic on $\C$. \\ The detailed calculation is the same as the proof Thm.3.5 and left to the reader. \subsection{Characterization of ($\text{P}_\text{I}$)} In order to apply the Poincar\'{e} linearization theorem to Eq.(\ref{3-8}), eigenvalues of the Jacobi matrix (\ref{3-9}) have to satisfy certain conditions and the other components of the matrix are not important. However, to prove the meromorphy of solutions, the $(2,3)$-component of the Jacobi matrix also plays an important role. If the $(2,3)$-component of the Jacobi matrix were zero, the function $f_1(w,u,z)$ defined in the proof of Thm.3.2 becomes $f_1 = z + w \varphi_2(\cdots ) - z_*$ (i.e. the term $w$ does not appear). As a result, the implicit function theorem is not applicable and we can not prove Thm.3.2. To see the geometric role of the $(2,3)$-component, let us consider the dynamical system \begin{equation} \left\{ \begin{array}{l} \dot{x} = 6y^2 + z, \\ \dot{y} = x, \\ \dot{z} = \beta, \end{array} \right. \label{3-41} \end{equation} where $\beta \in \C$ is a constant. When $\beta \neq 0$, this is reduced to ($\text{P}_\text{I}$) by a suitable scaling. This system defines a family of integral curves on $\C^3$. We regard $\C^3$ as a vector bundle; $z$-space is a base and $(x,y)$-space is a fiber. As long as $\beta \neq 0$, each integral curve is a local section of the bundle, while if $\beta = 0$, integral curves are tangent to a fiber and we can not solve the system as a function of $z$. Now we change the coordinates by (\ref{3-1}) and $\hat{X}_2 = X_2 - 2$. Then, Eq.(\ref{3-41}) is brought into the system \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{\hat{X}}_2 = 6\hat{X}_2 - Z_2 + \frac{-3\hat{X}_2^2 + \hat{X}_2Z_2}{2+\hat{X}_2}, \\ \displaystyle \dot{Z}_2 = 4Z_2 - \beta \varepsilon _2 + \frac{\beta \hat{X}_2\varepsilon _2}{2+\hat{X}_2}, \\ \displaystyle \dot{\varepsilon }_2 = 5\varepsilon _2. \end{array} \right. \end{equation} Hence, integral curves give local sections if and only if the $(2,3)$-component of the Jacobi matrix of the above system is not zero. This suggests that the $(2,3)$-component is closely related with the Painlev\'{e} property. On the $(x, y, z)$-coordinates of $\C P^3(3,2,4,5)$, give the ODE \begin{equation} \frac{dx}{dz} = f(x, y, z), \quad \frac{dy}{dz} = g(x, y, z), \label{3-43} \end{equation} where $f$ and $g$ are holomorphic in $x,y$ and meromorphic in $z$. We suppose that this equation defines a meromorphic ODE on $\C P^3(3,2,4,5)$. This means that the equations expressed in the other inhomogeneous coordinates are also meromorphic. We will show later that these equations are rational (recall that a meromorphic function on a projective space is rational). Thus, there are relatively prime polynomials $h_1, h_2, h_3$ such that the equation written in the $(X_2, Z_2, \varepsilon _2)$-coordinates is given by $dX_2/d\varepsilon _2 = h_1(X_2, Z_2, \varepsilon _2)/h_3(X_2, Z_2, \varepsilon _2)$ and $dZ_2/d\varepsilon _2 = h_2(X_2, Z_2, \varepsilon _2)/h_3(X_2, Z_2, \varepsilon _2)$. As before, we introduce a vector field \begin{eqnarray*} \left\{ \begin{array}{l} \dot{X}_2 = h_1(X_2, Z_2, \varepsilon _2), \\ \dot{Z}_2 = h_2(X_2, Z_2, \varepsilon _2), \\ \dot{\varepsilon }_2 = h_3(X_2, Z_2, \varepsilon _2). \end{array} \right. \end{eqnarray*} We call it the associated vector field with $dX_2/d\varepsilon _2 = h_1/h_3, dZ_2/d\varepsilon _2 = h_2/h_3$. The next theorem shows that ($\text{P}_\text{I}$) is characterized by the (i) geometry of $\C P^3(3,2,4,5)$ and (ii) a local condition at a fixed point. Note that there are infinitely many equations satisfying only the condition (i) below. It is remarkable that the condition (ii) seems to be very weak, however, it completely determines an equation. \\[0.2cm] \textbf{Theorem.\thedef.} Consider the ODE (\ref{3-43}), where $f$ and $g$ are holomorphic in $x,y$ and meromorphic in $z$. Suppose the following two conditions: \\[0.2cm] \textbf{(i)} Eq.(\ref{3-43}) defines a meromorphic ODE on $\C P^3(3,2,4,5)$. \\ \textbf{(ii)} The associated polynomial vector field in the $(X_2, Z_2, \varepsilon _2)$-coordinates has a fixed point of the form $(X_2, Z_2, \varepsilon _2) = (X_*, 0,0)$. Eigenvalues and the $(2,3)$-component of the Jacobi matrix at this point are not zero. \\[0.2cm] Then, Eq.(\ref{3-43}) is of the form \begin{eqnarray} \frac{dx}{dz} = a y^2 + bz, \quad \frac{dy}{dz} = cx, \label{3-44} \end{eqnarray} where $a\neq 0, c\neq 0$ and $b$ are constants. When $b\neq 0$, this is equivalent to ($\text{P}_\text{I}$), and when $b=0$, this is equivalent to the integrable equation $y'' = 6y^2$. \\[0.2cm] \textbf{Proof.} At first, we show that $f$ and $g$ are polynomial in $x, y$ and rational in $z$. In the $(X_2, Y_2, \varepsilon _2)$-coordinates, Eq.(\ref{3-43}) is written by \begin{eqnarray*} \left\{ \begin{array}{l} \displaystyle \frac{dX_2}{d\varepsilon _2} = \frac{1}{5\varepsilon _2}\left( 3X_2 - 2\varepsilon _2^{1/5} \frac{f(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5})} {g(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5})}\right), \\ \displaystyle \frac{dZ_2}{d\varepsilon _2} = \frac{1}{5\varepsilon _2}\left( 4Z_2 - 2\varepsilon _2^{2/5} \frac{1} {g(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5})}\right). \end{array} \right. \end{eqnarray*} By the condition (i), the right hand sides are meromorphic in $\varepsilon _2$. In particular, $f(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5})$ and $g(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5})$ are meromorphic as a function of $\varepsilon _2^{1/5}$. On the other hand, they are obviously meromorphic in $\varepsilon _2^{-1/5}$. Thus, they are rational in $\varepsilon _2^{1/5}$. Therefore, the right hand sides above are rational in $\varepsilon _2$. This implies that $\varepsilon _2^{-2/5}g$ and $\varepsilon _2^{-1/5}f$ are rational in $\varepsilon _2$ and we can set \begin{eqnarray*} & & \varepsilon _2^{-1/5} f(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5}) = \frac{\sum h_i^1(X_2, Z_2) \varepsilon _2 ^i}{\sum h_i^2(X_2, Z_2) \varepsilon _2 ^i}, \\ & & \varepsilon _2^{-2/5} g(X_2\varepsilon _2^{-3/5}, \varepsilon _2^{-2/5}, Z_2\varepsilon _2^{-4/5}) = \frac{\sum h_i^3(X_2, Z_2) \varepsilon _2 ^i}{\sum h_i^4(X_2, Z_2) \varepsilon _2 ^i}, \end{eqnarray*} where all $\sum$ are finite sums, and $h_i^1, h_i^2, h_i^3, h_i^4$ are meromorphic in $X_2, Z_2$. In the original chart, $f$ and $g$ are written as \begin{eqnarray} & & f(x, y, z) = \frac{\sum h_i^1(xy^{-3/2}, zy^{-2})y^{-5i/2-1/2}} {\sum h_i^2(xy^{-3/2}, zy^{-2})y^{-5i/2}}, \nonumber\\ & & g(x, y, z) = \frac{\sum h_i^3(xy^{-3/2}, zy^{-2})y^{-5i/2-1}} {\sum h_i^4(xy^{-3/2}, zy^{-2})y^{-5i/2}}. \label{3-45} \end{eqnarray} Next, we move to the $(X_3, Y_3, \varepsilon _3)$-coordinates. Eq.(\ref{3-43}) is written as \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \frac{dX_3}{d\varepsilon _3} = \frac{1}{5\varepsilon _2^2}\left( 3X_3\varepsilon _3 - 4\varepsilon _3^{4/5} f(X_3\varepsilon _3^{-3/5}, Y_3\varepsilon _3^{-2/5}, \varepsilon _3^{-4/5}) \right), \\ \displaystyle \frac{dY_3}{d\varepsilon _3} = \frac{1}{5\varepsilon _3^2}\left( 2Y_3\varepsilon _3 - 4\varepsilon _3^{3/5} g(X_3\varepsilon _3^{-3/5}, Y_3\varepsilon _3^{-2/5}, \varepsilon _3^{-4/5})\right). \end{array} \right. \label{3-46} \end{eqnarray} Substituting (\ref{3-45}) yields \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \frac{dX_3}{d\varepsilon _3} = \frac{1}{5\varepsilon _2^2}\left( 3X_3\varepsilon _3 - 4 \frac{\sum h_i^1 (X_3Y_3^{-3/2}, Y_3^{-2}) Y_3^{-5i/2-1/2}\varepsilon _3^{i+1}} {\sum h_i^2 (X_3Y_3^{-3/2}, Y_3^{-2}) Y_3^{-5i/2}\varepsilon _3^{i}} \right), \\ \displaystyle \frac{dY_3}{d\varepsilon _3} = \frac{1}{5\varepsilon _3^2}\left( 2Y_3\varepsilon _3 - 4 \frac{\sum h_i^3 (X_3Y_3^{-3/2}, Y_3^{-2}) Y_3^{-5i/2-1}\varepsilon _3^{i+1}} {\sum h_i^4 (X_3Y_3^{-3/2}, Y_3^{-2}) Y_3^{-5i/2}\varepsilon _3^{i}} \right). \end{array} \right. \label{3-47} \end{eqnarray} The right hand sides are obviously rational in $\varepsilon _3$. By the same argument as before, they are also rational in $Y_3$. Hence, the right hand sides of Eq.(\ref{3-46}) are rational in $Y_3, \varepsilon _3$, and we can set \begin{eqnarray*} & & \varepsilon _3^{4/5} f(X_3\varepsilon _3^{-3/5}, Y_3\varepsilon _3^{-2/5}, \varepsilon _3^{-4/5}) = \frac{\sum h_{ij}^5(X_3)Y_3^i \varepsilon _3 ^j}{\sum h_{ij}^6(X_3)Y_3^i \varepsilon _3 ^j}, \\ & & \varepsilon _3^{3/5} g(X_3\varepsilon _3^{-3/5}, Y_3\varepsilon _3^{-2/5}, \varepsilon _3^{-4/5}) = \frac{\sum h_{ij}^7(X_3)Y_3^i \varepsilon _3 ^j}{\sum h_{ij}^8(X_3)Y_3^i \varepsilon _3 ^j}, \end{eqnarray*} where $\sum$ are finite sums and $h_{ij}^5, h_{ij}^6, h_{ij}^7, h_{ij}^8$ are meromorphic in $X_3$. Finally, we move to the $(Y_1, Z_1, \varepsilon _1)$-coordinates. Repeating the same procedure, we can verify that $\varepsilon _1^{-1/5} f(\varepsilon _1^{-3/5}, Y_1\varepsilon _1^{-2/5}, Z_1\varepsilon _1^{-4/5})$ and $\varepsilon _1^{-2/5} g(\varepsilon _1^{-3/5}, Y_1\varepsilon _1^{-2/5}, Z_1\varepsilon _1^{-4/5})$ are rational in $Y_1,Z_1, \varepsilon _1$. This proves that $f(x, y, z)$ and $g(x, y, z)$ are rational functions. By the assumption, they are polynomial in $x$ and $y$. Now we can write $g$ and $f/g$ as quotients of polynomials as \begin{equation} g(X,Y,Z) = \frac{\sum a_{ijk} X^iY^jZ^k}{\sum b_k Z^k}, \quad \frac{f(X,Y,Z)}{g(X,Y,Z)} = \frac{\sum p_{ijk} X^iY^jZ^k}{\sum q_{ijk} X^iY^jZ^k}. \label{3-48} \end{equation} Our purpose is to determine coefficients $a_{ijk}, b_k, p_{ijk}$ and $q_{ijk}$ by the conditions (i) and (ii). In the $(X_2, Y_2, Z_2)$-coordinates, Eq.(\ref{3-43}) with (\ref{3-48}) is given by \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \frac{dX_2}{d\varepsilon _2} = \frac{1}{5\varepsilon _2}\left( 3X_2 - 2\varepsilon _2^{1/5} \frac{\sum p_{ijk} X_2^i Z_2^k \varepsilon _2^{-(3i+2j+4k)/5} } {\sum q_{ijk} X_2^i Z_2^k \varepsilon _2^{-(3i+2j+4k)/5}}\right), \\ \displaystyle \frac{dZ_2}{d\varepsilon _2} = \frac{1}{5\varepsilon _2}\left( 4Z_2 - 2\varepsilon _2^{2/5} \frac{\sum b_k Z_2^k \varepsilon _2^{-4k/5}} {\sum a_{ijk} X_2^i Z_2^k \varepsilon _2^{-(3i+2j+4k)/5}}\right). \end{array} \right. \label{3-49} \end{eqnarray} Due to the condition (i), the right hand sides are rational in $\varepsilon _2$. This yields the conditions for coefficients as \begin{equation} \left\{ \begin{array}{l} p_{ijk} \neq 0 \quad \text{only when} \quad 3i+2j+4k-1 = 5m + \delta, \quad (m=-1,0,\cdots , M), \\ q_{ijk} \neq 0 \quad \text{only when} \quad 3i+2j+4k = 5m' + \delta, \quad (m'=0,1,\cdots , M'), \\ b_{k} \neq 0 \quad \text{only when} \quad 4k-2 = 5n + \delta', \quad (n=-1,0,\cdots , N), \\ a_{ijk} \neq 0 \quad \text{only when} \quad 3i+2j+4k = 5n' + \delta', \quad (n'=0,1,\cdots , N'), \end{array} \right. \label{3-50} \end{equation} where $\delta , \delta ' \in \{ 0,1,2,3,4\}$. More precisely, the first line means that $p_{ijk} \neq 0$ only when there are integers $m$ and $\delta $ such that $(i,j,k)$ satisfies $3i+2j+4k-1 = 5m + \delta$, where $\delta \in \{ 0,1,2,3,4\}$ is independent of $(i,j,k)$. Since Eq.(\ref{3-49}) is rational, there is the largest integer $m$ satisfying $3i+2j+4k-1 = 5m + \delta$ and $p_{ijk} \neq 0$. In (\ref{3-50}), the largest integer is denoted by $M$. Integers $M', N$ and $N'$ play a similar role. Then, Eq.(\ref{3-49}) is rewritten as \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \frac{dX_2}{d\varepsilon _2} = \frac{1}{5\varepsilon _2}\left( 3X_2 - 2 \frac{\sum p_{ijk} X_2^i Z_2^k \varepsilon _2^{-m} } {\sum q_{ijk} X_2^i Z_2^k \varepsilon _2^{-m'}}\right), \\ \displaystyle \frac{dZ_2}{d\varepsilon _2} = \frac{1}{5\varepsilon _2}\left( 4Z_2 - 2 \frac{\sum b_k Z_2^k \varepsilon _2^{-n}} {\sum a_{ijk} X_2^i Z_2^k \varepsilon _2^{-n'}}\right). \end{array} \right. \label{3-51} \end{eqnarray} In order to confirm the condition (ii), we shall rewrite it as a polynomial vector field. \textbf{(I).} When $M>M'$ and $N>N'$, the associated polynomial vector field is of the form \begin{equation} \left\{ \begin{array}{l} \dot{X}_2 = 3X_2 (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N-n'}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M-m'}) \\ \qquad \qquad \qquad -2 (\sum p_{ijk} X_2^iZ_2^k \varepsilon _2^{M-m}) (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N-n'}), \\[0.4cm] \dot{Z}_2 = 4Z_2 (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N-n'}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M-m'}) \\ \qquad \qquad \qquad -2 (\sum b_{k} Z_2^k \varepsilon _2^{N-n}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M-m'}), \\[0.4cm] \dot{\varepsilon }_2 = 5\varepsilon _2 (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N-n'}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M-m'}). \end{array} \right. \end{equation} Because of the condition (ii), we seek a fixed point of the form $(X_*,0,0)$ with nonzero eigenvalues. Since $N-n' > 0$ and $M-m' > 0$, the right hand side of the equation of $\varepsilon _2$ is of order $O(\varepsilon _2^3)$. Hence, the Jacobi matrix at the point $(X_*,0,0)$ has a zero eigenvalue. In a similar manner, we can verify that two cases $M>M', N\leq N'$ and $M\leq M', N>N'$ are excluded. In these cases, the right hand side of the equation of $\varepsilon _2$ is of order $O(\varepsilon _2^2)$ and the Jacobi matrix has a zero eigenvalue. \textbf{(II).} When $M\leq M'$ and $N\leq N'$, the associated polynomial vector field is of the form \begin{equation} \left\{ \begin{array}{l} \dot{X}_2 = 3X_2 (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N'-n'}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M'-m'}) \\ \qquad \qquad \qquad -2 (\sum p_{ijk} X_2^iZ_2^k \varepsilon _2^{M'-m}) (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N'-n'}), \\[0.4cm] \dot{Z}_2 = 4Z_2 (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N'-n'}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M'-m'}) \\ \qquad \qquad \qquad -2 (\sum b_{k} Z_2^k \varepsilon _2^{N'-n}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M'-m'}), \\[0.4cm] \dot{\varepsilon }_2 = 5\varepsilon _2 (\sum a_{ijk} X_2^iZ_2^k \varepsilon _2^{N'-n'}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M'-m'}). \end{array} \right. \label{3-53} \end{equation} Suppose that this has a fixed point of the form $(X_*,0,0)$. The $(2,3)$-component of the Jacobi matrix at this point is given by \begin{eqnarray*} -2 \cdot \frac{\partial }{\partial \varepsilon _2}\Bigl|_{(X_*, 0,0)} (\sum b_{k} Z_2^k \varepsilon _2^{N'-n}) (\sum q_{ijk} X_2^iZ_2^k \varepsilon _2^{M'-m'}). \end{eqnarray*} We require that this quantity is not zero. \textbf{(II-a).} Suppose that the polynomial $\sum b_{k} Z_2^k \varepsilon _2^{N'-n}$ includes a constant term. This means that $b_k\neq 0$ when $k=0$ and $n= N'$. Substituting it to the third condition of (\ref{3-50}) provides $-2 = 5N' + \delta'$. This proves that $N' = -1$ and $\delta ' = 3$. Then, the fourth condition of (\ref{3-50}) yields $3i+2j+4k = -2$. Since there are no nonnegative integers $i,j,k$ satisfying this relation, $a_{ijk} = 0$ for any $i,j,k$. In this case, we obtain $\dot{\varepsilon }_2 = 0$ and the Jacobi matrix has a zero eigenvalue. \textbf{(II-b).} When $\sum b_{k} Z_2^k \varepsilon _2^{N'-n}$ does not have a constant term, it has to include a monomial $\varepsilon _2$. Otherwise, the $(2,3)$-component of the Jacobi matrix becomes zero. This means that $b_k\neq 0$ when $k=0$ and $n= N'-1$. The third condition of (\ref{3-50}) provides $-2 = 5(N'-1) + \delta'$, which proves $N' = 0$ and $\delta ' = 3$. Since $N\leq N'$, we have $N = -1$ or $N=0$. In this case, (\ref{3-50}) becomes \begin{equation} \left\{ \begin{array}{l} b_{k} \neq 0 \quad \text{only when} \quad 4k-2 = -2\quad \text{or}\quad 3,\\ a_{ijk} \neq 0 \quad \text{only when} \quad 3i+2j+4k = 3. \end{array} \right. \end{equation} Therefore, nonzero numbers among these coefficients are only $b_0$ and $a_{100}$. This proves that $g(X,Y,Z)$ is given by $a_{100}X/b_0$. In what follows, we put $a_{100}/b_0 = c$. Since $g=cX$ and $f$ is polynomial in $X,Y$, $f/g$ can be written as \begin{eqnarray*} \frac{f(X,Y,Z)}{g(X,Y,Z)} = \frac{\sum p_{ijk}X^iY^jZ^k}{\sum q_{10k}XZ^k}. \end{eqnarray*} In this case, the equation of $\varepsilon _2$ in (\ref{3-22}) is given by \begin{equation} \dot{\varepsilon }_2 = 5a_{100}X_2^2\varepsilon _2 \cdot \left( \sum q_{10k}Z_2^k \varepsilon _2^{M'-m'} \right). \end{equation} The polynomial $\sum q_{10k}Z_2^k \varepsilon _2^{M'-m'}$ has to include a constant term so that the Jacobi matrix at $(X_*, 0,0)$ does not have a zero eigenvalue. This means that $q_{10k} \neq 0$ when $k = 0$ and $m' = M'$. Thus (\ref{3-50}) provides $3 = 5M' + \delta $. This shows $M' = 0$ and $\delta =3$. Then, (\ref{3-50}) becomes \begin{equation} \left\{ \begin{array}{l} p_{ijk} \neq 0 \quad \text{only when} \quad 3i+2j+4k-1 = 3,\\ q_{10k} \neq 0 \quad \text{only when} \quad 3+4k = 3. \end{array} \right. \end{equation} Therefore, nonzero numbers among these coefficients are only $p_{020}, p_{001}$ and $q_{100}$. Putting $a := cp_{020}/q_{100}$ and $b:= cp_{001}/q_{100}$, we obtain the ODE (\ref{3-44}) as a necessary condition for (i) and (ii). It is straightforward to confirm that (\ref{3-44}) actually satisfies the conditions (i) and (ii) when $a\neq 0, c\neq 0$. This completes the proof. $\Box$ \section{The space of initial conditions} In this section, we construct the spaces of initial conditions for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) by the weighted blow-ups of $\C P^3(p,q,r,s)$. For a polynomial system, a manifold $E(z)$ parameterized by $z\in \C$ is called the space of initial conditions if any solutions give global holomorphic sections on the fiber bundle $\mathcal{P} = \{ (x,z) \, | \, x\in E(z) , z\in \C\}$ over $\C$. For the Painlev\'{e} equations, it was first constructed by Okamoto \cite{Oka} by blow-ups of a Hirzebruch surface eight times and by removing a certain divisor called vertical leaves. Different approaches are proposed by Duistermaat and Joshi \cite{Dui} and Iwasaki and Okada \cite{Iwa}. They also performed blow-ups many times. See \cite{Sak, Sai} for algebro-geometric approach. Here, we will obtain the spaces of initial conditions by weighted blow-ups only one time for ($\text{P}_\text{I}$), two times for ($\text{P}_\text{II}$) and three times for ($\text{P}_\text{IV}$). These numbers are the same as the numbers of types of Laurent series given in Sec.2.3. We will easily find a symplectic structure of the space of initial conditions. For ($\text{P}_\text{I}$), we will recover Painlev\'{e}'s coordinates in a purely geometric manner. We find a symplectic structure of the space of initial conditions and show that Painlev\'{e}'s coordinates are the Darboux coordinates of the symplectic structure. \subsection{The first Painlev\'{e} equation} Recall that ($\text{P}_\text{I}$) written in the $(X_2, Z_2, \varepsilon _2)$-coordinates has two fixed points $(\pm 2,0,0)$. Putting $\hat{X}_2 = X_2\pm 2$, we obtain \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{\hat{X}}_2 = 6\hat{X}_2 \pm Z_2 + \frac{\pm 3\hat{X}_2^2 + \hat{X}_2Z_2}{2 \mp \hat{X}_2}, \\ \displaystyle \dot{Z}_2 = 4Z_2 \pm \varepsilon _2 + \frac{\hat{X}_2\varepsilon _2}{2 \mp \hat{X}_2}, \\ \displaystyle \dot{\varepsilon }_2 = 5\varepsilon _2. \end{array} \right. \label{4-1} \end{equation} (In Sec.3, we used only $\hat{X}_2 = X_2- 2$). The origin is a fixed point of the vector field and it is a singularity of the foliation defined by integral curves. We apply a blow-up to this point. At first, we change the coordinates by the linear transformation \begin{eqnarray*} \left\{ \begin{array}{l} \hat{X}_2 = u \mp \frac{1}{2}v - \frac{1}{2}w, \\ Z_2 = v, \\ \varepsilon _2 = w. \end{array} \right. \end{eqnarray*} This yields \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \dot{u} = 6u + f_1(u,v,w), \\ \displaystyle \dot{v} = 4v \pm w + f_2(u,v,w), \\ \displaystyle \dot{w} = 5w, \end{array} \right. \label{4-2} \end{eqnarray} where $f_1$ and $f_2$ denote nonlinear terms. Note that the linear part is \textit{not} diagonalized; since we know that the $(2,3)$-component is important (Thm.3.9), we remove only the $(1,2)$-component of the linear part of (\ref{4-1}). Now we introduce the weighted blow-up with weights $6,4,5$, which are taken from eigenvalues of the Jacobi matrix, defined by the following transformations \begin{eqnarray} \left\{ \begin{array}{llll} u & = u_1^6 & = v_2^6 u_2 & = w_3^6 u_3, \\ v & = u_1^4v_1 & = v_2^4 & = w_3^4 v_3, \\ w & = u_1^5w_1 & = v_2^5 w_2 & = w_3^5. \end{array} \right. \label{4-3} \end{eqnarray} The exceptional divisor $\{ u_1 = 0\} \cup \{ v_2 = 0\} \cup \{ w_3= 0\}$ is a 2-dim weighted projective space $\C P^2(6,4,5)$ and the blow-up of $(u,v,w)$-space is a (singular) line bundle over $\C P^2(6,4,5)$. We mainly use the $(u_3, v_3, w_3)$-coordinates. In the $(u_3, v_3, w_3)$-coordinates, Eq.(\ref{4-2}) is written as \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du}{dv} = \frac{1}{8} \left( v^2w \pm 3vw^2 + 2w^3 \mp 8uvw^3 -10uw^4 + 12 u^2w^5 \right), \\[0.2cm] \displaystyle \frac{dw}{dv} = \frac{1}{4}\left( \pm 4 \pm vw^4 + w^5 - 2uw^6 \right), \\ \end{array} \right. \label{4-4} \end{equation} where the subscripts for $u_3, v_3, w_3$ are omitted for simplicity. The relation between the original chart $(x, y, z)$ and $(u_3, v_3, w_3)$ is \begin{equation} \left\{ \begin{array}{l} \displaystyle x = u_3w_3^3 \mp 2w_3^{-3} \mp \frac{1}{2}v_3 w_3 - \frac{1}{2} w_3^2, \\ \displaystyle y = w_3^{-2}, \\ \displaystyle z = v_3, \end{array} \right. \label{4-5} \end{equation} or \begin{equation} \left\{ \begin{array}{l} \displaystyle u_3 = xy^{3/2} \pm 2y^3 \pm \frac{1}{2}zy + \frac{1}{2}y^{1/2}, \\ \displaystyle w_3 = y^{-1/2}, \\ \displaystyle v_3 = z. \end{array} \right. \label{4-6} \end{equation} It is remarkable that the independent variable $z$ is not changed despite the fact that $z$ is changed in each step of transformations. Now we have recovered Painlev\'{e}'s coordinates (\ref{4-5}) which was introduced in his paper to prove the Painlev\'{e} property of ($\text{P}_\text{I}$). He found this transformation by observing a Laurent series of a solution, see also Gromak, Laine and Shimomura \cite{Gro}. At a first glance, $\C^2_{(x,y)} \cup \C^2_{(u_3, w_3)}$ glued by (\ref{4-5}) does not define a manifold because (\ref{4-5}) is not one-to-one but one-to-two. Nevertheless, we can show that (\ref{4-5}) defines a certain algebraic surface. Recall that the $(X_2, Z_2,\varepsilon _2)$-space should be divided by the $\Z_2$ action $(X_2, Z_2,\varepsilon _2) \mapsto (-X_2, Z_2, -\varepsilon _2)$ due to the orbifold structure of $\C P^3(3,2,4,5)$. This action induces a certain $\Z_2$ action on the $(u_3, v_3, w_3)$-space. If we divide the $(u_3, v_3, w_3)$-space by the $\Z_2$ action, we can prove that (\ref{4-5}) becomes a one-to-one mapping. Then, $\C^2_{(x,y)}$ and $\C^2_{(u_3, w_3)}/\Z_2$ are glued by (\ref{4-5}) to define an algebraic surface, which gives the space of initial conditions for ($\text{P}_\text{I}$). The space $\C^2_{(u_3, w_3)}/ \Z_2$ is realized as a nonsingular algebraic surface defined by \begin{equation} M(z) : V^2 = UW^4 + 2z W^3 + 4W \label{4-7} \end{equation} with the parameter $z$ (the proof is given below). By using $(U,V,W)$, the relations (\ref{4-5}),(\ref{4-6}) are rewritten as \begin{equation} \left\{ \begin{array}{l} \displaystyle x = VW^{-2} - \frac{1}{2}W, \\ \displaystyle y = W^{-1}. \\ \end{array} \right. ,\quad \left\{ \begin{array}{l} \displaystyle V = xy^{-2} + \frac{1}{2}y^{-3}, \\ \displaystyle W = y^{-1}. \end{array} \right. \label{4-8} \end{equation} Hence, $\C^2_{(x,y)}$ and $M(z)$ glued by this relation defines a nonsingular algebraic surface denoted by $E(z)$. The surface $M(z)$ admits a holomorphic symplectic form \begin{equation} -\frac{1}{W^4}dV \wedge dW = \frac{1}{4UW^3 + 6z W^2 + 4}dV \wedge dU. \end{equation} We can verify that \begin{equation} dx \wedge dy = -\frac{1}{W^4}dV \wedge dW. \end{equation} Thus, $E(z)$ also has a holomorphic symplectic form. ($\text{P}_\text{I}$) written by $(V,W)$ is \begin{equation} \left\{ \begin{array}{ll} \displaystyle \frac{dV}{dz} = 6 + zW^2 + \frac{1}{4W}(W^3 + 4V)(W^3 - 2V) = W^4 \frac{\partial H}{\partial W} \\[0.4cm] \displaystyle \frac{dW}{dz} = \frac{1}{2}W^3 - V = -W^4 \frac{\partial H}{\partial V}, \\ \end{array} \right. \label{4-11} \end{equation} where $H$ is given by \begin{eqnarray} H &=& \frac{1}{2}x^2 - 2y^3 - zy \nonumber \\ &=& \frac{V^2}{2W^4} - \frac{V}{2W} + \frac{1}{8}W^2 - \frac{2}{W^3} - \frac{z}{W}. \end{eqnarray} Hence, Eq.(\ref{4-11}) is a Hamiltonian system with respect to the symplectic form $-W^{-4}dV \wedge dW$ as well as the original ($\text{P}_\text{I}$) written by $(x, y)$. Let $z_*$ be a pole of a solution of ($\text{P}_\text{I}$). As $z \to z_*$, $x, y \to \infty$ and $V,W \to 0$. By using (\ref{4-7}), it is easy to verify that Eq.(\ref{4-11}) is holomorphic even at $W = 0$. This proves that $E(z) = \C^2_{(x,y)} \cup M(z)$ is the desired space of initial conditions. Note that the system (\ref{4-4}) is already a Hamiltonian system with the Hamiltonian \begin{eqnarray} H = \frac{1}{8} \left( \pm 8u \pm 2uvw^4 + 2uw^5 - 2u^2 w^6 - \frac{1}{2}v^2 w^2 \mp vw^3 - \frac{1}{2}w^4\right) . \end{eqnarray} The transformation (\ref{4-5}) yields \begin{equation} dx \wedge dy = -2 du_3 \wedge dw_3. \end{equation} This means that $(u_3, w_3)$-coordinates are the Darboux coordinates for the form $-W^{-4}dV \wedge dW$ (if we remove the factor $-2$ by a suitable scaling). These results are summarized as follows. \\[0.2cm] \textbf{Theorem.\thedef.} \noindent \textbf{(i)} The space $\C^2_{(u_3, w_3)}$ divided by the $\Z_2$ action induced from the orbifold structure of $\C P^3(3,2,4,5)$ gives the algebraic surface $M(z)$. The space of initial conditions $E(z)$ for ($\text{P}_\text{I}$) is given by $\C^2_{(x, x)} \cup M(z)$ glued by (\ref{4-8}). \noindent \textbf{(ii)} $M(z)$ and $E(z)$ have holomorphic symplectic forms and ($\text{P}_\text{I}$) is a Hamiltonian system with respect to the form. Painlev\'{e}'s coordinates defined by (\ref{4-5}) are the Darboux coordinates of the symplectic form on $M(z)$. \noindent \textbf{(iii)} Consider an ODE (\ref{3-43}) defined on the $(x, y, z)$-coordinates, where $f$ and $g$ are polynomials in $x$ and $y$. If it is also expressed as a polynomial ODE in the Painlev\'{e}'s coordinates, then (\ref{3-43}) is ($\text{P}_\text{I}$). \\ Recently, a similar result is obtained by Iwasaki and Okada \cite{Iwa} by a different approach. Symplectic atlases for the Painlev\'{e} equations are found by Takano et al. \cite{Tak1, Tak2, Tak3} for the second Painlev\'{e} to sixth Painlev\'{e} equations, while left open for ($\text{P}_\text{I}$). \\[0.2cm] \textbf{Proof.} Due to the orbifold structure of $\C P^3(3,2,4,5)$, the $(X_2, Z_2, \varepsilon _2)$-space should be divided by the $\Z_2$ action $(X_2, Z_2, \varepsilon _2) \mapsto (-X_2, Z_2, -\varepsilon _2)$. It is straightforward to show that in the $(u_3, v_3, w_3)$-coordinates, this action is written by \begin{equation} \left( \begin{array}{@{\,}c@{\,}} u_3 \\ v_3 \\ w_3 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} -u_3 + v_3 w_3^{-2} + 4w_3^{-6} \\ v_3 \\ -w_3 \end{array} \right) . \end{equation} Since $v_3$ is fixed, we consider $\C^2_{(u_3, w_3)} /\Z_2$. Polynomial invariants of this action are generated by \begin{equation} \left\{ \begin{array}{l} \displaystyle U = u_3(u_3 w_3^6 - v_3w_3^4 - 4) + \frac{1}{4}w_3^2v_3^2, \\ \displaystyle V = w_3^7 (u_3- \frac{1}{2}v_3 w_3^{-2} - 2w_3^{-6}) , \\ \displaystyle W = w_3^2. \end{array} \right. \end{equation} They satisfy the equation (\ref{4-7}), which proves $\C^2_{(u_3, w_3)} /\Z_2 = M(z)$. The rest of (i) and (ii) have already been shown. Let us prove (iii). In what follows, we omit the subscripts for $u_3, v_3, w_3$. By (\ref{4-5}) with the upper sign, Eq.(\ref{3-43}) is written in Painlev\'{e}'s coordinates as \begin{eqnarray} \frac{d}{dz}\left( \begin{array}{@{\,}c@{\,}} u \\ w \end{array} \right) &=& \left( \begin{array}{@{\,}c@{\,}} \displaystyle w^{-3}\cdot f + \frac{3}{2}uw^2 \cdot g - \frac{1}{4} z \cdot g + 3w^{-4} \cdot g - \frac{1}{2}w \cdot g + \frac{1}{2} w^{-2} \\ \displaystyle -\frac{1}{2} w^3 \cdot g \end{array} \right) \nonumber \\ &=:& T(f,g) + \left( \begin{array}{@{\,}c@{\,}} \frac{1}{2}w^{-2} \\ 0 \end{array} \right) =: \hat{T}(f,g). \end{eqnarray} Our purpose is to show that if $\hat{T}(f,g)$ is a polynomial, then Eq.(\ref{3-43}) is ($\text{P}_\text{I}$). Since we know that $\hat{T}(f,g)$ is a polynomial for ($\text{P}_\text{I}$), it is sufficient to show the uniqueness. We define operators $T$ and $\hat{T}$ as above. The operator $T$ is a linear mapping from the space of polynomials $\C [x, y] \times \C [x, y]$ into the space of Laurent polynomials $\C [u,w,w^{-1}] \times \C [u,w,w^{-1}] $ for each $z$. Let \begin{equation} \Pi : \C [u,w,w^{-1}] \times \C [u,w,w^{-1}] \to \C [u,w^{-1}] \times \C [u,w^{-1}] \end{equation} be the natural projection to the principle part. If there are two pairs of polynomials $(f_1, g_1)$ and $(f_2, g_2)$ such that $\hat{T}(f_i, g_i),\, (i = 1,2)$ are polynomials, then \begin{eqnarray*} \hat{T}(f_1, g_1) - \hat{T}(f_2, g_2) = T(f_1 - f_2, g_1- g_2) \end{eqnarray*} is also a polynomial and $\Pi \circ T(f_1 - f_2, g_1- g_2) = 0$. Hence, it is sufficient to prove $\mathrm{Ker}\, \Pi \circ T = \{ 0\}$. For this purpose, we show that images of monomials of the form $(X^mY^n, 0), (0, X^mY^n),\,\, (m,n = 0,1,\cdots )$ are linearly independent. They are calculated as \begin{eqnarray*} T (X^mY^n, 0) &=& \left( \begin{array}{@{\,}c@{\,}} w^{-2n-3} (uw^3 - 2w^{-3} - \frac{1}{2} z w - \frac{1}{2} w^2)^m \\ 0 \end{array} \right), \\ T(0, X^mY^n) &=& \left( \begin{array}{@{\,}c@{\,}} * \\ -\frac{1}{2}w^{-2n+3} (uw^3 - 2w^{-3} - \frac{1}{2} z w - \frac{1}{2} w^2)^m \end{array} \right). \end{eqnarray*} It is easy to verify that the principle parts of them are linearly independent. $\Box$ \\[0.2cm] \textbf{Remark.\thedef.} We have constructed the space of initial conditions by the weighted blow-up at the fixed point $(X_2, Z_2, \varepsilon _2) = (2,0,0)$. We can also construct it by using the fixed point $(Y_1, Z_1, \varepsilon _1) = ((1/4)^{1/3} , 0,0)$, which represents the same point as $(X_2, Z_2, \varepsilon _2) = (2,0,0)$. By the same procedure as before (an affine transformation and the weighted blow-up in the $(Y_1, Z_1, \varepsilon _1)$-coordinates), we obtain the one-to-three transformation \begin{equation} \left\{ \begin{array}{l} \displaystyle x = w_3^{-3}, \\ \displaystyle y = u_3 w_3^4 - \frac{2^{-1/3}}{3}v_3 w_3^2 + 2^{-2/3} w_3^{-2}, \\ \displaystyle z = v_3. \end{array} \right. \label{4-19} \end{equation} Due to the orbifold structure, the $(Y_1, Z_1, \varepsilon _1)$-space should be divided by the $\Z_3$ action (\ref{3-2}). This induces the $\Z_3$ action in the $(u_3, v_3, w_3)$-space and we can show that $\C^2_{(x, y)}$ and $\C^2_{(u_3, w_3)}/\Z_3$ glued by (\ref{4-19}) gives the same algebraic surface $E(z)$ as before. \subsection{The second Painlev\'{e} equation} Recall that ($\text{P}_\text{II}$) written in the $(X_2, Z_2, \varepsilon _2)$-coordinates has two fixed points $(\pm 1,0,0)$. Putting $\hat{X}_2 = X_2\pm 1$, we have obtained Eq.(\ref{3-23}). The origin is a fixed point of (\ref{3-23}) and it is a singularity of the foliation defined by integral curves. We apply a blow-up to this point. At first, we change the coordinates by the linear transformation \begin{eqnarray} \left\{ \begin{array}{l} \hat{X}_2 = u \mp \frac{1}{2}v - \left( \frac{1}{2} \pm \alpha \right)w, \\ Z_2 = v, \\ \varepsilon _2 = w. \end{array} \right. \label{4-20} \end{eqnarray} Then, we obtain \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \dot{u} = 4u + f_1(u,v,w), \\ \displaystyle \dot{v} = 2v \pm w + f_2(u,v,w), \\ \displaystyle \dot{w} = 3w, \end{array} \right. \label{4-21} \end{eqnarray} where $f_1$ and $f_2$ denote nonlinear terms. Now we introduce the weighted blow-up with weights $4,2,3$, which are taken from eigenvalues of the Jacobi matrix, defined by the following transformations \begin{eqnarray} \left\{ \begin{array}{llll} u & = u_1^4 & = v_2^4 u_2 & = w_3^4 u_3, \\ v & = u_1^2v_1 & = v_2^2 & = w_3^2 v_3, \\ w & = u_1^3w_1 & = v_2^3 w_2 & = w_3^3. \end{array} \right. \label{4-22} \end{eqnarray} The exceptional divisor $\{ u_1 = 0\} \cup \{ v_2 = 0\} \cup \{ w_3= 0\}$ is a 2-dim weighted projective space $\C P^2(4,2,3)$ and the blow-up of $(u,v,w)$-space is a (singular) line bundle over $\C P^2(4,2,3)$. Note that we performed the blow-ups at two points $(X_2, Z_2, \varepsilon _2) = (1,0,0)$ and $(-1,0,0)$. If we want to distinguish the sign of $\hat{X}_2 = X_2 \pm 1$, the notation $(u^{\pm}_i,v^{\pm}_i,w^{\pm}_i)$ for $i=1,2,3$ will be used. In the $(u_3, v_3, w_3)$-coordinates, Eq.(\ref{4-21}) is written as \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du}{dv} = \frac{1}{4} \left( 4uw-1\mp 2\alpha \right) \left( \mp v+2uw^2-(1\pm 2\alpha ) w\right) = -\frac{\partial \widetilde{H}}{\partial w}, \\[0.2cm] \displaystyle \frac{dw}{dv} = \frac{1}{2}\left( -2uw^4+(1\pm 2\alpha ) w^3\pm vw^2 \pm 2 \right) = \frac{\partial \widetilde{H}}{\partial u}, \\ \end{array} \right. \label{4-23} \end{equation} where the subscripts for $u_3, v_3, w_3$ are omitted for simplicity. This is a Hamiltonian system with the Hamiltonian \begin{equation} \widetilde{H} = \frac{1}{2}(-u^2w^4+(1\pm 2\alpha )uw^3\pm uvw^2 \pm 2u) \mp \frac{1}{4}(1\pm 2\alpha ) (vw\pm \frac{1}{2}(1\pm 2\alpha )w^2), \end{equation} as well as the original ($\text{P}_\text{II}$). The relation between the original chart $(x, y, z)$ and $(u_3, v_3, w_3)$ is \begin{equation} \left\{ \begin{array}{l} \displaystyle x = u_3w_3^2 \mp w_3^{-2} \mp \frac{1}{2}v_3 - (\frac{1}{2} \pm \alpha ) w_3, \\ \displaystyle y = w_3^{-1}, \\ \displaystyle z = v_3, \end{array} \right. \label{4-25} \end{equation} or \begin{equation} \left\{ \begin{array}{l} \displaystyle u_3 = xy^{2} \pm y^4 \pm \frac{1}{2}zy^2 + (\frac{1}{2}\pm \alpha )y, \\ \displaystyle w_3 = y^{-1}, \\ \displaystyle v_3 = z. \end{array} \right. \label{4-26} \end{equation} It is remarkable that the system (\ref{4-23}) is polynomial, and the independent variable $z$ is not changed despite the fact that $z$ is changed in each step of transformations. Now we distinguish the choice of the sign of $\hat{X}_2 = X_2 \pm 1$. For the upper sign $\hat{X}_2 = X_2 + 1$ and the lower one $\hat{X}_2 = X_2 - 1$, we use the notation $(u_3^+, v_3^+, w_3^+)$ and $(u_3^-, v_3^-, w_3^-)$, respectively. Eq.(\ref{4-26}) should be \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle u^+_3 = xy^{2} + y^4 + \frac{1}{2}zy^2 + (\frac{1}{2} + \alpha )y, \\ \displaystyle w^+_3 = y^{-1}, \\ \displaystyle v^+_3 = z. \end{array} \right. \left\{ \begin{array}{l} \displaystyle u^-_3 = xy^{2} - y^4 - \frac{1}{2}zy^2 + (\frac{1}{2} - \alpha )y, \\ \displaystyle w^-_3 = y^{-1}, \\ \displaystyle v^-_3 = z. \end{array} \right. \label{4-27} \end{eqnarray} Thus, $\C^2_{(x,y)}, \C^2_{(u^+_3, w^+_3)}$ and $\C^2_{(u^-_3, w^-_3)}$ are glued by (\ref{4-27}) to define an algebraic surface $E(z)$, which gives the space of initial conditions for ($\text{P}_\text{II}$). The transformation (\ref{4-27}) yields \begin{equation} dy \wedge dx = du^{\pm}_3 \wedge dw^{\pm}_3, \end{equation} where $z$ is regarded as a parameter, and \begin{equation} dy \wedge dx - dH \wedge dz = du^{\pm}_3 \wedge dw^{\pm}_3 - d\widetilde{H} \wedge dz, \end{equation} where $z$ is regarded as a coordinate. These results are summarized as follows. \\[0.2cm] \textbf{Theorem.\thedef.} \noindent \textbf{(i)} The space of initial conditions $E(z)$ for ($\text{P}_\text{II}$) is given by $\C^2_{(x, x)} \cup \C^2_{(u^+_3, w^+_3)}\cup \C^2_{(u^-_3, w^-_3)}$ glued by (\ref{4-27}). \noindent \textbf{(ii)} The transformation (\ref{4-27}) is symplectic, and ($\text{P}_\text{II}$) written in $(u_3^+, w_3^+)$ and $(u_3^-, w_3^-)$ are also polynomial Hamiltonian systems. \noindent \textbf{(iii)} Consider a Hamiltonian system (\ref{3-43}) defined on the $(x, y, z)$-coordinates, where $f$ and $g$ are polynomials in $x$ and $y$. If it is also expressed as a polynomial Hamiltonian system in the $(u^{\pm}_3, w^{\pm}_3)$-coordinates, then (\ref{3-43}) is ($\text{P}_\text{II}$). \\ Part (iii) is proved in the same way as Thm.4.1 and omitted. The same result is also obtained by Takano et al. \cite{Tak1, Tak2, Tak3} using a slightly different coordinates. \subsection{The fourth Painlev\'{e} equation} ($\text{P}_\text{IV}$) have been written as three dimensional vector fields (\ref{3-33}), (\ref{3-34}) and (\ref{3-35}). They have three fixed points $(Y_1, Z_1, \varepsilon _1) = (0,0,0), (1,0,0)$ and $(X_2, Z_2, \varepsilon _2) = (0,0,0)$, which correspond to movable singularities. Let us construct the space of initial conditions for ($\text{P}_\text{IV}$) by weighted blow-ups at the three fixed points. \textbf{(i)} $(Y_1, Z_1, \varepsilon _1) = (0,0,0)$. For Eq.(\ref{3-33}), we change the coordinates by the linear transformation \begin{eqnarray} \left\{ \begin{array}{l} Y_1 = U_1 + 2\kappa_0 W_1, \\ Z_1 = V_1, \\ \varepsilon _1 = W_1, \end{array} \right. \end{eqnarray} which results in \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \dot{U}_1 = 3U_1 + f_1(U_1, V_1, W_1), \\ \displaystyle \dot{V}_1 = V_1 + W_1 + f_2(U_1, V_1, W_1), \\ \displaystyle \dot{W}_1 = 2W_1, \end{array} \right. \label{4-32} \end{eqnarray} where $f_1$ and $f_2$ denote nonlinear terms. We introduce the weighted blow-up with weights $3,1,2$ by \begin{eqnarray} \left\{ \begin{array}{llll} U_1 & = u_1^3 & = v_2^3 u_2 & = w_3^3 u_3, \\ V_1 & = u_1v_1 & = v_2 & = w_3 v_3, \\ W_1 & = u_1^2w_1 & = v_2^2 w_2 & = w_3^2. \end{array} \right. \label{4-33} \end{eqnarray} The exceptional divisor $\{ u_1 = 0\} \cup \{ v_2 = 0\} \cup \{ w_3= 0\}$ is a 2-dim weighted projective space $\C P^2(3,1,2)$. In the $(u_3, v_3, w_3)$-coordinates, Eq.(\ref{4-32}) is written as \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du}{dv} = -(4\theta _\infty- 8 \kappa_0 ) uw + 2uv + 3u^2w^2 + 4\kappa_0(\kappa_0-\theta _\infty) = -\frac{\partial \widetilde{H}_1}{\partial w}, \\[0.2cm] \displaystyle \frac{dw}{dv} = 1-2uw^3-2vw + (2\theta _\infty-4\kappa_0) w^2 = \frac{\partial \widetilde{H}_1}{\partial u}, \\ \end{array} \right. \label{4-34} \end{equation} where the subscripts for $u_3, v_3, w_3$ are omitted for simplicity. This is a Hamiltonian system with the Hamiltonian \begin{equation} \widetilde{H}_1 = u-u^2w^3-2uvw+(2\theta _\infty-4\kappa_0) uw^2-4\kappa_0(\kappa_0-\theta _\infty)w \end{equation} as well as the original ($\text{P}_\text{IV}$). The relation between the original chart $(x, y, z)$ and $(u_3, v_3, w_3)$ is \begin{equation} \left\{ \begin{array}{l} \displaystyle x =w_3^{-1}, \\ \displaystyle y = u_3w_3^2 + 2\kappa_0 w_3, \\ \displaystyle z = v_3. \end{array} \right. \label{4-36} \end{equation} We can verify the equalities \begin{eqnarray*} & & dx \wedge dy = du_3 \wedge w_3, \\ & & dx \wedge dy + dH\wedge dz = du_3 \wedge dw_3 + d \widetilde{H}_1 \wedge dz, \end{eqnarray*} where $z$ is regarded as a parameter in the former relation, and as a coordinate in the latter one. \textbf{(ii)} $(Y_1, Z_1, \varepsilon _1) = (1,0,0)$. In this case, putting $\hat{Y}_1 = Y_1 -1$ for Eq.(\ref{3-33}) yields \begin{eqnarray} \left\{ \begin{array}{l} \displaystyle \dot{\hat{Y}}_1 = 3\hat{Y}_1 + 4Z_1 + (2\kappa_0 - 2\theta _\infty) \varepsilon _1 + f_1(\hat{Y}_1, Z_1, \varepsilon _1), \\ \displaystyle \dot{Z}_1 = Z_1 - \varepsilon _1 + f_2(\hat{Y}_1, Z_1, \varepsilon _1), \\ \displaystyle \dot{\varepsilon }_1 = 2\varepsilon _1, \end{array} \right. \end{eqnarray} where $f_1$ and $f_2$ denote nonlinear terms. After a certain linear transformation $(\hat{Y}_1,Z_1, \varepsilon _1) \mapsto (U_2, V_2, W_2)$, which removes the $(1,2)$ and $(1,3)$-components of the linear part as before, we introduce the weighted blow-up with weights $3,1,2$ by \begin{eqnarray} \left\{ \begin{array}{llll} U_2 & = u_4^3 & = v_5^3 u_5 & = w_6^3 u_6, \\ V_2 & = u_4v_4 & = v_5 & = w_6 v_6, \\ W_2 & = u_4^2w_4 & = v_5^2 w_5 & = w_6^2. \end{array} \right. \end{eqnarray} In the $(u_6, v_6, w_6)$-coordinates, we obtain \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du}{dv} = 3u^2w^2-2uv-4(2\kappa_0-\theta _\infty -2) uw + 4(1-2\kappa_0+\theta _\infty +\kappa_0^2 - \kappa_0 \theta _\infty) = -\frac{\partial \widetilde{H}_2}{\partial w}, \\[0.2cm] \displaystyle \frac{dw}{dv} = -1-2uw^3 + 2vw + 2(2\kappa_0-\theta _\infty -2)w^2 = \frac{\partial \widetilde{H}_2}{\partial u}, \\[0.2cm] \widetilde{H}_2 := -u-u^2w^3 + 2uvw + 2(2\kappa_0-\theta _\infty -2) uw^2 - 4(1-2\kappa_0+\theta _\infty +\kappa_0^2 - \kappa_0 \theta _\infty) w. \end{array} \right. \end{equation} where the subscripts for $u_6, v_6, w_6$ are omitted for simplicity. The relation with the original chart $(x, y, z)$ is \begin{equation} \left\{ \begin{array}{l} \displaystyle x =w_6^{-1}, \\ \displaystyle y =w_6^{-1}+u_6w_6^2-2v_6-2(\kappa_0-\theta _\infty -1)w_6 , \\ \displaystyle z = v_6, \end{array} \right. \label{4-40} \end{equation} which is symplectic as the case (i). \textbf{(iii)} $(X_2, Z_2, \varepsilon _2) = (0,0,0)$. After a certain linear transformation $(X_2, Z_2, \varepsilon _2) \mapsto (U_3, V_3, W_3)$ as before, we introduce the weighted blow-up with weights $3,1,2$ by \begin{eqnarray} \left\{ \begin{array}{llll} U_3 & = u_7^3 & = v_8^3 u_8 & = w_9^3 u_9, \\ V_3 & = u_7v_7 & = v_8 & = w_9 v_9, \\ W_3 & = u_7^2w_7 & = v_8^2 w_8 & = w_9^2. \end{array} \right. \end{eqnarray} In the $(u_9, v_9, w_9)$-coordinates, we obtain \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du}{dv} =3u^2w^2 -2uv -(4\kappa_0- 8\theta _\infty) uw + 4\theta _\infty (\theta _\infty - \kappa_0) = -\frac{\partial \widetilde{H}_3}{\partial w}, \\[0.2cm] \displaystyle \frac{dw}{dv} = 1-2uw^3 + 2vw + (2\kappa_0 - 4\theta _\infty) w^2 = \frac{\partial \widetilde{H}_3}{\partial u}, \\[0.2cm] \widetilde{H}_3 := u-u^2w^3 + 2uvw + (2\kappa_0 - 4\theta _\infty) uw^2 - 4\theta _\infty (\theta _\infty-\kappa_0) w, \end{array} \right. \end{equation} where the subscripts for $u_9, v_9, w_9$ are omitted. The relation with the original chart $(x, y, z)$ is \begin{equation} \left\{ \begin{array}{l} \displaystyle x =u_9w_9^2 + 2\theta _\infty w_9, \\ \displaystyle y =w_9^{-1}, \\ \displaystyle z = v_9, \end{array} \right. \label{4-44} \end{equation} which is symplectic as the case (i). Thus, $\C^2_{(x,y)}, \C^2_{(u_3, w_3)}, \C^2_{(u_6, w_6)}$ and $\C^2_{(u_9, w_9)}$ are glued by above symplectic transformations to define an algebraic surface $E(z)$, which gives the space of initial conditions for ($\text{P}_\text{IV}$). These results are summarized as follows. \\[0.2cm] \textbf{Theorem.\thedef.} \noindent \textbf{(i)} The space of initial conditions $E(z)$ for ($\text{P}_\text{IV}$) is given by $\C^2_{(x,y)} \cup \C^2_{(u_3, w_3)} \cup \C^2_{(u_6, w_6)} \cup \C^2_{(u_9, w_9)}$ glued by (\ref{4-36}), (\ref{4-40}) and (\ref{4-44}). \noindent \textbf{(ii)} These transformations are symplectic. \noindent \textbf{(iii)} Consider a Hamiltonian system (\ref{3-43}) defined on the $(x, y, z)$-coordinates, where $f$ and $g$ are polynomials in $x$ and $y$. If it is also expressed as polynomial Hamiltonian systems in the $(u_3, w_3), (u_6, w_6)$ and $(u_9, w_9)$-coordinates, then (\ref{3-43}) is ($\text{P}_\text{IV}$). \\ Part (iii) is proved in the same way as Thm.4.1 and omitted. \section{Characteristic index} As usual, the weight $(p,q,r,s)$ denotes $(3,2,4,5)$, $(2,1,2,3)$ and $(1,1,1,2)$ for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$), respectively. Recall the decomposition \begin{eqnarray*} \C P^3(p,q,r,s) = \C^3 / \Z_s \cup \C P^2(p,q,r). \end{eqnarray*} The Painlev\'{e} equation is defined on the covering space of $ \C^3 / \Z_s$, and $\C P^2(p,q,r)$ is attached at infinity. We have seen that there are fixed points of the vector field on $\C P^2(p,q,r)$ which correspond to movable singularities. In Sec.3 and 4, the eigenvalues $(\lambda _1, \lambda _2, \lambda _3)$ of the Jacobi matrices at the fixed points play an important role, where $(\lambda _1, \lambda _2, \lambda_3) = (6,4,5), (4,2,3)$ and $(3,1,2)$ for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$), respectively. In Sec.3, they are used to apply the Poincar\'{e} linearization theorem. In Sec.4, they determine the weight of the weighted blow-up. We call the eigenvalues $(\lambda _1, \lambda _2, \lambda _3)$ the characteristic index for ($\text{P}_\text{J}$), see Table.2. Obviously, they are invariant under the automorphism on $\C P^3(p,q,r,s)$. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|} \hline & $\C P^3(p,q,r,s)$ & $\lambda _1, \lambda _2, \lambda _3$ & $(\lambda _1 + \lambda _2)/\lambda _3$ & $\kappa$ \\ \hline \hline ($\text{P}_\text{I}$) & $\C P^3(3,2,4,5)$ & $6,4,5$ & 2 & 6 \\ \hline ($\text{P}_\text{II}$) & $\C P^3(2,1,2,3)$ & $4,2,3$ & 2 & 4 \\ \hline ($\text{P}_\text{IV}$) & $\C P^3(1,1,1,2)$ & $3,1,2$ & 2 & 3 \\ \hline \end{tabular} \end{center} \caption{The characteristic index $(\lambda _1, \lambda _2, \lambda _3)$ and the Kovalevskaya exponent $\kappa$.} \end{table} We observe the following properties: \\[0.2cm] \textbf{(i)} $r = \lambda _2$ and $s = \lambda _3$. \\ \textbf{(ii)} $\lambda _1 = \kappa = \mathrm{deg} (H_\text{J}) = s+1$, where $\kappa$ is the Kovalevskaya exponent and $\mathrm{deg} (H_\text{J})$ is a weighted degree of the Hamiltonian given in Sec.2.1. \\ \textbf{(iii)} $(\lambda _1 + \lambda _2)/\lambda _3$ is an integer. \\ \textbf{(iv)} $p+q = s$. \\ In the forthcoming paper \cite{Chi4}, the properties (i) and (ii) are proved for a general $m$-dimensional system satisfying certain conditions on the Newton diagram; two numbers in the characteristic index coincide with $r$ and $s$ determined by the Newton diagram, and the others coincide with the Kovalevskaya exponents. Part (iii) and (iv) are related to the following proposition. \\[0.2cm] \textbf{Proposition.\thedef.} For ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$), the symplectic form $dx \wedge dy + dH_J \wedge dz$ is a rational form on $\C P^3(p,q,r,s)$. This can be proved by a straightforward calculation. For example, on the $(Y_1, Z_1, \varepsilon _1)$ chart, we have \begin{eqnarray*} dx \wedge dy = d (\varepsilon _1^{-p/s}) \wedge d(Y_1\varepsilon _1^{-q/s}). \end{eqnarray*} In order for it to be rational in $\varepsilon _1$, $p+q$ should be a multiple of $s$, and this is true for ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$). Similarly, we can verify that $dH \wedge dz$ is rational, which is a consequence of Part (iii) above. The above properties (i) to (iv) are true for higher order first and second Painlev\'{e} equations. For example, the characteristic indices and the Kovalevskaya exponents for the fourth order and sixth order first Painlev\'{e} equations, $(\text{P}_\text{I})_4$ and $(\text{P}_\text{I})_6$, from the first Painlev\'{e} hierarchy $(\text{P}_\text{I})_{2n}$, and the fourth order second Painlev\'{e} equation $(\text{P}_\text{II})_4$ from the second Painlev\'{e} hierarchy $(\text{P}_\text{II})_{2n}$ are shown in Table 3. In Table 3, the equality $p_j + q_j = s$ holds and $\nu := (\lambda _1+ \cdots + \lambda _{2n})/\lambda _{2n+1}$ is an integer. Furthermore, $r=\lambda _{2n}, s=\lambda _{2n+1}$ and the Kovalevskaya exponents coincide with $\lambda _1, \cdots ,\lambda _{2n-1}$. See \cite{Chi4} for the detail. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|} \hline $(\text{P}_\text{J})_{2n}$ & $\C P^{2n+1}(p_1, q_1, \cdots , p_n, q_n, r,s)$ & $\lambda _1, \cdots , \lambda _{2n+1}$ & $\nu$ & $\kappa$ \\ \hline \hline $(\text{P}_\text{I})_4$ & $\C P^5(5,2,3,4,6,7)$ & $8,5,2,6,7$ & 3 & 8,5,2 \\ \hline $(\text{P}_\text{I})_6$ & $\C P^7(7,2,5,4,3,6,8,9)$ & $10,7,5,4,2,8,9$ & 4 & 10,7,5,4,2 \\ \hline $(\text{P}_\text{II})_4$ & $\C P^5(2,3,4,1,4,5)$ & $6,3,2,4,5$ & 3 & 6,3,2 \\ \hline \end{tabular} \end{center} \caption{The characteristic index $(\lambda _1, \cdots , \lambda _{2n+1})$, the Kovalevskaya exponent $\kappa$ and $\nu := (\lambda _1+ \cdots + \lambda _{2n})/\lambda _{2n+1}$.} \end{table} \section{The Boutroux coordinates} In his celebrated paper, Boutroux \cite{Bou} introduced the coordinate transformations $y = u t^{2/5},\, z = t^{4/5}$ for ($\text{P}_\text{I}$), and $y = u t^{1/3},\, z = t^{2/3}$ for ($\text{P}_\text{II}$) to investigate the asymptotic behavior of solutions around the essential singularity $z = \infty$. His coordinate transformations are essentially the same as the third local chart of $\C P^3(p,q,r,s)$, see Eq.(\ref{3-1}) and (\ref{3-17}), and put $\varepsilon _3 = 1/t$. Hence, we call the third local chart $(X_3, Y_3, \varepsilon _3)$ of $\C P^3(p,q,r,s)$ the Boutroux coordinates even for ($\text{P}_\text{IV}$). On the Boutroux coordinates, ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) expressed as the autonomous vector fields are given by Eq.(\ref{3-7}), (\ref{3-22}) and (\ref{3-35}), respectively. Recall that the set $\C P^2(p,q,r) = \{ \varepsilon _1 = 0\} \cup \{ \varepsilon _2 = 0\}\cup \{ \varepsilon _3 = 0\}$ is attached at infinity of the original chart $(x,y,z)$. In particular, the set $\{ \varepsilon _3 \, ( = z^{-s/r}) = 0 \}$ describes the behavior around the irregular singular point $z = 0$. Putting $\varepsilon _3 = 0$ in Eq.(\ref{3-7}), (\ref{3-22}) and (\ref{3-35}), we obtain \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_3 =24Y_3^2 + 4 , \\ \displaystyle \dot{Y}_3 = 4X_3 , \\ \end{array} \right. \label{Bou1} \end{equation} \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_3 =4Y_3^3 + 2Y_3, \\ \displaystyle \dot{Y}_3 = 2X_3, \\ \end{array} \right. \label{Bou2} \end{equation} and \begin{equation} \left\{ \begin{array}{l} \displaystyle \dot{X}_3 = -X_3^2+2X_3Y_3+2X_3, \\ \displaystyle \dot{Y}_3 = -Y_3^2+2X_3Y_3-2Y_3, \\ \end{array} \right. \label{Bou4} \end{equation} respectively. It is remarkable that they are autonomous Hamiltonian systems with the Hamiltonian functions $\mathcal{H}_J$ shown in Table 4. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|} \hline & space & $\mathcal{H}_J$ & symmetry \\ \hline \hline ($\text{P}_\text{I}$) & $\C P^3(3,2,4,5)$ & $2X_3^2 - 8Y_3^3 - 4Y_3$ & $\Z_4$ \\ \hline ($\text{P}_\text{II}$) & $\C P^3(2,1,2,3)$ & $X_3^2 - Y_3^4 - Y_3^2$ & $\Z_2\times \Z_2$ \\ \hline ($\text{P}_\text{IV}$) & $\C P^3(1,1,1,2)$ & $X_3^2Y_3 - X_3Y_3^2 - 2X_3Y_3$ & $\mathfrak{S}_3$ \\ \hline \end{tabular} \end{center} \caption{Hamiltonian functions defined on the set $\{ \varepsilon _3 = 0\}$. See Section 7 for the symmetry.} \end{table} Therefore, each leaf of the foliation on the set $\{ \varepsilon _3 = 0\} \subset \C P^2(p,q,r)$ is determined by the level set $\{ \mathcal{H}_J = c\},\, c\in \C $ of the Hamiltonian, which is an elliptic curve for a generic value of $c\in \C$. In particular, $\mathcal{H}_I = c$ is the Weierstrass form and $\mathcal{H}_{II} = c$ is the Jacobi form. In what follows, we will see that the space of initial conditions written in the Boutroux coordinates has been already constructed by the weighted blow-ups introduced in Sec.4. \subsection{The first Painlev\'{e} equation} We have proved that $(x, y)$-space and $(u_3, w_3)$-space are glued to give the space of initial conditions. In this subsection, it is shown that $(X_3, Y_3)$-space and $(u_2, v_2)$-space give the space of initial conditions for ($\text{P}_\text{I}$) written in the Boutroux coordinates. On the $(X_3, Y_3, \varepsilon _3)$-coordinates, ($\text{P}_\text{I}$) is given as \begin{equation} \frac{dX_3}{d\varepsilon _3} = \frac{1}{5\varepsilon _3^2} \left( -24Y_3^2 - 4 + 3X_3\varepsilon _3 \right), \quad \frac{dY_3}{d\varepsilon _3} = \frac{1}{5\varepsilon _3^2} \left( -4X_3 + 2Y_3\varepsilon _3\right) . \label{5-1} \end{equation} It has an irregular singular point $\varepsilon _3 = 0$. Since $z = \varepsilon _3^{-4/5}$, this equation also has the Painlev\'{e} property with a possible branch point $\varepsilon _3 = 0$. Recall that the $(u_2, v_2, w_2)$-coordinates are defined by (\ref{4-3}). In these coordinates, ($\text{P}_\text{I}$) is written as \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du_2}{dw_2} = \frac{1}{5w_2^2} \left( -\frac{1}{2}v_2 - 6u_2^2v_2^5 + 6u_2w_2 \mp \frac{3}{2}v_2^2w_2 + 5u_2 v_2^4w_2 \pm 4u_2v_2^3 - v_2^3w_2^2 \right), \\ \displaystyle \frac{dv_2}{dw_2} = \frac{1}{5w_2^2}\left( \mp 4 \mp v_2^4 + 2u_2v_2^6 - v_2 w_2 - v_2^5 w_2 \right) . \end{array} \right. \label{5-2} \end{equation} This is a polynomial ODE with an irregular singular point $w_2 = 0$. The relation between two charts are \begin{equation} \left\{ \begin{array}{l} \displaystyle X_3 = v_2^{-3} (v_2^6 u_2 \mp \frac{1}{2}v_2^4 - \frac{1}{2}v_2^5 w_2 \mp 2), \\ \displaystyle Y_3 = v_2^{-2}, \\ \displaystyle \varepsilon _3 = w_2, \end{array} \right. \label{5-3} \end{equation} or \begin{equation} \left\{ \begin{array}{l} \displaystyle u_2 = X_3Y_3^{3/2} + \frac{1}{2}Y_3^{1/2} \varepsilon _3 \pm \frac{1}{2}Y_3 \pm 2Y_3^3, \\ \displaystyle v_2 = Y_3^{-1/2}, \\ \displaystyle w_2 = \varepsilon _3. \end{array} \right. \label{5-4} \end{equation} We should divide the $(u_2, v_2, w_2)$-space by the $\Z_2$ action as before. The $\Z_2$ action induced from the orbifold structure is given by \begin{equation} \left( \begin{array}{@{\,}c@{\,}} u_2 \\ v_2 \\ w_2 \end{array} \right) \mapsto \left( \begin{array}{@{\,}c@{\,}} -u_2 + v_2^{-2} + 4v_2^{-6} \\ -v_2 \\ w_2 \end{array} \right) . \end{equation} We define invariants of this action to be \begin{equation} \left\{ \begin{array}{l} \displaystyle U = u_2 (u_2 v_2^6 - v_2^4 - 4) + \frac{1}{4}v_2^2 \\ \displaystyle V = v_2^7( u_2 - \frac{1}{2} v_2^{-2} - 2v_2^{-6}) \\ \displaystyle W = v_2^2. \end{array} \right. \end{equation} This defines a nonsingular algebraic surface $M = \C ^2_{(u_2, v_2)} / \Z_2$ \begin{equation} M : V^2 = UW^4 + 2W^3 + 4W. \end{equation} Note that it is independent of a parameter. Eqs.(\ref{5-2}) to (\ref{5-4}) are rewritten as \begin{equation} \left\{ \begin{array}{l} \displaystyle X_3 = W^{-2}V - \frac{1}{2}W\varepsilon _3 \\ \displaystyle Y_3 = W^{-1}, \end{array} \right. \quad \left\{ \begin{array}{l} \displaystyle V = X_3Y_3^{-2}+\frac{1}{2}Y_3^{-3}\varepsilon _3 \\ \displaystyle W = Y_3^{-1}, \end{array} \right. \end{equation} and \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{dV}{d\varepsilon _3} = \frac{1}{5\varepsilon _3^2} \left( 8 \frac{V^2}{W} - \varepsilon _3 V - 2\varepsilon _3 VW^2 - 4W^2 - \varepsilon _2^2 W^5 - 24 \right), \\ \displaystyle \frac{dW}{d\varepsilon _3} = \frac{1}{5\varepsilon _3^2} \left( -2W\varepsilon _3 + 4V - 2W^3 \varepsilon _3 \right), \end{array} \right. \label{5-9} \end{equation} respectively. Hence, $\C^2_{(X_3, Y_3)} \cup M$ gives the space of initial conditions for the Boutroux coordinates. Note that Eqs.(\ref{5-1}) and (\ref{5-9}) are not Hamiltonian systems, though they are reduced to Hamiltonian systems as $\varepsilon _3 \to 0$. \subsection{The second Painlev\'{e} equation} On the $(X_3, Y_3, \varepsilon _3)$ chart, ($\text{P}_\text{II}$) is written as \begin{equation} \frac{dX_3}{d\varepsilon _3} = \frac{4Y_3^3 + 2Y_3 + 2\alpha \varepsilon _3 - 2X_3\varepsilon _3}{-3\varepsilon _3^2}, \quad \frac{dY_3}{d\varepsilon _3} = \frac{2X_3 - Y_3\varepsilon _3}{-3\varepsilon _3^2}. \label{5-10} \end{equation} The space of initial conditions for this system is also obtained by the weighted blow up as follows: Recall that $(u_2^{\pm},v_2^{\pm},w_2^{\pm})$ is defined by (\ref{4-22}). The relation between $(X_3,Y_3, \varepsilon _3)$ and $(u_2^{\pm},v_2^{\pm},w_2^{\pm})$ is given by \begin{equation} \left\{ \begin{array}{l} \displaystyle u_2^{\pm} = X_3Y_3^2 \pm Y_3^4 \pm \frac{1}{2} Y_3^2 + (\frac{1}{2} \pm \alpha ) \varepsilon _3 Y_3 \\ v_2^{\pm} = Y_3^{-1} \\ w_2^{\pm} = \varepsilon _3. \end{array} \right. \label{5-11} \end{equation} It is remarkable that the independent variable $\varepsilon _3$ is not changed. The equation written in $(u_2^{\pm},v_2^{\pm},\varepsilon _3)$ is \begin{equation} \left\{ \begin{array}{l} \displaystyle \frac{du}{d\varepsilon }= \frac{-1}{3\varepsilon ^2} \left( 4u^2v^3 \mp 2uv \pm (\frac{1}{2} \pm \alpha ) \varepsilon -6 (\frac{1}{2} \pm \alpha ) \varepsilon uv^2 -4 \varepsilon u + \frac{\varepsilon ^2 v}{2} \pm 2 \alpha \varepsilon ^2v + 2\alpha ^2 \varepsilon ^2v \right) \\[0.4cm] \displaystyle \frac{dv}{d\varepsilon }= \frac{-1}{3\varepsilon ^2} \left( \pm 2 \pm v^2- 2uv^4 + \varepsilon v + \varepsilon v^3 \pm 2\alpha \varepsilon v^3 \right), \\ \end{array} \right. \label{5-12} \end{equation} where the subscript and the superscript for $u_2^{\pm},v_2^{\pm}, \varepsilon _3$ are omitted. Since this equation is polynomial in $u$ and $v$, the space of initial conditions is obtained by glueing $\C^2_{(X_3, Y_3)}$, $\C^2_{(u_2^+, v_2^+)}$ and $\C^2_{(u_2^-, v_2^-)}$ by the relation (\ref{5-11}). \subsection{The fourth Painlev\'{e} equation} We call the third local chart $(X_3, Y_3, \varepsilon _3)$ for ($\text{P}_\text{IV}$) the Boutroux coordinates as in the first and second Painlev\'{e} equations. In this chart, ($\text{P}_\text{IV}$) is written as \begin{equation} \left\{ \begin{array}{ll} \displaystyle \frac{dX_3}{d\varepsilon _3} = \frac{-X_3^2+2X_3Y_3+2X_3-2\theta _\infty \varepsilon _3-X_3\varepsilon _3}{-2\varepsilon _3^2} \\[0.4cm] \displaystyle \frac{dY_3}{d\varepsilon _3} = \frac{-Y_3^2+2X_3Y_3-2Y_3-2\kappa_0 \varepsilon _3-Y_3\varepsilon _3}{-2\varepsilon _3^2}. \\ \end{array} \right. \end{equation} In Sec.4.3, $(u_2, v_2, w_2), (u_5, v_5, w_5)$ and $(u_8, v_8, w_8)$ coordinates are defined through the weighted blow-ups. The relations between them and $(X_3,Y_3, \varepsilon _3)$ are given by \begin{equation} \left\{ \begin{array}{ll} X_3 = v_2^{-1} & = v_5^{-1} \\ Y_3 = u_2v_2^2 + 2\kappa_0 w_2 v_2 & = u_5v_5^2 + 2(1 + \theta _\infty - \kappa_0) w_5v_5 + v_5^{-1} -2 \\ \varepsilon _3 = w_2 & = w_5 , \end{array} \right. \end{equation} and \begin{equation} \left\{ \begin{array}{ll} X_3 = u_8v_8^2 + 2\theta _\infty w_8 v_8 \\ Y_3 = v_8^{-1} \\ \varepsilon _3 = w_8 \end{array} \right. \end{equation} In particular, the independent variable $\varepsilon _3$ is not changed. The equations written in $(u_2,v_2, w_2), (u_5, v_5, w_5)$ and $(u_8,v_8,w_8)$ are polynomial in $(u_2,v_2), (u_5, v_5)$ and $(u_8,v_8)$, respectively. Thus, the space of initial conditions is obtained by glueing $\C^2_{(X_3, Y_3)}, \C^2_{(u_2, v_2)}, \C^2_{(u_5, v_5)}$ and $\C^2_{(u_8, v_8)}$ by the above relation. \section{The extended affine Weyl group} It is known that there exists a group of rational transformations acting on $\C^3$ which changes the Painlev\'{e} equation to another Painlev\'{e} equation of the same type with different parameters. The transformation group is isomorphic to the extended affine Weyl group. See \cite{Tsu} for the complete list of the actions of the groups for the second to the sixth Painlev\'{e} equations written in Hamiltonian forms. In this section, we study the actions of the extended affine Weyl groups for ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$). For a classical root system $R$, the affine Weyl group and the extended affine Weyl group are denoted by $W(R^{(1)})$ and $\widetilde{W}(R^{(1)})$, respectively. Let $G = \mathrm{Aut} (R^{(1)})$ be the Dynkin automorphism group of the extended Dynkin diagram. We have $\widetilde{W}(R^{(1)}) \cong G \ltimes W(R^{(1)})$. For the second Painlev\'{e} equation, \begin{eqnarray*} & & \widetilde{W}(A_1^{(1)}) \cong G \ltimes W(A_1^{(1)}) = \langle s_1, \pi \rangle, \\ & & G = \mathrm{Aut} (A_1^{(1)}) = \langle \pi \rangle \cong \Z_2. \end{eqnarray*} The action of the group is given in Table 5. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|} \hline & $\alpha$ & $x$ & $y$ \\ \hline \hline $s_1$ & $-\alpha +1$ & $\displaystyle x+\frac{(2\alpha -1)y}{y^2-x+z/2} + \frac{(\alpha -1/2)^2}{(y^2-x+z/2)^2}$ & $\displaystyle y+ \frac{\alpha -1/2}{y^2-x+z/2}$ \\ \hline $\pi$ & $-\alpha $ & $-x$ & $-y$ \\ \hline \end{tabular} \end{center} \caption{The action of the extended affine Weyl group for ($\text{P}_\text{II}$) .} \end{table} For the fourth Painlev\'{e} equation, \begin{eqnarray*} & & \widetilde{W}(A_2^{(1)}) \cong G \ltimes W(A_2^{(1)}) = \langle s_0, s_1, s_2, \sigma_1, \sigma_2 \rangle, \\ & & G = \mathrm{Aut} (A_2^{(1)}) = \langle \sigma_1, \sigma_2 \rangle \cong \mathfrak{S}_3. \end{eqnarray*} The action of the group is given in Table 6. \begin{table}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|c|} \hline & $\kappa_0$ & $\theta _\infty$ & $x$ & $y$ & $z$ \\ \hline \hline $s_0$ & $1+\theta _\infty$ & $\kappa_0-1$ & $\displaystyle x+\frac{2(1-\kappa_0+\theta _\infty)}{x-y-2z}$ & $\displaystyle y+\frac{2(1-\kappa_0+\theta _\infty)}{x-y-2z}$ & $z$ \\ \hline $s_1$ & $-\kappa_0$ & $\theta _\infty-\kappa_0$ & $x-2\kappa_0/y$ & $y$ & $z$ \\ \hline $s_2$ & $\kappa_0 - \theta _\infty$ & $-\theta _\infty$ & $x$ & $y-2\theta _\infty/x$ & $z$ \\ \hline $\pi$ & $-\theta _\infty$ & $\kappa_0-\theta _\infty -1 $ & $-x+y+2z$ & $-x$ & $z$ \\ \hline $\sigma_1$ & $-\theta _\infty$ & $-\kappa_0$ & $-iy$ & $-ix$ & $iz$ \\ \hline $\sigma_2$ & $\kappa_0$ & $\kappa_0 - \theta _\infty - 1$ & $ix$ & $i(x-y-2z)$ & $iz$ \\ \hline \end{tabular} \end{center} \caption{The action of the extended affine Weyl group for ($\text{P}_\text{IV}$) .} \end{table} In the next theorem, $R^{(1)}$ and $(p,q,r,s)$ denote $A^{(1)}_1$ and $(2,1,2,3)$ for ($\text{P}_\text{II}$), and $A^{(1)}_2$ and $(1,1,1,2)$ for ($\text{P}_\text{IV}$), respectively. \\[0.2cm] \textbf{Theorem.\thedef.} (i) The transformation group $\widetilde{W}(R^{(1)})$ given above is extended to a rational transformation group on $\C P^3(p,q,r,s)$. (ii) For each element $s$ in $W(R^{(1)})$, the action of $s$ on the infinity set $\C P^2(p,q,r)$ is trivial: $s|_{\C P^2(p,q,r)} = \mathrm{id}$. Hence, the transformation group $\widetilde{W}(R^{(1)}) \cong \mathrm{Aut} (R^{(1)}) \ltimes W(R^{(1)})$ is reduced to $\mathrm{Aut} (R^{(1)})$ on $\C P^2(p,q,r)$. (iii) The foliation on $\C P^2(p,q,r)$ is $\mathrm{Aut} (R^{(1)})$-invariant. \\[0.2cm] \textbf{Proof.} We give a proof for ($\text{P}_\text{II}$). A proof for ($\text{P}_\text{IV}$) is done in the same way. (i) The original chart $(x,y,z)$ for ($\text{P}_\text{II}$) has to be divided by the $\Z_3$-action $(x,y,z)\mapsto (\omega ^2 x, \omega y, \omega ^2z),\, \omega :=e^{2\pi i/3}$ because of the orbifold structure, see Eq.(\ref{2-5}). It is easy to verify that the actions of the generators $s_1$ and $\pi$ of $\widetilde{W}(A_1^{(1)})$ shown in Table 5 commute with the $\Z_3$-action, so that they induce actions on the quotient space $\C^3 / \Z_3$. Further, these actions are extended to the whole space $\C P^3(2,1,2,3)$. For example, the action of $s_1$ on the third local chart is expressed by \begin{equation} (X_3, Y_3,\varepsilon _3) \mapsto (X_3 + \frac{(2\alpha -1)Y_3\varepsilon _3}{Y_3^2-X_3+1/2} +\frac{(\alpha -1/2)^2 \varepsilon _3^2}{(Y_3^2-X_3+1/2)^2}, Y_3 + \frac{(\alpha -1/2)\varepsilon _3}{Y_3^2-X_3+1/2}, \varepsilon _3 ). \label{7-1} \end{equation} This is rational and commutes with the action (\ref{3-19}) defining the orbifold structure. It is straightforward to calculate the action on the other charts $(Y_1, Z_1, \varepsilon _1)$ and $(X_2, Z_2, \varepsilon _2)$, which proves that the action $s_1$ is well defined on $\C P^3(2,1,2,3)$. The same is also true for the action of $\pi$. (ii) The set $\C P^2 (2,1,2)$ is given by $\{ \varepsilon _1 = 0\} \cup \{ \varepsilon _2 = 0\} \cup\{ \varepsilon _3 = 0\}$. The action (\ref{7-1}) is reduced to the trivial action as $\varepsilon _3 \to 0$. Similarly, the action of $s_1$ on the other charts $(Y_1, Z_1, \varepsilon _1)$ and $(X_2, Z_2, \varepsilon _2)$ becomes trivial as $\varepsilon _1 \to 0$ and $\varepsilon _2 \to 0$. On the other hand, the action of $\pi$ on $\C P^2 (2,1,2)$ is not trivial. For example, the action of $\pi$ on $(X_3, Y_3,\varepsilon _3)$ is $(X_3, Y_3,\varepsilon _3)\mapsto (-X_3, -Y_3,\varepsilon _3)$. (iii) The action of $\widetilde{W}(A_1^{(1)})$ transforms ($\text{P}_\text{II}$) into ($\text{P}_\text{II}$) with a different parameter. However, the foliation on $\C P^2 (2,1,2)$ is independent of the parameter $\alpha $, see Eq.(\ref{Bou2}). Thus, the foliation on $\C P^2 (2,1,2)$ is $\mathrm{Aut} (A_1^{(1)})$-invariant. $\Box$ \\ In Table 4 in Sec.6, the symmetry groups of the foliations on $\C P^2 (p,q,r)$ generated by ($\text{P}_\text{I}$), ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$) are shown. The foliation generated by $\mathcal{H}_I$ is invariant under the $\Z_4$-action (\ref{3-4}) which arises from the orbifold structure. The foliation generated by $\mathcal{H}_{II}$ is invariant under the $\Z_2$-action (\ref{3-19}), and $\mathrm{Aut} (A_1^{(1)}) \cong \Z_2$ given by $(X_3, Y_3)\mapsto (-X_3, -Y_3)$. Similarly, the foliation generated by $\mathcal{H}_{IV}$ is invariant under the action of $\mathrm{Aut} (A_2^{(1)}) \cong \mathfrak{S}_3$, while there are no symmetry induced from the orbifold structure. The foliation $\{ \mathcal{H}_{IV} = c\}$ for real $c \in \R$ is represented in Fig.1. Note that $\mathrm{Aut} (A_2^{(1)})$ is isomorphic to the dihedral group $D_3$ of a triangle. In Fig.1, the zero level set $\mathcal{H}_{IV} = 0$ consists of three lines, which creates a triangle. $\mathrm{Aut} (A_2^{(1)})$ acts on the triangle as the dihedral group $D_3$. \begin{figure} \begin{center} \includegraphics[]{fig1-1.eps} \caption[]{The foliation $\{ \mathcal{H}_{IV} = c\}$ for $c \in \R$.} \label{fig1} \end{center} \end{figure} \section{Cellular decomposition and Dynkin diagrams} In this section, the weighted blow-up of $\C P^3(3,2,4,5)$ with weights $6,4,5$ defined in Sec.4.1 is called the total space for ($\text{P}_\text{I}$) and denoted by $\mathcal{M}_I$. We calculate the cellular decomposition of it. It will be shown that $\mathcal{M}_I$ is decomposed into the fiber space $\mathcal{P} = \{ (x,z) \, | \, x\in E(z), z\in \C \}$ for ($\text{P}_\text{I}$), an elliptic fibration over the moduli space of complex tori defined by the Weierstrass equation and a projective line. We also shows that the extended Dynkin diagram of type $\tilde{E}_8$ is hidden in the space $\mathcal{M}_I$. \subsection{The elliptic fibration} Let us calculate a cellular decomposition of $\mathcal{M}_I$. $\C P^3(3,2,4,5)$ is decomposed as (\ref{2-7}). Furthermore, $\C P^2(3,2,4)$ is decomposed as $\C P^2(3,2,4) = \C^2/\Z_4 \cup \C P^1(3,2)$. Since $\C P^1(3,2)$ is isomorphic to the Riemann sphere, we obtain \begin{eqnarray*} \C P^3(3,2,4,5) &=& \C^3 / \Z_5 \,\, \cup \,\, \C^2 / \Z_4 \,\, \cup \,\, \C \,\, \cup \,\, \{ p\}, \end{eqnarray*} where $\{ p\}$ denotes the point $(X_2, Z_2, \varepsilon _2) = (2,0,0)$, at which the weighted blow-up was performed. In local coordinates, $\C^3 / \Z_5$ is the $(x, y, z)$-space divided by the $\Z_5$ action, and $\C^2 / \Z_4$ is the set $\{ (X_3, Y_3, \varepsilon _3) \, | \, \varepsilon _3 = 0\}$ divided by the $\Z_4$ action. On the set $\{ (X_3, Y_3, \varepsilon _3) \, | \, \varepsilon _3 = 0\}$, the foliation is defined by the Hamiltonian system (\ref{Bou1}) with the Hamiltonian function $H = 2X^2 - 8Y^3 - 4Y$. This equation is actually invariant under the $\Z_4$ action given by (\ref{3-4}). Next, due to the definition of the weighted blow-up, we have \begin{eqnarray*} & & \mathcal{M}_I = \C^3 / \Z_5 \,\, \cup \C^2 / \Z_4 \,\, \cup \,\, \C \,\, \cup \,\, \C P^2(6,4,5). \end{eqnarray*} Since $\C P^2(6,4,5) = \C^2 / \Z_5 \cup \C P^1(6,4)$, we obtain \begin{eqnarray} = & & \,\, \C^3 / \Z_5 \quad \cup \quad\quad \C^2 / \Z_4 \quad\quad\, \cup \quad \C \nonumber \\ &\cup & \,\, \C^2 / \Z_5 \quad \cup \,\, \C P^1(6,4)\backslash \{ q\} \,\, \cup \,\,\,\, \{ q\}, \label{7-2} \end{eqnarray} where $\{ q\}$ denotes the point $(u_1, v_1, w_1) = (0,0,0)$. In local coordinates, $\C^2 / \Z_5$ is given as $\{ (u_3, v_3, w_3) \, | \, w_3 = 0\}$ divided by $\Z_5$. This implies that the first column $\C^3 / \Z_5 \cup \C^2 / \Z_5$ is just the fiber space $\mathcal{P}$ for ($\text{P}_\text{I}$) divided by the $\Z_5$ action, the fiber space over $\C$ whose fiber is the space of initial conditions. The last column $\C \cup \{ q\}$ is the Riemann sphere. Let us investigate the second column $\C^2 / \Z_4 \cup \C P^1(6,4)\backslash \{ q\}$. On the space $\C^2 / \Z_4$, the equation (\ref{Bou1}) divided by the $\Z_4$ action is defined, and $\C P^1(6,4)$ is attached at ``infinity". Note that each integral curve $X^2 = 4Y^3 + 2Y - g_3$ of (\ref{Bou1}) defines an elliptic curve, where $g_3$ is an integral constant; compare with the Weierstrass normal form $X^2 = 4Y^3 - g_2 Y - g_3$. The Weierstrass normal form defines a complex torus when $g_2^3 - 27 g_3^2 \neq 0$. Two complex tori defined by $(g_2, g_3)$ and $(g_2', g_3')$ are isomorphic to one another if there is $\lambda \neq 0$ such that $(g_2, g_3) = (\lambda ^4 g'_2, \lambda ^6 g'_3)$. Hence, $\C P^1(6,4)\backslash \{ \text{one point}\}$ is a moduli space of complex tori. According to Eq.(\ref{Bou1}), the $(X_3,Y_3)$-space is foliated by a family of elliptic curves (including two singular curves $g_3 = \pm (8/27)^{1/2}$) defined by the Weierstrass normal form $X^2 = 4Y^3 + 2Y - g_3$. By the $\Z_4$ action $(X_3, Y_3) \mapsto (iX_3, -Y_3)$, the normal form is mapped to $X^2 = 4Y^3 + 2Y + g_3$. This means that by the $\Z_4$ action induced from the orbifold structure, two elliptic curves having parameters $(-2, g_3)$ and $(-2, -g_3)$ are identified. However, the equality $(-2, g_3 ) = (-2 \lambda ^4, g'_3 \lambda ^6)$ holds for some $\lambda $ if and only if $g_3 = g_3'$ or $g'_3 = -g_3$. This proves that two elliptic curves identified by the $\Z_4$ action are isomorphic with one another, and $\C^2 / \Z_4$ is foliated by isomorphism classes of elliptic curves (including a singular one, while the case $g_2 = 0$ is excluded). The set $\C P^1(6,4)\backslash \{ q\}$ is expressed as $\{ (u_2, v_2, w_2) \, | \, v_2 = w_2 = 0\}$ divided by the $\Z_2$ action $u_2 \mapsto -u_2$. We can show that each isomorphism class of an elliptic curve $X^2 = 4Y^3 + 2Y \mp g_3$ intersects with the moduli space $\C P^1(6,4)\backslash \{ q\}$ at the point $(u_2, v_2, w_2) = (g_3/4, 0,0) \sim (-g_3/4,0,0)$. This proves that the second column of (\ref{7-2}) gives an elliptic fibration whose base space is the moduli space $\C P^1(6,4)\backslash \{ q\}$ and fibers are isomorphism classes of elliptic curves including the singular curve, but excluding the curve of $g_2 = 0$. \\[0.2cm] \textbf{Theorem.\thedef.} The total space $\mathcal{M}_I$ is decomposed into the disjoint union of the fiber space for ($\text{P}_\text{I}$) divided by $\Z_5$, an elliptic fibration obtained from the Weierstrass normal form as above, and $\C P^1$. \\ Similar results also hold for ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$), for which elliptic curves are not defined by the Weierstrass form but Hamiltonians represented in Table 4. \subsection{The extended Dynkin diagram} It is known that an extended Dynkin diagram is associated with each Painlev\'{e} equation (Sakai \cite{Sak}). For example, the diagram of type $\tilde{E}_8$ is associated with ($\text{P}_\text{I}$). Okamoto obtained $\tilde{E}_8$ as follows: In order to construct the space of initial conditions of ($\text{P}_\text{I}$), he performed blow-ups eight times to a Hirzebruch surface. After that, vertical leaves, which are the pole divisor of the symplectic form, are removed. The configuration of irreducible components of the vertical leaves is described by the Dynkin diagram of type $\tilde{E}_8$, see Fig.\ref{fig3}(a). Our purpose is to find the diagram $\tilde{E}_8$ hidden in the total space $\mathcal{M}_I$. Recall that $\mathcal{M}_I$ is covered by seven local coordinates; the inhomogeneous coordinates (\ref{3-1}) of $\C P^3(3,2,4,5)$ and $(u_i, v_i, w_i)$ defined by (\ref{4-3}). These local coordinates should be divided by the suitable actions due to the orbifold structure. The actions on $(Y_1, Z_1, \varepsilon _1)$, $(X_2, Z_2, \varepsilon _2)$ and $(X_3,Y_3,\varepsilon _3)$ are listed in Eqs.(\ref{3-2}) to (\ref{3-4}). The action on $(u_1, v_1, w_1)$ is given by \begin{eqnarray} (u_1, v_1, w_1) \mapsto (\zeta u_1, \zeta^2 v_1, \zeta w_1), \quad \zeta = e^{2 \pi i/6}. \end{eqnarray} Each fiber (the space of initial conditions) for ($\text{P}_\text{I}$) is not invariant under the $\Z_5$ action (\ref{1-4}) except for the fiber on $z = 0$. Hence, we consider the closure of the fiber on $z = 0$ in $\mathcal{M}_I$. The closure is a 2-dim orbifold expressed as \begin{equation} N:= \{ (x, y, 0)\} \cup \{(Y_1 ,0, \varepsilon _1)\} \cup \{ (X_2, 0, \varepsilon _2) \} \cup \{ (u_3, 0, w_3)\} \cup \{ (u_1, 0, w_1)\}. \end{equation} $N$ is a compactification of the space of initial conditions $E(0) = \{ (x, y, 0)\} \cup \{ (u_3, 0, w_3)\}$ obtained by attaching a 1-dim space \begin{equation} D:= \{ (Y_1,0,0)\} \cup \{(X_2 ,0,0)\, | \, X_2 \neq \pm 2 \} \cup \{ (u_1, 0,0) \}, \quad X_2 = Y_1^{-3/2} = u_1^6 \mp 2. \end{equation} $N$ has three orbifold singularities on $D$ given by \begin{equation} (Y_1, \varepsilon _1) = (0,0), \quad (X_2, \varepsilon _2) = (0,0), \quad (u_1, w_1) = (0,0). \end{equation} Let us calculate the minimal resolution of these singularities. For example, the singularity $(X_2, \varepsilon _2) = (0,0)$ is defined by the $\Z_2$ action $(X_2, \varepsilon _2) \mapsto (-X_2, -\varepsilon _2)$; i.e. this is a $A_1$ singularity, and it is resolved by the standard one time blow-up. The self-intersection number of the exceptional divisor is $-2$. Similarly, singularities $(Y_1, \varepsilon _1) = (0,0)$ and $(u_1, w_1) = (0,0)$ are resolved by one time blow-ups, whose self-intersection numbers of the exceptional divisors are $-3$ and $-6$, respectively. From the minimal resolution of singularities of $N$, we remove the space of initial conditions $E(0)$. Then, we obtain the union of four projective lines, whose configuration is described as Fig.\ref{fig3}(b), see also Fig.\ref{fig2}. \begin{figure} \begin{center} \includegraphics[]{fig1-2.eps} \caption[]{The minimal resolution of singularities of the closure of $E(0)$.} \label{fig2} \end{center} \end{figure} Although the diagram Fig.\ref{fig3}(b) is different from $\tilde{E}_8$, it is remarkable that the self-intersection numbers $-2, -3, -6$ are the same as the lengths of arms from the center of $\tilde{E}_8$. The same results hold for ($\text{P}_\text{II}$) and ($\text{P}_\text{IV}$). Let $\mathcal{M}_\text{II}$ be the total space for ($\text{P}_\text{II}$) obtained by two points blow-up with the weights $(4,2,3)$ of $\C P^3(2,1,2,3)$ constructed in Sec.4.2. Consider a fiber $E(0)$ (the space of initial conditions) on $z = 0$ and take the closure $N$ of it in $\mathcal{M}_\text{II}$. From the minimal resolution of $N$ at the orbifold singularities, we remove $E(0)$. Then, we obtain the union of four projective lines, whose configuration and self-intersection numbers are described in Fig.\ref{fig3}(b). Although the diagram Fig.\ref{fig3}(b) is different from $\tilde{E}_7$, the self-intersection numbers $-4, -4, -6$ are the same as the lengths of arms from the center of $\tilde{E}_7$. A similar result is true for ($\text{P}_\text{IV}$). \begin{figure} \begin{center} \includegraphics[]{fig1-3.eps} \caption[]{(a) The usual extended Dynkin diagram. Each vertex denotes $\C P^1$, and two $\C P^1$ are connected by an edge if they intersect with each other. All self-intersection numbers are $-2$. (b) The diagram obtained from our total space $\mathcal{M}_\text{J}$. Each number denotes the self-intersection number.} \label{fig3} \end{center} \end{figure}
\section{Introduction} Since the seminal work of Owen (\citeyear{Owe88}, \citeyear{Owe90}), there have been many advances in empirical likelihood method that have brought applications of the method to virtually every area of statistical research. It has been widely observed [e.g., \citet{HalLaS90,QinLaw94,CorDavSpa95,Owe01,LiuChe10}] that empirical likelihood ratio confidence regions can have poor accuracy, especially in small sample and multidimensional situations. In particular, there is a persistent undercoverage problem in that coverage probabilities of empirical likelihood ratio confidence regions tend to be lower than the nominal levels. In this paper, we tackle a fundamental problem underlying the poor accuracy and undercoverage, that is, the empirical likelihood is defined on only a part of the parameter space. We call this the \emph{mismatch problem} between the domain of the empirical likelihood and the parameter space. We solve this problem through a geometric approach that expands the domain to the full parameter space. Our solution brings about substantial improvements in accuracy of the empirical likelihood inference and is particularly useful for small sample and multidimensional situations. To see the mismatch problem, consider the example of empirical likelihood for the mean based on $n$ observations $X_1, \ldots, X_n$ of a random vector $X\in\mathbb{R}^d$. The underlying parameter space $\Theta$ is $\mathbb{R}^d$ itself. But for a given $\theta\in\mathbb {R}^d$, when the convex hull of the $(X_i-\theta)$ does not contain $0$, the empirical likelihood $L(\theta)$, the empirical likelihood ratio $R(\theta)=n^nL(\theta)$ and the empirical log-likelihood ratio $l(\theta)=-2\log R(\theta)$ are all undefined. When this occurs, an established convention assigns $L(\theta)=0$, and technically $L(\theta)$, $R(\theta)$ and $l(\theta)$ are now defined over the full parameter space. Nevertheless, to highlight the difference between the natural domain of the empirical likelihood and the parameter space, we define the common domain of $L(\theta)$, $R(\theta)$ and $l(\theta)$ as \begin{equation}\label{1-2} \Theta_n=\bigl\{\theta\dvtx \theta\in\Theta\mbox{ and } l(\theta)<+ \infty\bigr\}. \end{equation} With this definition, the mismatch can now be expressed as $\Theta _n\subset\Theta$. For the mean, $\Theta_n$ is the interior of the convex hull of the $X_i$, which is indeed a proper subset of $\Theta =\mathbb{R}^d$. The mismatch $\Theta_n\subset\Theta$ holds for empirical likelihoods in general as the basic formulation common to all empirical likelihoods has a convex hull constraint on the origin, such as the one for the mean above, which may be violated by some $\theta$ values in the parameter space $\Theta$. The convex hull constraint violation underlying the mismatch is well known in the empirical likelihood literature. It was first noted in \citet{Owe90} for the case of the mean. See also \citet{Owe01}. To assess its impact on coverage probabilities of empirical likelihood ratio confidence regions, \citet{Tsa04} investigated bounds on coverage probabilities resulting from the convex hull constraint. To bypass this constraint, \citet{Bar07} introduced a \emph{penalized empirical likelihood} (PEL) for the mean which removes the convex hull constraint in the formulation of the original empirical likelihood (OEL) of Owen (\citeyear{Owe90}, \citeyear{Owe01}) and replaces it with a penalizing term based on the Mahalanobis distance. For parameters defined by general estimating equations, \citet{CheVarAbr08} introduced an \emph{adjusted empirical likelihood} (AEL) which retains the formulation of the OEL but adds a pseudo-observation to the sample. The AEL is just the OEL defined on the augmented sample. But due to the clever construction of the pseudo-observation, the convex hull constraint will never be violated by the AEL. \citet{EmeOwe09} showed that the AEL statistic has a boundedness problem which may lead to trivial 100\% confidence regions. They proposed an extension of the AEL involving adding two pseudo-observations to the sample to address the boundedness problem. \citet{CheHua13} also addressed the boundedness problem by modifying the adjustment factor in the pseudo-observation. \citet{LiuChe10} proved a surprising result that under a certain level of adjustment, the AEL confidence region achieves the second order accuracy of the Bartlett corrected empirical likelihood (BEL) region by \citet{DiCHalRom91N2}. Recently, \citet{LahMuk12} showed that under certain dependence structures, a~modified PEL for the mean works in the extremely difficult case of large dimension and small sample size. The PEL and the AEL are both defined on $\mathbb{R}^d$, and are thus free from the mismatch problem. In this paper, we propose a new \emph{extended empirical likelihood} (EEL) that is also free from the mismatch problem. We derive this EEL through the domain expansion method of \citet{Tsa13} which expands the domain of the OEL but retains its important geometric characteristics. This EEL makes effective use of the dimension information in the data and can attain the second order accuracy of the BEL. The most important aspect of this EEL, however, is that there is an easy-to-use first order version which is substantially more accurate than the OEL. This first order EEL is also more accurate than available second order empirical likelihood methods in most small sample and multidimensional applications and is comparable in accuracy to the latter when the sample size is large. The focus of the present paper is on the construction of EEL for the mean through which we introduce the basic idea of and important tools for expanding the OLE domain to the full parameter space. Under certain conditions, EEL for other parameters may also be constructed but this will be discussed elsewhere. For brevity, we will use ``OEL $l(\theta)$'' and ``EEL $l^*(\theta)$'' to refer to the original and extended empirical log-likelihood ratios for the mean, respectively. Throughout this paper, we assume that the parameter space $\Theta$ is $\mathbb{R}^d$. The case where $\Theta$ is a known subset of $\mathbb{R}^d$ can be handled by finding EEL $l^*(\theta)$ defined on $\mathbb{R}^d$ first and then, for $\theta \notin\Theta$ only, redefine it as $l^*(\theta)=+\infty$. \section{Preliminaries} We review several key results and assumptions for developing the EEL defined on the full parameter space. \subsection{Empirical likelihood for the mean} Let $X\in\mathbb{R}^d$ be a random vector with mean $\theta_0$ and covariance matrix $\Sigma_0$. Two assumptions we will need are: \begin{longlist}[$(A_2)$] \item[$(A_1)$] $\Sigma_0$ is a finite covariance matrix with full rank $d$; and \item[$(A_2)$] $\lim\sup_{\|t\|\rightarrow\infty}|E[\exp\{it^TX\}]| <1$ and $E\|X\|^{15}<+\infty$. \end{longlist} These are also assumptions under which the OEL for the mean is Bartlett correctable [\citet{DiCHalRom91N1}, \citet{r4}]. Let $X_1, \ldots, X_n$ be independent copies of $X$ where $n>d$. Let $\Theta_n$ be the collection of points in the interior of the convex hull of the $X_i$. For a $\theta\in\mathbb{R}^d$, \citet{Owe90} defined the empirical likelihood ratio $R(\theta)$ as \begin{equation}\label{2-50} R(\theta) = \sup\Biggl\{ \prod_{i=1}^{n} nw_i \Big| \sum_{i=1}^nw_i(X_i- \theta)=0, w_i \geq0, \sum_{i=1}^nw_i=1 \Biggr\}. \end{equation} It may be verified that $0<R(\theta)\leq1$ iff $\theta\in\Theta_n$. Also, $R(\theta)=0$ if $\theta\notin\Theta_n$. Hence, the domain of the OEL $l(\theta)=-2\log R(\theta)$ is $\Theta_n$. For a $\theta\in\Theta_n$, the method of Lagrange multipliers may be used to show that \begin{equation}\label{2-58} l(\theta) = 2 \sum^n_{i=1} \log\bigl\{1+ \lambda^T(X_i-\theta)\bigr\}, \end{equation} where the multiplier $\lambda=\lambda(\theta)\in\mathbb{R}^d$ satisfies \begin{equation}\label{2-56} \sum_{i=1}^n \frac{X_i-\theta}{1+\lambda^T(X_i-\theta)} = 0. \end{equation} Under assumption ($A_1$), \citet{Owe90} showed that OEL $l(\theta)$ satisfies \begin{equation}\label{2-60} l(\theta_0) \stackrel{\mathcal{D}} {\longrightarrow} \chi^2_d\qquad\mbox{as $n\rightarrow+\infty$}. \end{equation} For an $\alpha\in(0,1)$, let $c$ be the ($1-\alpha$)th quantile of the $\chi^2_d$ distribution. Then, the $100(1-\alpha)\%$ OEL confidence region for $\theta_0$ is given by \begin{equation}\label{1-3} {\mathcal C}_{1-\alpha}=\bigl\{\theta\dvtx \theta\in\Theta_n \mbox{ and } l(\theta) \leq c\bigr\}. \end{equation} Under assumptions ($A_1$) and ($A_2$), DiCiccio, Hall and Romano (\citeyear{DiCHalRom91N1,DiCHalRom91N2}) showed that the coverage error of ${\mathcal C}_{1-\alpha}$ is $O(n^{-1})$, that is, \begin{equation}\label{2-64} P(\theta_0 \in{\mathcal C}_{1-\alpha})=P\bigl(l( \theta_0)\leq c\bigr)=P\bigl(\chi^2_d \leq c \bigr)+O\bigl(n^{-1}\bigr). \end{equation} More importantly, they showed that the empirical likelihood is Bartlett correctable. To give a brief account of this surprising result, let \begin{equation}\label{2-68} {\mathcal C}_{1-\alpha}'=\bigl\{\theta\dvtx l(\theta) \bigl(1-bn^{-1}\bigr) \leq c\bigr\} \end{equation} be the Bartlett corrected empirical likelihood ratio confidence region where $b$ is the \emph{Bartlett correction constant} and $(1-bn^{-1})$ is the \emph{Bartlett correction factor}, DiCiccio, Hall and Romano (\citeyear{DiCHalRom91N1,DiCHalRom91N2}) showed that ${\mathcal C}_{1-\alpha}'$ has a coverage error of only $O(n^{-2})$, that is, \begin{equation}\label{2-70} P\bigl(\theta_0 \in{\mathcal C}_{1-\alpha}' \bigr)=P\bigl[l(\theta_0) \bigl(1-bn^{-1}\bigr)\leq c \bigr]=P\bigl(\chi^2_d \leq c\bigr)+O\bigl(n^{-2} \bigr). \end{equation} In practice, the Bartlett correction constant $b$ cannot be determined since it depends on the moments of $X$ which are not available in the nonparametric setting of the empirical likelihood. However, replacing the Bartlett correction factor in (\ref{2-70}) with $[1-bn^{-1}+O_p(n^{-3/2})]$ does not affect the $O(n^{-2})$ term in its right-hand side, that is, \begin{equation}\label{2-72} P\bigl\{l(\theta_0)\bigl[1-bn^{-1}+O_p \bigl(n^{-3/2}\bigr)\bigr]\leq c\bigr\}=P\bigl(\chi^2_d \leq c\bigr)+O\bigl(n^{-2}\bigr). \end{equation} This allows us to replace $b$ in (\ref{2-68}) and (\ref{2-70}) with a $\sqrt{n}$-consistent estimate $\hat{b}$ without invalidating (\ref {2-70}). See \citet{DiCHalRom91N2} and \citet{HalLaS90} for detailed discussions on Bartlett correction. \subsection{Extended empirical likelihood} The OEL confidence region ${\mathcal C}_{1-\alpha}$ in (\ref{1-3}) is confined to the OEL domain $\Theta_n$. This is a main cause of the undercoverage problem associated with ${\mathcal C}_{1-\alpha}$ [\citet{Tsa04}]. To alleviate the problem, \citet{Tsa13} proposed to expand $\Theta _n$ which will lead to larger EL confidence regions. Let $h_n\dvtx \mathbb {R}^d \rightarrow\mathbb{R}^d$ be a bijective mapping and define a new empirical log-likelihood ratio $l^*(\theta)$ through the OEL $l(\theta)$ as follows: \begin{equation}\label{1-5} l^*(\theta)=l\bigl(h^{-1}_n(\theta)\bigr)\qquad \mbox{for $\theta\in\mathbb{R}^d$}. \end{equation} Then the domain for the new empirical log-likelihood ratio is $\Theta _n^*= h_n(\Theta_n)$. Here, $h_n$ plays the role of reassigning or extending the OEL values of points in $\Theta_n$ to points in $\Theta_n^*$. Because of this, \citet{Tsa13} named $l^*(\theta)$ the extended empirical log-likelihood ratio or simply EEL. In particular, \citet{Tsa13} used the following $\tilde{\theta}$-centred similarity mapping $h_n^*\dvtx \mathbb {R}^d\rightarrow\mathbb{R}^d$ \begin{equation}\label{2-80} h_n^*(\theta)=\tilde{\theta}+\gamma_n(\theta-\tilde{ \theta}), \end{equation} where $\tilde{\theta}$ is the sample mean and $\gamma_n\in\mathbb {R}^1$ is a constant (which we will refer to as the \emph{expansion factor}) satisfying $\gamma_n \geq0$ and $\gamma_n \rightarrow1$ as $n \rightarrow+\infty$. If we choose $\gamma_n>1$, then $\Theta_n \subset\Theta^*_n\subset\mathbb{R}^d$, and $\Theta^*_n$ alleviates the mismatch problem of $\Theta_n$. The EEL confidence region for $\theta _0$ is given by \begin{equation}\label{1-7} {\mathcal C}_{1-\alpha}^*=\bigl\{\theta\dvtx \theta\in\Theta^*_n \mbox{ and } l^*(\theta) \leq c\bigr\}. \end{equation} The advantages of the EEL based on $h_n^*$ in (\ref{2-80}) are: (1) the EEL confidence regions are similarly transformed OEL confidence regions, as such they retain the natural centre and shape of the OEL confidence regions, (2) the EEL can be applied to empirical likelihood inference for a wide range of parameters, and (3) with a properly selected constant $\gamma_n$, EEL confidence regions can achieve the second order accuracy of $O(n^{-2})$. Nevertheless, the EEL based on $h_n^*$ is only a partial solution to the mismatch problem because the domain of this EEL $\Theta^*_n$ is also a proper subset of $\mathbb{R}^d$. A~second order version of this EEL has been found to have good accuracy in one- and two-dimensional problems. But it also tends to undercover and no accurate first order version of this EEL is available. These motivated us to consider an EEL defined on the full parameter space. \section{Extended empirical likelihood on the full parameter space} Consider a bijective mapping from the OEL domain to the parameter space, $h_n\dvtx \Theta_n \rightarrow\Theta=\mathbb{R}^d$. Under such a mapping, the EEL $l^*(\theta)$ given by (\ref{1-5}) is well defined throughout $\mathbb{R}^d$ and is thus free from the mismatch problem. In this section, we first construct such a mapping using $h_n^*$ in (\ref{2-80}). We call it the \emph{composite similarity mapping} and denote it by $h^C_n\dvtx \Theta_n\rightarrow\mathbb{R}^d$. We then study the asymptotic properties of the EEL $l^*(\theta)$ based on $h^C_n$. \subsection{The composite similarity mapping} The simple similarity mapping $h_n^*$ in (\ref{2-80}) maps OEL domain $\Theta_n$ onto a similar but bounded region in $\mathbb{R}^d$. If we think of $\Theta_n$ as a region consisting of distinct and nested contours of the OEL, then $h_n^*$ expands all contours with the same constant expansion factor $\gamma_n$. In order to map $\Theta_n$ onto the full $\mathbb{R}^d$, we need to expand contours on the outside more and more so that the images of the contours will fill up the entire $\mathbb{R}^d$. To achieve this, consider level-$\tau$ contour of the OEL $l(\theta)$, \begin{equation}\label{3-2} c(\tau)=\bigl\{\theta\dvtx \theta\in\Theta_n\mbox{ and } l(\theta)=\tau\bigr\}, \end{equation} where $\tau\geq0$. The contours form a partition of the OEL domain, \begin{equation}\label{3-3} \Theta_n = \bigcup_{\tau\in[0,+\infty)} c(\tau). \end{equation} In light of (\ref{3-3}), the centre of $\Theta_n$ is $c(0)=\{\tilde {\theta}\}$ and the outwardness of a $c(\tau)$ with respect to the centre is indexed by $\tau$; the larger the $\tau$ value, the more outward $c(\tau)$ is. If we allow the expansion factor $\gamma_n$ to be a continuous monotone increasing function of $\tau$ and allow $\gamma _n$ to go to infinity when $\tau$ goes to infinity, then (such a variation of) $h_n^*$ will map $\Theta_n$ onto $\mathbb{R}^d$. Hence, we define the composite similarity mapping $h^C_n\dvtx \Theta _n\rightarrow\mathbb{R}^d$ as follows: \begin{equation}\label{3-15} h_n^C(\theta)=\tilde{\theta}+\gamma\bigl(n,l(\theta) \bigr) (\theta-\tilde{\theta})\qquad\mbox{for $\theta\in \Theta_n$}, \end{equation} where $\gamma(n,l(\theta))$ is given by \begin{equation}\label{3-16} \gamma\bigl(n,l(\theta)\bigr)=1+\frac{l(\theta)}{2n}. \end{equation} Function $\gamma(n,l(\theta))$ is the new expansion factor which depends continuously on $\theta$ through the value of $l(\theta)$ or $\tau=l(\theta)$. For convenience, we will emphasis the dependence of $\gamma(n,l(\theta))$ on $l(\theta)$ instead of $\theta$ or $\tau$. This new expansion factor has the two desired properties discussed above: \begin{eqnarray} \label{3-e1} & & \mbox{for a fixed $n$, if $l(\theta_1)<l(\theta _2)$, then $\gamma\bigl(n,l(\theta_2)\bigr)<\gamma \bigl(n,l(\theta_2)\bigr)$; and} \\ \label{3-e2} & & \mbox{for a fixed $n$, $\gamma\bigl(n,l(\theta)\bigr)\rightarrow +\infty$ as $l(\theta) \rightarrow+\infty$.} \end{eqnarray} The inclusion of the denominator $2n$ in (\ref{3-16}) ensures that the expansion factor converges to 1, reflecting the fact that there is no need for domain expansion for large sample sizes. Also, the constant $2$ in the denominator provides extra adjustment to the speed of expansion and may be replaced with other positive constants (see Figure \ref{fig1}). We choose to use 2 here as the corresponding $\gamma(n,l(\theta))$ in (\ref{3-16}) is asymptotically equivalent to a likelihood based expansion factor $\gamma(n, L(\theta))=L(\theta )^{-1/n}$ which we had first considered and was found to give accurate numerical results. The definition of $\gamma(n, l(\theta))$ in (\ref {3-16}) uses $l(\theta)$ instead of $L(\theta)$ because of convenience for theoretical investigations. A more general form of $\gamma(n,l(\theta))$ will be considered later. \begin{figure} \includegraphics{1143f01.eps} \caption{\textup{(a)} Contours of the OEL $l(\theta)$ (for which the expansion factor is 1.0); \textup{(b)} contours of the EEL $l^*(\theta)$ with expansion factor $\gamma(11,l(\theta))=1+l(\theta)/3n$; \textup{(c)} contours of the EEL $l^*(\theta)$ with $\gamma(11,l(\theta))=1+l(\theta)/2n$; \textup{(d)} contours of the EEL $l^*(\theta)$ with $\gamma(11,l(\theta))=1+l(\theta)/n$. All four plots are based on the same random sample of 11 points shown in small circles. The star in the middle is the sample mean. The expansion factor increases as we go from plot \textup{(b)} to plot \textup{(d)}, and correspondingly the EEL contours also become bigger in scale from plot \textup{(b)} to \textup{(d)}. But the centre and shapes of contours are the same in all plots.}\label{fig1} \end{figure} Theorem \ref{thm1} below summarizes the key properties of the composite similarity mapping $h_n^C$. Its proof and that of subsequent theorems and lemmas may all be found in the \hyperref[app]{Appendix}. \begin{theorem} \label{thm1} Under assumption ($A_1$), the composite similarity mapping $h^C_n\dvtx \Theta_n\rightarrow\mathbb{R}^d$ defined by (\ref{3-15}) and (\ref {3-16}) satisfies: \begin{longlist} \item it has a unique fixed point at the mean $\tilde{\theta}$; \item it is a similarity mapping for each individual $c(\tau)$; and \item it is a bijective mapping from $\Theta_n$ to $\mathbb{R}^d$. \end{longlist} \end{theorem} Because of (ii) above, we call $h_n^C$ the composite similarity mapping as it may be viewed as a continuous sequence of simple similarity mappings from $\mathbb{R}^d$ to $\mathbb{R}^d$ indexed by $\tau=l(\theta)\in[0, +\infty)$. The ``$\tau$th'' mapping from this sequence has expansion factor $\gamma(n,l(\theta))=\gamma(n,\tau)$. It is just the simple similarity mapping $h_n^*$ in (\ref{2-80}) with $\gamma_n=\gamma(n,\tau)$, and is used exclusively to map the ``$\tau $th'' OEL contour $c(\tau)$. The latter has been implicitly built into $h_n^C$ since for all $\theta\in c(\tau)$, $l(\theta)=\tau$ which implies the corresponding expansion factor of $h_n^C$ is the constant $\gamma(n,\tau)$ that defines the ``$\tau$th'' mapping. It should be noted that $h_n^C$ is not a similarity mapping from $\mathbb{R}^d$ to $\mathbb{R}^d$ itself due to the dependence of the expansion factor $\gamma(n,l(\theta))$ on $\theta$ and its domain $\Theta_n$ which is only a bounded subset of $\mathbb{R}^d$. \subsection{The extended empirical likelihood under the composite similarity mapping} By Theorem \ref{thm1}, $h_n^C\dvtx \Theta_n\rightarrow\mathbb{R}^d$ is bijective. Hence, it has an inverse which we denote by $h_n^{-C}\dvtx \mathbb{R}^d \rightarrow\Theta_n$. The EEL $l^*(\theta)$ under $h_n^C$ is \[ l^*(\theta)=l\bigl(h_n^{-C}(\theta)\bigr) \qquad \mbox{for $\theta\in\mathbb{R}^d$}, \] which is defined throughout $\mathbb{R}^d$. The contours of $l^*(\theta)$ are larger in scale but have the same centre and identical shape as that of OEL $l(\theta)$. Figure \ref{fig1} compares their contours with a sample of 11 two-dimensional observations. It shows that geometrically, mapping $h_n^C$ is anchored at the sample mean $\tilde{\theta}$ as it is the fixed point of $h_n^C$ that is not moved. From this anchoring point, the mapping pushes out/expands each OEL contour $c(\tau)$ proportionally in all directions at an expansion factor of $\gamma(n,\tau)$ to form an EEL contour. The boundary points of $\Theta_n$ are all pushed out to the infinity. In the following, we will use $\theta'$ to denote the image of a $\theta\in\mathbb{R}^d$ under the inverse transformation $h_n^{-C}$, that is, $h_n^{-C}(\theta)=\theta' \in\Theta_n$. Of particular interest is the image of the unknown true mean $\theta_0$, \begin{equation}\label{image} h_n^{-C}(\theta_0)=\theta_0'. \end{equation} Because the inverse transformation $h_n^{-C}$ does not have an analytic expression, that for $\theta'_0$ is also not available. Nevertheless, Lemma \ref{lem2} gives an asymptotic assessment on its distance to $\theta_0$. The proof of Lemma \ref{lem2} will need Lemma \ref{lem1} below which shows that inside $\Theta_n$, the OEL $l(\theta)$ is a ``monotone increasing'' function along each ray originating from the mean $\tilde{\theta}$. \begin{lemma} \label{lem1} Under assumption ($A_1$), for a fixed point $\theta\in\Theta_n$ and any value $\alpha\in[0,1]$ the OEL $l(\theta)$ satisfies \[ l\bigl(\tilde{\theta}+\alpha(\theta-\tilde{\theta})\bigr)\leq l(\theta). \] \end{lemma} Denote by $[\tilde{\theta}, \theta_0]$ the line segment connecting $\tilde{\theta}$ and $\theta_0$. Lemma \ref{lem2} below shows that $\theta_0'$ is on $[\tilde{\theta}, \theta_0]$ and is asymptotically very close to $\theta_0$. \begin{lemma} \label{lem2} Under assumption ($A_1$), point $\theta_0'$ defined by equation (\ref{image}) satisfies \textup{(i)} $\theta_0' \in[\tilde {\theta}, \theta_0]$ and \textup{(ii)} $\theta'_0-\theta_0 = O_p(n^{-3/2})$. \end{lemma} Using $\theta_0'$, the EEL $l^*(\theta_0)$ can now be expressed as \begin{equation}\label{3-50} l^*(\theta_0)=l\bigl(h_n^{-C}( \theta_0)\bigr)= l\bigl(\theta'_0\bigr)=l \bigl(\theta_0+\bigl(\theta_0'- \theta_0\bigr)\bigr). \end{equation} The following theorem gives the asymptotic distribution of $l^*(\theta_0)$: \begin{theorem} \label{thm2} Under assumption ($A_1$) and with the composite similarity mapping $h_n^C$ defined by (\ref{3-15}) and (\ref{3-16}), the EEL $l^*(\theta)$ satisfies \begin{equation}\label{3-58} l^*(\theta_0) \stackrel{\mathcal{D}} {\longrightarrow} \chi^2_d \qquad\mbox{as $n\rightarrow+\infty$}. \end{equation} \end{theorem} The proof of Theorem \ref{thm2} is based on the observation that $\| \theta_0'-\theta_0\|$ is asymptotically very small. This and (\ref {3-50}) imply that $l^*(\theta_0)=l(\theta_0)+o_p(1)$. This proof demonstrates an advantage of the EEL: it has the simple relationship with the OLE shown in (\ref{3-50}) through which we make use of known asymptotic properties of the OEL to study the EEL. Our derivation of a second order EEL below further explores this advantage. \subsection{Second order extended empirical likelihood} It may be verified that Theorems \ref{thm1} and \ref{thm2} also hold under any composite similarity mapping defined by (\ref{3-15}) and the following general form of the expansion factor, \begin{equation}\label{3-90} \gamma\bigl(n,l(\theta)\bigr)=1+\frac{\kappa[l(\theta)]^{\delta (n)}}{n^m}, \end{equation} where $\kappa$ and $m$ are both positive constants and $\delta(n)$ is a bounded function of $n$ satisfying $\varepsilon_n< \delta(n) \leq a$ for some constants $a>\varepsilon_n>0$. The availability of a whole family of $\gamma(n,l(\theta))$ functions for the construction of the EEL provides an opportunity to optimize our choice of this function to achieve the second order accuracy. Theorem \ref{thm3} below gives the optimal choice. \begin{theorem} \label{thm3} Assume $(A_1)$ and $(A_2)$ hold and denote by $l^*_s(\theta)$ the EEL under the composite similarity mapping (\ref {3-15}) with expansion factor \begin{equation}\label{3-80} \gamma_s\bigl(n,l(\theta)\bigr)=1+\frac{b }{2n}\bigl[l(\theta) \bigr]^{\delta(n)}, \end{equation} where $\delta(n)=O(n^{-1/2})$ and $b$ is the Bartlett correction constant in (\ref{2-68}). Then \begin{equation}\label{3-80-1} l^*_s(\theta_0)=l(\theta_0) \bigl[1-bn^{-1}+O_p\bigl(n^{-3/2}\bigr)\bigr] \end{equation} and \begin{equation}\label{3-81} P\bigl(l_s^*(\theta_0)\leq c\bigr)=P\bigl( \chi^2_d \leq c\bigr)+O\bigl(n^{-2}\bigr). \end{equation} \end{theorem} In our subsequent discussions, we will refer to an $l^*_s(\theta)$ defined by the expansion factor $\gamma_s(n,l(\theta))$ in (\ref{3-80}) as a \emph{second order EEL} on the full parameter space. The EEL $l^*(\theta)$ defined by $\gamma(n, l(\theta))$ in (\ref {3-16}) will henceforth be referred to as the \emph{first order EEL} on the full parameter space. The utility of the $\delta(n)$ in $\gamma_s(n,l(\theta))$ is to provide an extra adjustment for the speed of the domain expansion which ensures that $l^*_s(\theta)$ will behave asymptotically like the BEL and hence will have the second order accuracy of the BEL. For convenience, we set $\delta(n)=n^{-1/2}$. The resulting second order EEL turns out to be competitive in accuracy to the BEL and the second order AEL. For small sample and/or high dimension situations, confidence regions based on this second order EEL can have undercoverage problems like those based on the OEL and BEL. Fine-tuning of $\delta(n)$ for such situations is needed and methods of fine-tuning are discussed in \citet{Wu13}. Finally, we noted after Theorem \ref{thm2} that the first order EEL $l^*(\theta_0)$ can be expressed in terms of the OLE $l(\theta_0)$ as $l^*(\theta_0)=l(\theta_0)+o_p(1)$. The $o_p(1)$ term can be improved and in fact we have $l^*(\theta_0)=l(\theta_0)+O_p(n^{-1})$. See the proof of Theorem \ref{thm2} in the \hyperref[app]{Appendix}. An even stronger connection between $l^*(\theta_0)$ and $l(\theta_0)$ is given by Corollary \ref{cor1} below. \begin{cor} \label{cor1} Under assumptions ($A_1$) and ($A_2$), the first order EEL $l^*(\theta)$ satisfies \begin{equation}\label{3-81-1} l^*(\theta_0)=l(\theta_0)\bigl[1-l( \theta_0)n^{-1}+O_p\bigl(n^{-3/2} \bigr)\bigr]. \end{equation} \end{cor} The proof of Corollary \ref{cor1} follows from that for Theorem \ref {thm3}. This result provides a partial explanation for the remarkable numerical accuracy of confidence regions based on the first order EEL $l^*(\theta)$. \begin{table} \tabcolsep=0pt \caption{Simulated coverage probabilities for one-dimensional examples}\label{tab1} {\fontsize{8.6pt}{11pt}\selectfont{ \begin{tabular*}{\tablewidth}{@{\extracolsep{\fill}}l c c c c c d{1.4} c c c c@{}} \hline & $\bolds{n}$ & \textbf{Level} & \textbf{OEL} & \textbf{\textit{EEL}$\bolds{_1}$} & \multicolumn{1}{c}{\textbf{BEL}} & \multicolumn{1}{c}{\textbf{AEL}} & \multicolumn{1}{c}{\textbf{EEL}$\bolds{_2}$} & \multicolumn{1}{c}{\textbf{BEL}$\bolds{^*}$} & \multicolumn{1}{c}{\textbf{AEL}$\bolds{^*}$} & \multicolumn{1}{c@{}}{\textbf{EEL}$\bolds{^*_2}$} \\ \hline $N(0,1)$ &10 &0.90 &0.8506 &\textbf{0.8914} &0.8753 &0.8788 &0.8813 &0.8767 &0.8867 &0.8824 \\ & &0.95 &0.9039 &\textbf{0.9452} &0.9246 &0.9294 &0.9317 &0.9242 &0.9352 &0.9324 \\ & &0.99 &0.9580 &\textbf{0.9867} &0.9677 &0.9753 &0.9738 &0.9656 &0.9771 &0.9734 \\ &30 &0.90 &0.8920 &\textbf{0.9071} &0.9007 &0.9008 &0.9022 &0.9017 &0.9019 &0.9030 \\ & &0.95 &0.9398 &\textbf{0.9548} &0.9461 &0.9461 &0.9476 &0.9466 &0.9468 &0.9474 \\ & &0.99 &0.9866 &\textbf{0.9925} &0.9882 &0.9883 &0.9885 &0.9883 &0.9884 &0.9884 \\ &50 &0.90 &0.8941 &\textbf{0.9024} &0.8995 &0.8996 &0.9003 &0.8992 &0.8993 &0.9000 \\ & &0.95 &0.9447 &\textbf{0.9541} &0.9481 &0.9479 &0.9486 &0.9483 &0.9484 &0.9490 \\ & &0.99 &0.9880 &\textbf{0.9920} &0.9892 &0.9892 &0.9892 &0.9894 &0.9894 &0.9895 \\ [4pt] $t_5$ &10 &0.90 &0.8277 &\textbf{0.8765} &0.9226 &0.9979 &0.9209 &0.8520 &0.8873 &0.8782 \\ & &0.95 &0.8882 &\textbf{0.9394} &0.9599 &1.000 &0.9651 &0.9036 &0.9367 &0.9307 \\ & &0.99 &0.9556 &\textbf{0.9851} &0.9820 &1.000 &0.9887 &0.9499 &0.9798 &0.9751 \\ &30 &0.90 &0.8690 &\textbf{0.8852} &0.8999 &0.9028 &0.9017 &0.8852 &0.8885 &0.8882 \\ & &0.95 &0.9265 &\textbf{0.9436} &0.9476 &0.9509 &0.9502 &0.9385 &0.9428 &0.9420 \\ & &0.99 &0.9797 &\textbf{0.9888} &0.9875 &0.9901 &0.9886 &0.9831 &0.9863 &0.9861 \\ &50 &0.90 &0.8862 &\textbf{0.8967} &0.9040 &0.9048 &0.9052 &0.8977 &0.8983 &0.8987 \\ & &0.95 &0.9410 &\textbf{0.9491} &0.9515 &0.9518 &0.9518 &0.9465 &0.9471 &0.9474 \\ & &0.99 &0.9861 &\textbf{0.9918} &0.9907 &0.9913 &0.9913 &0.9881 &0.9882 &0.9886 \\[4pt] $\chi_1^2$ &10 &0.90 &0.7764 &\textbf{0.8174} &0.8726 &1.000 &0.8634 &0.6792 &0.8456 &0.8291 \\ & &0.95 &0.8314 &\textbf{0.8781} &0.9068 &1.000 &0.9030 &0.7239 &0.8918 &0.8779 \\ & &0.99 &0.8973 &\textbf{0.9378} &0.9417 &1.000 &0.9461 &0.7677 &0.9391 &0.9253 \\ &30 &0.90 &0.8594 &\textbf{0.8759} &0.8887 &0.8901 &0.8890 &0.8658 &0.8847 &0.8829 \\ & &0.95 &0.9115 &\textbf{0.9249} &0.9319 &0.9343 &0.9330 &0.9105 &0.9278 &0.9272 \\ & &0.99 &0.9659 &\textbf{0.9764} &0.9759 &0.9786 &0.9769 &0.9565 &0.9735 &0.9733 \\ &50 &0.90 &0.8722 &\textbf{0.8833} &0.8936 &0.8941 &0.8943 &0.8887 &0.8912 &0.8909 \\ & &0.95 &0.9318 &\textbf{0.9411} &0.9441 &0.9458 &0.9459 &0.9388 &0.9419 &0.9415 \\ & &0.99 &0.9779 &\textbf{0.9847} &0.9837 &0.9845 &0.9845 &0.9804 &0.9831 &0.9830 \\[4pt] $0.3N(0,1)$ &10 &0.90 &0.8470 &\textbf{0.8908} &0.8551 &0.8556 &0.8569 &0.8761 &0.8826 &0.8821 \\ $\quad{}+0.7N(2,1)$ & &0.95 &0.9036 &\textbf{0.9433} &0.9094 &0.9097 &0.9127 &0.9215 &0.9299 &0.9285 \\ & &0.99 &0.9564 &\textbf{0.9867} &0.9592 &0.9601 &0.9631 &0.9657 &0.9760 &0.9741 \\ &30 &0.90 &0.8930 &\textbf{0.9054} &0.8956 &0.8956 &0.8960 &0.9016 &0.9013 &0.9017 \\ & &0.95 &0.9438 &\textbf{0.9582} &0.9455 &0.9455 &0.9460 &0.9501 &0.9501 &0.9507 \\ & &0.99 &0.9873 &\textbf{0.9943} &0.9883 &0.9883 &0.9884 &0.9901 &0.9901 &0.9909 \\ &50 &0.90 &0.8965 &\textbf{0.9048} &0.8989 &0.8988 &0.8990 &0.9014 &0.9014 &0.9016 \\ & &0.95 &0.9465 &\textbf{0.9556} &0.9475 &0.9476 &0.9477 &0.9494 &0.9496 &0.9499 \\ & &0.99 &0.9876 &\textbf{0.9911} &0.9879 &0.9877 &0.9879 &0.9883 &0.9883 &0.9886 \\ \hline \end{tabular*}}} \end{table} \section{Numerical examples and comparisons}\label{sec4} We now present a simulation study comparing the EEL with the OEL, BEL and AEL. Throughout this section, we use $l^*_1(\theta)$ or EEL$_1$ to denote the first order EEL with expansion factor (\ref{3-16}), and use $l^*_2(\theta)$ or EEL$_2$ to denote the second order EEL given by expansion factor (\ref{3-80}) where $\delta(n)=n^{-1/2}$. \subsection{Low-dimensional examples} Tables \ref{tab1} and \ref{tab2} contain simulated coverage probabilities of confidence regions for the mean based on first order methods OEL, EEL$_1$ and second order methods BEL, AEL, EEL$_2$, BEL$^*$, AEL$^*$ and EEL$^*_2$. Here, BEL, AEL and EEL$_2$ are based on the theoretical Bartlett correction constant $b$, and BEL$^*$, AEL$^*$ and EEL$^*_2$ are based on $\tilde{b}_n$ which is a bias corrected estimate for $b$ given by \citet{LiuChe10}. \begin{table} \tabcolsep=0pt \caption{Simulated coverage probabilities for two-dimensional examples}\label{tab2} {\fontsize{8.6pt}{11pt}\selectfont{ \begin{tabular*}{\tablewidth}{@{\extracolsep{\fill}}l c c c c c d{1.4} c c c c@{}} \hline & $\bolds{n}$ & \textbf{Level} & \textbf{OEL} & \textbf{\textit{EEL}$\bolds{_1}$} & \multicolumn{1}{c}{\textbf{BEL}} & \multicolumn{1}{c}{\textbf{AEL}} & \multicolumn{1}{c}{\textbf{EEL}$\bolds{_2}$} & \multicolumn{1}{c}{\textbf{BEL}$\bolds{^*}$} & \multicolumn{1}{c}{\textbf{AEL}$\bolds{^*}$} & \multicolumn{1}{c@{}}{\textbf{EEL}$\bolds{^*_2}$} \\ \hline $\mathit{BV}1$ &10 &0.90 &0.7134 &\textbf{0.8118} &0.7965 &0.9777 &0.8212 &0.7561 &0.8350 &0.7989 \\ & &0.95 &0.7717 &\textbf{0.8758} &0.8407 &1.000 &0.8718 &0.8000 &0.8941 &0.8478 \\ & &0.99 &0.8484 &\textbf{0.9422} &0.8945 &1.000 &0.9268 &0.8570 &0.9680 &0.9083 \\ &30 &0.90 &0.8549 &\textbf{0.8888} &0.8824 &0.8856 &0.8872 &0.8786 &0.8813 &0.8822 \\ & &0.95 &0.9120 &\textbf{0.9426} &0.9313 &0.9348 &0.9361 &0.9296 &0.9336 &0.9337 \\ & &0.99 &0.9689 &\textbf{0.9868} &0.9772 &0.9798 &0.9796 &0.9757 &0.9783 &0.9779 \\ &50 &0.90 &0.8699 &\textbf{0.8917} &0.8869 &0.8874 &0.8894 &0.8859 &0.8869 &0.8886 \\ & &0.95 &0.9259 &\textbf{0.9428} &0.9354 &0.9361 &0.9374 &0.9344 &0.9351 &0.9363 \\ & &0.99 &0.9806 &\textbf{0.9908} &0.9846 &0.9852 &0.9856 &0.9839 &0.9848 &0.9851 \\[4pt] $\mathit{BV}2$ &10 &0.90 &0.7513 &\textbf{0.8521} &0.8035 &0.8573 &0.8282 &0.7942 &0.8451 &0.8229 \\ & &0.95 &0.8061 &\textbf{0.9095} &0.8627 &0.9397 &0.8861 &0.8499 &0.9103 &0.8833 \\ & &0.99 &0.8879 &\textbf{0.9693} &0.9202 &1.000 &0.9430 &0.9116 &0.9721 &0.9405 \\ &30 &0.90 &0.8714 &\textbf{0.9019} &0.8864 &0.8872 &0.8897 &0.8864 &0.8881 &0.8906 \\ & &0.95 &0.9256 &\textbf{0.9549} &0.9406 &0.9413 &0.9428 &0.9403 &0.9409 &0.9432 \\ & &0.99 &0.9789 &\textbf{0.9907} &0.9823 &0.9826 &0.9838 &0.9820 &0.9829 &0.9836 \\ &50 &0.90 &0.8826 &\textbf{0.9037} &0.8935 &0.8937 &0.8954 &0.8938 &0.8939 &0.8952 \\ & &0.95 &0.9348 &\textbf{0.9528} &0.9423 &0.9426 &0.9438 &0.9423 &0.9425 &0.9435 \\ & &0.99 &0.9839 &\textbf{0.9914} &0.9862 &0.9864 &0.9871 &0.9861 &0.9864 &0.9864 \\[4pt] $\mathit{BV}3$ &10 &0.90 &0.7001 &\textbf{0.7979} &0.7979 &1.000 &0.8162 &0.7333 &0.8363 &0.7911 \\ & &0.95 &0.7608 &\textbf{0.8581} &0.8374 &1.000 &0.8624 &0.7765 &0.8922 &0.8375 \\ & &0.99 &0.8331 &\textbf{0.9263} &0.8817 &1.000 &0.9151 &0.8286 &0.9639 &0.8942 \\ &30 &0.90 &0.8429 &\textbf{0.8775} &0.8749 &0.8789 &0.8788 &0.8719 &0.8780 &0.8764 \\ & &0.95 &0.9015 &\textbf{0.9363} &0.9266 &0.9326 &0.9319 &0.9221 &0.9282 &0.9280 \\ & &0.99 &0.9648 &\textbf{0.9817} &0.9740 &0.9776 &0.9760 &0.9709 &0.9750 &0.9740 \\ &50 &0.90 &0.8619 &\textbf{0.8836} &0.8807 &0.8820 &0.8836 &0.8787 &0.8801 &0.8808 \\ & &0.95 &0.9212 &\textbf{0.9403} &0.9351 &0.9364 &0.9379 &0.9325 &0.9346 &0.9347 \\ & &0.99 &0.9758 &\textbf{0.9848} &0.9810 &0.9816 &0.9817 &0.9802 &0.9806 &0.9810 \\[4pt] $\mathit{BV}4$ &10 &0.90 &0.6408 &\textbf{0.7371} &0.7882 &1.000 &0.7940 &0.6240 &0.8382 &0.7596 \\ & &0.95 &0.7030 &\textbf{0.8027} &0.8212 &1.000 &0.8377 &0.6637 &0.8896 &0.8051 \\ & &0.99 &0.7788 &\textbf{0.8808} &0.8580 &1.000 &0.8914 &0.7129 &0.9576 &0.8602 \\ &30 &0.90 &0.8229 &\textbf{0.8595} &0.8709 &0.8820 &0.8760 &0.8598 &0.8738 &0.8681 \\ & &0.95 &0.8857 &\textbf{0.9191} &0.9215 &0.9329 &0.9255 &0.9079 &0.9212 &0.9170 \\ & &0.99 &0.9520 &\textbf{0.9734} &0.9689 &0.9819 &0.9717 &0.9591 &0.9696 &0.9667 \\ &50 &0.90 &0.8494 &\textbf{0.8707} &0.8758 &0.8783 &0.8783 &0.8716 &0.8755 &0.8740 \\ & &0.95 &0.9060 &\textbf{0.9251} &0.9256 &0.9287 &0.9282 &0.9221 &0.9259 &0.9251 \\ & &0.99 &0.9675 &\textbf{0.9807} &0.9781 &0.9797 &0.9793 &0.9753 &0.9778 &0.9765 \\ \hline \end{tabular*}}} \end{table} Table \ref{tab1} gives four one-dimensional (1-$d$) examples. Table \ref{tab2} contains four bivariate ($\mathit{BV}$ or 2-$d$) examples; the first three were taken from \citet{LiuChe10}, the fourth is a ``2-$d$ chi-square'', and here are the details: ($\mathit{BV}_1$): $X_1|D\sim N(0,D^2)$ and $X_2|D\sim$ Gamma$(D^{-1},1)$. ($\mathit{BV}_2$): $X_1|D\sim$ Poisson$(D)$ and $X_2|D\sim$ Poisson$(D^{-1})$. ($\mathit{BV}_3$): $X_1|D\sim$ Gamma$(D,1)$ and $X_2|D\sim$ Gamma$(D^{-1},1)$. ($\mathit{BV}_4$): $X_1$ and $X_2$ are independent copies of a $\chi^2_1$ random variable. The $D$ in $\mathit{BV}_1$, $\mathit{BV}_2$ and $\mathit{BV}_3$ is a uniform random variable on $[1,2]$ which is used to induce dependence between $X_1$ and $X_2$. We included $n=10, 30, 50$ representing, respectively, small, medium and large sample sizes. Each entry in the tables is based on 10,000 random samples of size $n$, shown in column 2, from the distribution in column 1. Here are our observations: \begin{longlist}[(2)] \item[(1)] \textit{BEL}, \textit{AEL and EEL$_2$}: For $n=30$ and $50$, these three theoretical second order methods are extremely close in terms of coverage accuracy. This is to be expected as their coverage errors are all $O(n^{-2})$ which is very small for medium or large sample sizes. For $n=10$, the AEL statistic suffers from a boundedness problem [\citet{EmeOwe09}] which may lead to trivial 100\% confidence regions or inflated coverage probabilities. This explains the 1.000's in various places in the AEL column and renders the AEL unsuitable for such small sample sizes. Between BEL and EEL$_2$, the latter is more accurate, especially for the 2-$d$ examples. Overall, EEL$_2$ is the most accurate theoretical second order method. \item[(2)] \textit{BEL$^*$}, \textit{AEL$^*$ and EEL$^*_2$}: For $n=30$ and $50$, the AEL$^*$ and EEL$^*_2$ are slightly more accurate than the BEL$^*$, especially in 2-$d$ examples. For $n=10$, the AEL$^*$ has higher coverage probabilities but these are inflated by and unreliable due to the boundedness problem. Also, EEL$^*_2$ is more accurate than BEL$^*$. For the ``2-$d$ chi-square'' in example $\mathit{BV}_4$, EEL$^*_2$ is at least 12\% more accurate than the BEL$^*$. Overall, EEL$^*_2$ is the most reliable and accurate among the three. \item[(3)] \textit{OEL and EEL$_1$}: These first order methods are simpler than the second order methods as they do not require computation of the theoretical or estimated Bartlett correction factor. The EEL$_1$ is consistently and substantially more accurate than the OEL. In particular, for 2-$d$ examples with $n=10$, the EEL$_1$ is more accurate by about 10\%. \begin{table} \tabcolsep=0pt \caption{Simulated coverage probabilities for the mean of $d$-dimensional multivariate normal distributions}\label{tab3} {\fontsize{8.6pt}{11pt}\selectfont{ \begin{tabular*}{\tablewidth}{@{\extracolsep{\fill}}l c c c c c d{1.4} c c d{1.4} c@{}} \hline & $\bolds{n}$ & \textbf{Level} & \textbf{OEL} & \textbf{\textit{EEL}$\bolds{_1}$} & \multicolumn{1}{c}{\textbf{BEL}} & \multicolumn{1}{c}{\textbf{AEL}} & \multicolumn{1}{c}{\textbf{EEL}$\bolds{_2}$} & \multicolumn{1}{c}{\textbf{BEL}$\bolds{^*}$} & \multicolumn{1}{c}{\textbf{AEL}$\bolds{^*}$} & \multicolumn{1}{c@{}}{\textbf{EEL}$\bolds{^*_2}$} \\ \hline $d=5$ &10 &0.90 &0.3007 &\textbf{0.5897} &0.3839 &1.000 &0.5306 &0.3691 &1.000 &0.5005 \\ & &0.95 &0.3368 &\textbf{0.6794} &0.4135 &1.000 &0.5842 &0.4028 &1.000 &0.5500 \\ & &0.99 &0.3946 &\textbf{0.7984} &0.4498 &1.000 &0.6642 &0.4422 &1.000 &0.6287 \\ &30 &0.90 &0.7790 &\textbf{0.8862} &0.8258 &0.8554 &0.8455 &0.8273 &0.8585 &0.8468 \\ & &0.95 &0.8497 &\textbf{0.9436} &0.8880 &0.9208 &0.9047 &0.8884 &0.9256 &0.9052 \\ & &0.99 &0.9337 &\textbf{0.9881} &0.9532 &0.9889 &0.9629 &0.9539 &0.9899 &0.9634 \\ &50 &0.90 &0.8476 &\textbf{0.9036} &0.8752 &0.8803 &0.8820 &0.8757 &0.8808 &0.8825 \\ & &0.95 &0.9089 &\textbf{0.9522} &0.9297 &0.9341 &0.9349 &0.9297 &0.9349 &0.9354 \\ & &0.99 &0.9728 &\textbf{0.9913} &0.9804 &0.9833 &0.9830 &0.9804 &0.9839 &0.9831 \\[4pt] $d=10$ &20 &0.90 &0.1889 &\textbf{0.5367} &0.2845 &1.000 &0.4297 &0.2823 &1.000 &0.4235 \\ & &0.95 &0.2281 &\textbf{0.6260} &0.3209 &1.000 &0.4905 &0.3191 &1.000 &0.4824 \\ & &0.99 &0.2895 &\textbf{0.7708} &0.3747 &1.000 &0.5783 &0.3727 &1.000 &0.5717 \\ &30 &0.90 &0.4689 &\textbf{0.7752} &0.5944 &1.000 &0.6750 &0.5954 &1.000 &0.6752 \\ & &0.95 &0.5432 &\textbf{0.8594} &0.6627 &1.000 &0.7492 &0.6635 &1.000 &0.7480 \\ & &0.99 &0.6698 &\textbf{0.9442} &0.7670 &1.000 &0.8514 &0.7675 &1.000 &0.8527 \\ &50 &0.90 &0.7097 &\textbf{0.8806} &0.7921 &0.9531 &0.8189 &0.7933 &0.9582 &0.8198 \\ & &0.95 &0.7959 &\textbf{0.9393} &0.8577 &0.9968 &0.8827 &0.8588 &0.9974 &0.8838 \\ & &0.99 &0.9027 &\textbf{0.9864} &0.9392 &1.000 &0.9546 &0.9396 &1.000 &0.9549 \\[4pt] $d=15$ &30 &0.90 &0.1224 &\textbf{0.4850} &0.2199 &1.000 &0.3581 &0.2196 &1.000 &0.3569 \\ & &0.95 &0.1513 &\textbf{0.5761} &0.2504 &1.000 &0.4130 &0.2502 &1.000 &0.4124 \\ & &0.99 &0.2155 &\textbf{0.7490} &0.3100 &1.000 &0.5077 &0.3099 &1.000 &0.5054 \\ &50 &0.90 &0.4769 &\textbf{0.7983} &0.6177 &1.000 &0.6883 &0.6191 &1.000 &0.6894 \\ & &0.95 &0.5665 &\textbf{0.8776} &0.6971 &1.000 &0.7630 &0.6985 &1.000 &0.7646 \\ & &0.99 &0.7065 &\textbf{0.9600} &0.8097 &1.000 &0.8682 &0.8103 &1.000 &0.8686 \\ &100 &0.90 &0.7696 &\textbf{0.9031} &0.8325 &0.9309 &0.8472 &0.8328 &0.9341 &0.8484 \\ & &0.95 &0.8484 &\textbf{0.9514} &0.8985 &0.9852 &0.9086 &0.8989 &0.9865 &0.9096 \\ & &0.99 &0.9405 &\textbf{0.9900} &0.9639 &1.000 &0.9693 &0.9641 &1.000 &0.9692 \\ \hline \end{tabular*}}} \end{table} \item[(4)] \textit{EEL$_1$ versus EEL$^*_2$}: These are the most accurate practical first and second order methods, respectively. Surprisingly, EEL$_1$ turns out to be slightly more accurate than EEL$^*_2$. Only the (impractical) theoretical second order EEL$_2$ is comparable to EEL$_1$ in accuracy. This intriguing observation may be partially explained by Corollary \ref{cor1} where it was shown that $l^*_1(\theta_0)=l(\theta_0)[1+l(\theta_0)n^{-1}+O(n^{-3/2})]$, which resembles the Bartlett corrected OEL in (\ref{2-72}) with the constant $b$ replaced by $l(\theta_0)$. However, this does not account for its good accuracy for small sample sizes, which is due to the fact that EEL$_1$ makes good use of the dimension information through the composite similarity mapping. We will further elaborate on this in Section~\ref{sec4.2}. \item[(5)] \textit{EEL$_1$}: Overall, it is the most accurate among the eight methods that we have compared. Importantly, it is not just accurate in relative terms. It is sufficiently accurate in absolute terms for practical applications in most 1-$d$ examples, including cases of $n=10$. It is also quite accurate for 2-$d$ examples when $n=30, 50$. \end{longlist} \subsection{High-dimensional examples}\label{sec4.2} Table \ref{tab3} contains simulated coverage probabilities for the mean of three high-dimensional multivariate normal distributions ($d=5,10,15$). Our main interest here is to probe the small sample behaviour of all methods in high-dimension situations. Because of this, we have included only combinations of $n$ and $d$ where $n/d$, which we will refer to as the \emph{effective sample size}, is very small ($2\leq n/d\leq10$). The following are our observations based on Table \ref{tab3}: \begin{longlist}[(2)] \item[(1)] For these high-dimension examples, EEL$_1$ is the most accurate, surpassing even the theoretical second order EEL$_2$. Whereas the OEL uses dimension $d$ only once through the degrees of freedom in the chi-square calibration, EEL$_1$ uses $d$ twice. The expansion factor for EEL$_1$ is $1+l(\theta)/2n$ which implicitly depends on $d$; the 100($1-\alpha$)\% EEL$_1$ confidence region is just the 100($1-\alpha$)\% OEL confidence region expanded by a factor of $1+\chi ^2_{d,1-\alpha}/2n$. Hence, EEL$_1$ uses $d$ through the chi-square calibration of the OEL region {and} the expansion factor. \begin{table}[b] \caption{Average lengths of EL confidence intervals for $N(0,1)$ mean} \label{tab4} \begin{tabular*}{\tablewidth}{@{\extracolsep{\fill}}l c c c c c c@{}} \hline $\bolds{n}$ &\textbf{Level}& \textbf{OEL} & \textbf{EEL}$\bolds{_1}$ & \textbf{BEL}$\bolds{^*}$ & \textbf{AEL}$\bolds{^*}$ & \textbf{EEL}$\bolds{^*_2}$ \\ \hline 10 &0.90 &0.965 &1.096 &1.044 &$N/A$ (0.026) &1.077 \\ &0.95 &1.149 &1.370 &1.242 &$N/A$ (0.058) &1.298 \\ &0.99 &1.499 &1.996 &1.615 &$N/A$ (0.172) &1.731 \\[4pt] 30 &0.90 &0.589 &0.616 &0.606 &0.606 &0.608 \\ &0.95 &0.706 &0.752 &0.726 &0.727 &0.731 \\ &0.99 &0.940 &1.044 &0.967 &0.969 &0.976 \\[4pt] 50 &0.90 &0.460 &0.473 &0.467 &0.468 &0.468 \\ &0.95 &0.551 &0.572 &0.560 &0.560 &0.561 \\ &0.99 &0.732 &0.780 &0.744 &0.744 &0.747 \\ \hline \end{tabular*} \end{table} For a fixed $\alpha\in(0,1)$, the chi-square quantile $\chi ^2_{d,1-\alpha}$ and consequently the EEL$_1$ expansion factor $1+\chi ^2_{d,1-\alpha}/2n$ are increasing functions of $d$. Hence at a fixed $n$, EEL$_1$ automatically provides higher degrees of expansion for higher dimensions where this is indeed needed. \item[(2)] For multivariate normal means, Table \ref{tab3} shows that EEL$_1$ is accurate when the effective sample size satisfies $n/d\geq5$. However, when the underlying distribution is heavily skewed, the effective sample size needed to achieve similar accuracy needs to be 15 or larger. See Table \ref{tab2} for some 2-$d$ examples to this effect. \item[(3)] The AEL and AEL$^*$ broke down in most cases with 100\% coverage probabilities. This further illustrates the observation that AEL methods may not be suitable when the effective sample size is small. Among OEL, BEL, EEL$_2$ and EEL$^*_2$, the two EEL methods are consistently more accurate but they are not sufficiently accurate for practical applications except for the case of $(d,n)=(5,50)$. \end{longlist} \subsection{Confidence region size comparison} For the 1-$d$ examples in Table~\ref{tab1}, we computed the average interval lengths of the five practical methods OEL, EEL$_1$, BEL$^*$, AEL$^*$ and EEL$^*_2$. Table \ref{tab4} gives the average length of 1000 intervals of each method and $n$ combination for the $N(0,1)$ case. For $n=10$, the average for AEL$^*$ is not available due to occurrences of unbounded intervals; the number beside the $N/A$ is the proportion of times where this occurred. Not surprisingly, intervals with higher coverage probabilities in Table \ref{tab1} have larger average lengths. That of EEL$_1$ is the largest but it is not excessive relative to averages of other methods. As such, length is not a big disadvantage for EEL$_1$ as other methods have lower coverage probabilities. For $d>1$, sizes of the confidence regions are difficult to determine. But the relative size of an EEL region to the corresponding OEL region can be measured by the expansion factor. Table \ref{tab5} contains values of the expansion factor for 95\% EEL$_1$ regions at some combinations of $n$ and $d$. The expansion factor increases when $d$ goes up but decreases when $n$ goes up, responding to the need for more expansion in higher dimension situations and the need for less expansion when the sample size is large. \begin{table} \caption{Values of expansion factor for 95\% EEL$_1$ confidence regions}\label{tab5} \begin{tabular*}{\tablewidth}{@{\extracolsep{\fill}}l c c c c c c@{}} \hline $\bolds{n}$ &$\bolds{d=1}$ & $\bolds{d=2}$ & $\bolds{d=3}$ & $\bolds{d=5}$ & $\bolds{d=10}$ & $\bolds{d=15}$ \\ \hline 10 &1.192 &1.299 &1.390 &1.553 &$N/A$ & $N/A$ \\ 15 &1.128 &1.199 &1.260 &1.369 &1.610 & $N/A$ \\ 20 &1.096 &1.149 &1.195 &1.276 &1.457 &1.624 \\ 30 &1.064 &1.099 &1.130 &1.184 &1.305 &1.416 \\ 50 &1.038 &1.059 &1.078 &1.110 &1.183 &1.249\\ \hline \end{tabular*} \end{table} Finally, we briefly comment on the computation concerning the EEL $l^*_1(\theta)$. To compute $l^*_1(\theta)$ at a given $\theta\in \mathbb{R}^d$ which is just OEL $l(\theta')$ where $\theta'$ satisfies equation $h_n^{C}(\theta')=\theta$, we need to find the multivariate root for function $f(\theta')=h_n^{C}(\theta')-\theta$. This is seen as a nonlinear multivariate problem but it is easily reduced to a simpler \emph{univariate} problem due to the fact that $\theta' \in[\tilde {\theta}, \theta]$ (see Lemma \ref{thm2} and its proof). When using $l^*_1(\theta)$ for hypothesis testing or when simulating the coverage probabilities of the EEL confidence regions, we may use the fact that $l^*_1(\theta)\leq l(\theta)$. Hence, we can compute $l(\theta)$ first and if it is smaller than the critical value, then there is no need to compute $l^*_1(\theta)$ because it must also be smaller than the critical value. Incorporating these observations, our R code for computing the EEL runs quite fast. \section{Concluding remarks} The geometric motivation of the domain expansion method is simple: since the OEL confidence region tends to be too small, an expansion of the OEL confidence region should help to ease its undercoverage problem. What needed to be determined then are the manner in which the expansion should take place and the amount of expansion that would be appropriate. The composite similarity mapping of the present paper is an effective way to undertake the expansion as it solves the mismatch problem and retains all important geometric characteristics of the OEL contours. With the impressive numerical accuracy of the EEL$_1$, the particular amount of expansion represented by its expansion factor (\ref {3-16}) would be appropriate for general applications of the EEL method. The EEL is readily constructed for parameters defined by general estimating equations. For such parameters, we use the maximum empirical likelihood estimator (MELE) $\tilde{\theta}$ to define the composite similarity mapping in (\ref{3-15}). Under certain conditions on the estimating function which also guarantee the $\sqrt{n}$-consistency of the MELE, Lemma \ref{thm2} and all three theorems of this paper remain valid. A detailed treatment of the EEL in this setting may be found in a technical report by \citet{TsaWu13}. See also \citet{Tsa13} for an EEL for estimating equations under the simple similarity transformation (\ref{2-80}). For parameters outside of the standard estimating equations framework, the EEL on full parameter space may also be defined through a composite similarity mapping centred on the MELE, but its asymptotic properties need to be investigated for each case separately. To conclude, the simple first order EEL$_1$ is a practical and reliable method that is remarkably accurate when the effective sample size is not too small. It is also easy to use. Hence, we recommend it for real applications of the empirical likelihood method. An intriguing question that remains largely unanswered is why this first order method has such good accuracy relative to the OEL and the second order methods. Corollary \ref{cor1} and the first remark in Section \ref{sec4.2} suggested, respectively, possible asymptotic and finite sample reasons, but a more convincing theoretical explanation is needed. \begin{appendix}\label{app} \section*{Appendix} We give proofs for lemmas and theorems below. \begin{pf*}{Proof of Theorem \ref{thm1}} Assumption ($A_1$) and $n>d$ imply that, with probability 1, the convex hull of the $X_i$ is nondegenerate. This implies the OEL $l(\theta)$ has an open domain $\Theta_n\subset\mathbb{R}^d$, a condition that is required for the implementation of OEL domain expansion via composite similarity mapping. Subsequent proofs all require this condition which, hereafter, is assumed whenever $(A_1)$ is, and for brevity will not be explicitly restated. Part (i) is a simple consequence of the observation that $\gamma (n,l(\theta))\geq1$. To show~(ii), let $n$ and $\tau$ be fixed, and consider the level-$\tau$ OEL contour $c(\tau)$ defined by (\ref{3-2}). For $\theta\in c(\tau)$, $l(\theta)=\tau$. Thus, the composite similarity mapping $h_n^C$ simplifies to $h_n^C(\theta)=\tilde{\theta}+\gamma_n(\theta-\tilde{\theta})$ for $\theta\in c(\tau)$ where $\gamma_n=\gamma(n,\tau)$ is constant. This is the simple similarity mapping in (\ref{2-80}). To prove (iii), we need to show that $h_n^C\dvtx \Theta_n \rightarrow \mathbb{R}^d$ is both surjective and injective. We first show it is surjective, that is, for any given $\theta' \in\mathbb{R}^d$, there exists a $\theta'' \in\Theta_n$ such that $h_n^C(\theta'')=\theta'$. Consider the ray originating from $ \tilde{\theta}$ and through~$\theta '$. Introduce a univariate parametrization of this ray, \[ \theta= \theta(\zeta)= \tilde{\theta}+\zeta\vec{\theta}, \] where $\vec{\theta}$ is the unit vector $(\theta'-\tilde{\theta})/\| \theta'-\tilde{\theta} \|$ in the direction of the ray and $\zeta\in [0, \infty)$ is the distance between $\theta$ (a point on the ray) and $\tilde{\theta}$. Define \[ \zeta_b=\inf\bigl\{\zeta\dvtx \zeta\in[0,+\infty) \mbox{ and } \theta(\zeta) \notin\Theta_n \bigr\}. \] Then, $\theta(\zeta)\in\Theta_n$ for all $\zeta\in[0, \zeta_b)$. But $\theta(\zeta_b)\notin\Theta_n$ because $\Theta_n$ is open. It follows that $\zeta_b>0$ as it represents the distance between $\tilde{\theta }$, an interior point of the open~$\Theta_n$, and $\theta(\zeta_b)$ which is a boundary point of $\Theta_n$. Now, consider the following univariate function defined on $[0, \zeta_b)$: \[ f(\zeta) = \gamma\bigl(n,l\bigl(\theta(\zeta)\bigr)\bigr)\zeta. \] We have $f(0)=\gamma(n,l(\tilde{\theta}))\times0=\gamma(n,0)\times 0=0$. Also, by (\ref{3-e2}), \[ \lim_{\zeta\rightarrow\zeta_b} f(\zeta)= \lim_{\zeta\rightarrow\zeta _b}\gamma \bigl(n,l\bigl(\theta(\zeta)\bigr)\bigr) \zeta= \zeta_b \lim _{\zeta\rightarrow\zeta_b}\gamma\bigl(n,l\bigl(\theta(\zeta)\bigr)\bigr )=+\infty. \] Hence, by the continuity of $f(\zeta)$, for $\zeta'=\| \theta'-\tilde {\theta}\| \in[0,+\infty)$, there exists a $\zeta'' \in[0,\zeta_b)$ such that $f(\zeta'') = \zeta'$. Let $\theta''=\theta(\zeta'')$. Then $\theta'' \in\Theta_n$ since $\zeta'' \in[0,\zeta_b)$. Also, $h_n^C(\theta'') = \tilde{\theta} + \gamma(n, l(\theta''))(\theta ''-\tilde{\theta})= \theta'$. Hence, $\theta''$ is the desired point in $\Theta_n$ that satisfies $h_n^C(\theta'')=\theta'$ and $h_n^C$ is surjective. To show that $h_n^C$ is also injective, first note that for a given OEL contour $c(\tau)$, the mapping $h_n^{C}\dvtx c(\tau) \rightarrow c^*(\tau)$ is injective\vspace*{2pt} because for $\theta\in c(\tau)$, $h_n^C$ is equivalent to the similarity mapping in (\ref{2-80}) which is bijective from $\mathbb {R}^d$ to $\mathbb{R}^d$. By the partition of the OEL domain $\Theta_n$ in (\ref{3-3}), two different points $\theta_1$, $\theta_2$ from $\Theta_n$ are either [$a$] on the same contour $c(\tau)$ where $\tau=l(\theta_1)=l(\theta _2)$ or [$b$] on two separate contours $c(\tau_1)$ and $c(\tau_2)$, respectively, where $\tau_1=l(\theta_1)\neq l(\theta_2)=\tau_2$. Under [$a$], $h_n^C(\theta_1)\neq h_n^C(\theta_2)$ because $h_n^{C}\dvtx c(\tau) \rightarrow c^*(\tau)$ is injective. Under [$b$], $h_n^C(\theta_1)\neq h_n^C(\theta_2)$ also holds as (\ref{3-e1}) implies $c^*(\tau_1) \cap c^*(\tau_2)=\varnothing$. \end{pf*} \begin{pf*}{Proof of Lemma \ref{lem1}} For a fixed $\theta\in\Theta_n$, $l(\theta)$ is a fixed quantity in $[0,+\infty)$. Define an OEL confidence region for the mean using $l(\theta)$ as follows: \begin{equation}\label{3-20} {\cal C}_\theta=\bigl\{\theta'\dvtx \mbox{ $ \theta' \in\mathbb{R}^d$ and $l\bigl(\theta' \bigr) \leq l(\theta)$} \bigr\}. \end{equation} Then, ${\cal C}_\theta$ is a convex set in $\mathbb{R}^d$. See \citet{Owe90} and \citet{HalLaS90}. Since $l(\tilde{\theta})=0$ and $l(\theta) \geq0$, $\tilde{\theta}$ is in ${\cal C}_\theta$. Further, by construction, $\theta$ itself is also in~${\cal C}_\theta$. It follows from the convexity of ${\cal C}_\theta$ that for any $\alpha\in[0,1]$, \[ \theta^*=(1-\alpha)\tilde{\theta} + \alpha\theta= \tilde{\theta} + \alpha( \theta-\tilde{\theta}) \] must also be in ${\cal C}_\theta$. By (\ref{3-20}), $l(\theta^*) \leq l(\theta)$. Thus, $ l(\tilde{\theta}+\alpha(\theta-\tilde{\theta }))\leq l(\theta)$. \end{pf*} \begin{pf*}{Proof of Lemma \ref{lem2}} Since $\theta_0= h_n^C(\theta_0')=\tilde{\theta}+\gamma(n,l(\theta _0'))(\theta_0'-\tilde{\theta})$, \begin{equation}\label{3-25} \theta_0-\tilde{\theta}=\gamma\bigl(n,l\bigl(\theta_0' \bigr)\bigr) \bigl(\theta_0'-\tilde{\theta}\bigr). \end{equation} Noting that $\gamma(n,l(\theta))\geq1$, (\ref{3-25}) implies $\theta '_0$ is on the ray originating from $ \tilde{\theta}$ and through $\theta_0$ and $\|\theta_0-\tilde{\theta}\|\geq\|\theta_0'-\tilde{\theta }\|$. Hence, $\theta_0'\in[\tilde{\theta}, \theta_0]$. Without\vspace*{1pt} loss of generality, we assume that $\theta_0 \in\Theta_n$. See \citet{Owe90}. By the convexity of $\Theta_n$, $[\tilde{\theta}, \theta _0] \subset\Theta_n$. It follows from Lemma \ref{lem1} that \[ 0=l(\tilde{\theta}) \leq l\bigl(\theta_0'\bigr) \leq l(\theta_0). \] This and the fact that $l(\theta_0) = O_p(1)$ imply $l(\theta _0')=O_p(1)$. Hence, \begin{equation}\label{3-30} \gamma\bigl(n,l\bigl(\theta_0'\bigr)\bigr)=1+ \frac{l(\theta_0')}{2n}=1+O_p\bigl(n^{-1}\bigr). \end{equation} Replacing $(\theta_0'-\tilde{\theta})$ in (\ref{3-25}) with $(\theta _0'-\theta_0+\theta_0-\tilde{\theta})$, we obtain \begin{equation}\label{3-34} \bigl[1-\gamma\bigl(n,l\bigl(\theta_0'\bigr)\bigr) \bigr](\theta_0-\tilde{\theta})=\gamma\bigl(n,l\bigl(\theta _0'\bigr)\bigr) \bigl(\theta_0'- \theta_0\bigr). \end{equation} By $\tilde{\theta}-\theta_0=O_p(n^{-1/2})$, (\ref{3-30}) and (\ref {3-34}), we have $\theta_0'-\theta_0=O_p(n^{-3/2})$. \end{pf*} It may be verified using the same steps in the above proof that if the expansion factor $\gamma(n,l(\theta))$ in (\ref{3-16}) is replaced with a more general $\gamma(n,l(\theta))=1+O_p(n^{-m})$ such as that in (\ref {3-90}) where $m>0$, then Lemma \ref{lem2}(i) still holds and (ii) becomes $\theta_0'-\theta_0=O_p(n^{-m-1/2})$. In particular, under expansion factor $\gamma_s(n,l(\theta))$ in (\ref{3-80}), we also have $\theta_0'-\theta_0=O_p(n^{-3/2})$. \begin{pf*}{Proof of Theorem \ref{thm2}} Differentiating both sides of (\ref{2-58}), we obtain $\partial l(\theta)/\partial\theta=-2n\lambda^T$. By (ii) in Lemma \ref{lem2}, $\theta_0'-\theta_0=O_p(n^{-3/2})$. Applying Taylor's expansion to $l^*(\theta_0)=l(\theta_0+(\theta_0'-\theta_0))$ in (\ref {3-50}), we have \begin{equation}\label{3-61} l^*(\theta_0)=l(\theta_0)-2n\lambda( \theta_0)^T\bigl(\theta_0'- \theta_0\bigr) + o_p\bigl(\bigl\|\bigl(\theta_0'- \theta_0\bigr)\bigr\|\bigr). \end{equation} By \citet{Owe90}, $\lambda(\theta_0)=O_p(n^{-1/2})$. This and (\ref {3-61}) imply that $ l^*(\theta_0)=l(\theta_0)+ O_p(n^{-1})$, which together with (\ref {2-60}), imply Theorem \ref{thm2}. \end{pf*} For cases where $\theta_0'-\theta_0=o_p(n^{-1/2})$, we have $l^*(\theta _0)=l(\theta_0)+ o_p(1)$ which also implies Theorem \ref{thm2}. Since $\theta_0'-\theta_0=o_p(n^{-1/2})$ under expansion factor (\ref {3-90}), Theorem \ref{thm2} also holds for EEL defined by expansion factor (\ref{3-90}). \begin{pf*}{Proof of Theorem \ref{thm3}} First, note that $\gamma_s(n,l(\theta))$ in (\ref{3-80}) satisfies conditions (\ref{3-e1}) and (\ref{3-e2}). Thus it may be verified that Theorem \ref{thm1}, Lemma~\ref{lem2} and Theorem \ref{thm2} hold under the composite similarity mapping given by $\gamma_s(n,l(\theta))$. In particular, the EEL corresponding to this composite similarity mapping, $l^*_s(\theta_0)$, converges in distribution to a $\chi^2_d$ random variable. Since $\delta(n)=O(n^{-1/2})$ and $l(\theta_0')=l^*_s(\theta _0)=O_p(1)$, we have \begin{equation}\label{3-83} \bigl[l\bigl(\theta_0'\bigr)\bigr]^{\delta(n)}=1+O_p \bigl(n^{-1/2}\bigr). \end{equation} With expansion factor $\gamma_s(n,l(\theta))$ in (\ref{3-80}), equation (\ref{3-25}) becomes \[ \theta_0-\tilde{\theta}=\gamma_s\bigl(n,l\bigl( \theta_0'\bigr)\bigr) \bigl(\theta_0'- \tilde{\theta}\bigr). \] This implies \begin{eqnarray}\label{3-87-3} \theta_0'-\theta_0&=&\frac{b [l(\theta_0')]^{\delta(n)}}{2n} \bigl(\tilde{\theta}-\theta_0'\bigr)\nonumber\\[-8pt]\\[-8pt] &=&\frac{b [l(\theta_0')]^{\delta(n)}}{2n}( \tilde{\theta}-\theta_0)+\frac{b [l(\theta_0')]^{\delta(n)}}{2n}\bigl( \theta_0-\theta_0'\bigr).\nonumber \end{eqnarray} It follows from (\ref{3-83}), (\ref{3-87-3}) and $\theta_0'-\theta _0=O_p(n^{-3/2})$ that \begin{equation}\label{3-87-2} \theta_0'-\theta_0=\frac{b [l(\theta_0')]^{\delta(n)}}{2n}( \tilde{\theta}-\theta_0)+O_p\bigl(n^{-5/2} \bigr) =\frac{b}{2n}(\tilde{\theta}-\theta_0)+O_p \bigl(n^{-2}\bigr).\hspace*{-25pt} \end{equation} By assumptions ($A_1$) and ($A_2$), the OEL $l(\theta_0)$ has expansion \begin{equation}\label{3-85} l(\theta_0) = n(\tilde{\theta} -\theta_0)^T \Sigma_0^{-1}(\tilde{\theta} -\theta_0)+O_p \bigl(n^{-1/2}\bigr), \end{equation} and the Lagrange multipliers $\lambda$ at $\theta_0$ can be written as \begin{equation}\label{3-86} \lambda=\lambda(\theta_0) = \Sigma_0^{-1}( \tilde{\theta} -\theta_0) +O_p\bigl(n^{-1} \bigr). \end{equation} See \citet{HalLaS90} and \citet{DiCHalRom91N2}. By (\ref{3-87-2}), (\ref{3-86}) and the Taylor expansion (\ref{3-61}), we have \begin{eqnarray}\label{3-88} l^*_s(\theta_0)&=& l\bigl(\theta_0+\bigl( \theta_0'-\theta_0\bigr)\bigr)\nonumber\\ &=&l(\theta _0)-2n\lambda(\theta_0)^T\bigl( \theta_0'-\theta_0\bigr) + o_p \bigl(\bigl\|\bigl(\theta_0'-\theta_0\bigr)\bigr\| \bigr) \\ &=& l(\theta_0)-2n\lambda(\theta_0)^T \frac{b}{2n}(\tilde{\theta}-\theta_0)+O_p \bigl(n^{-3/2}\bigr).\nonumber \end{eqnarray} It follows from (\ref{3-85}), (\ref{3-86}), (\ref{3-88}) and $\tilde {\theta}-\theta_0=O_p(n^{-1/2})$ that \begin{eqnarray}\label{3-89} l^*_s(\theta_0) &=&l( \theta_0)-\frac{b}{n} \bigl\{n\bigl[(\tilde{\theta} - \theta_0)^T\Sigma_0^{-1} +O_p\bigl(n^{-1}\bigr)\bigr](\tilde{\theta}- \theta_0) \bigr\}\nonumber\\ &&{}+O_p\bigl(n^{-3/2}\bigr) \\ &=&l(\theta_0) \bigl[1-bn^{-1}+O_p\bigl(n^{-3/2}\bigr)\bigr].\nonumber \end{eqnarray} This proves (\ref{3-80-1}) and shows that $l^*_s(\theta_0)$ is equivalent to the BEL in the left-hand side of (\ref{2-72}). Finally, (\ref{3-81}) follows from (\ref{3-80-1}) and (\ref{2-72}). \end{pf*} \begin{pf*}{Proof of Corollary \ref{cor1}} It is convenient to view the expansion factor of the first order EEL as a special case of that for the second order EEL (\ref{3-80}) where $b=1$ and $\delta(n)=1$. The condition of $\delta(n)=O(n^{-1/2})$ imposed\vspace*{1pt} on the $\delta(n)$ in (\ref{3-80}) is not needed here. Noting that $l^*(\theta_0)=l(\theta_0)+ O_p(n^{-1})$ and $l(\theta _0')=l^*(\theta_0)$, equation (\ref{3-83}) in the proof Theorem \ref {thm3} is now \[ l\bigl(\theta_0'\bigr)=l(\theta_0)+ O_p\bigl(n^{-1}\bigr). \] Thus equation (\ref{3-87-2}) becomes \[ \theta_0'-\theta_0=\frac{l(\theta_0)}{2n}( \tilde{\theta}-\theta_0)+O_p\bigl(n^{-5/2} \bigr). \] Using the above equation and following the steps given by (\ref{3-85}) to (\ref{3-89}), we obtain Corollary \ref{cor1}. \end{pf*} \end{appendix} \section*{Acknowledgements} Min Tsao is the principal author of this paper. Fan Wu contributed the simulation results in Section \ref{sec4}. We thank two anonymous referees and an Associate Editor for their valuable comments which have led to improvements in this paper. We also thank the Editor, Professor Peter Hall, and the Associate Editor for their exceptionally quick handling of this manuscript.
\section{Computational details} First-principles density functional theory calculations were carried out using the Vienna \textit{ab initio} simulations package (VASP) \cite{kresse1994norm, kresse1999ultrasoft} within the projector-augmented wave method \cite{Bloechl_PAW}. \textbf{Atomic structure.} For calculations of the atomic structure of bulk and 2D K$_2$CuF$_4$\, the local density approximation (LDA) was employed to the exchange-correlation potential \cite{KohnShamLDA} because we found that this perfectly reproduces the experimentally known Jahn-Teller distortions. A k-mesh of 8x8x1 (5x5x5) grid points was used for single layer (bulk) K$_2$CuF$_4$\ and a kinetic energy cutoff of 700 eV (single layer) and 500 eV (bulk). We used the experimentally known unit cell volume of the bulk system and the same lattice parameters for the 2D system (lattice constants $a=b=5.8655$ \AA\ \cite{Hidaka1983343}). The atomic positions were relaxed until forces were smaller than 0.02 eV/\AA. \textbf{Cleavage energies, elastic theory.} The binding between K$_2$CuF$_4$\ layers is partially due to van der Waals interaction. LDA is known to reproduce the interlayer distances in weakly bound layered systems while GGA fails which we also see in our calculations. Therefore, we performed calculations using LDA. We also employed the LDA with van der Waals corrections within the DFT-D2 method of Grimme \cite{grimme2006semiempirical}. Local Coulomb interactions do not play a role for the binding of K$_2$CuF$_4$\ layers which we tested using the LDA+$U$ method \cite{LichtensteinLDAU}. \begin{figure*}[htb] \includegraphics[width=0.3\textwidth]{Fig1_sm.eps} \caption{Scheme of the geometry used to simulated cleavage fracture. Unit cells are marked by grey lines.} \label{fig:cleavagestruct} \end{figure*} For calculations of the bilayer binding energy, we simulated a bilayer system and relaxed the structure to the equilibrium configuration. Afterwards, we lifted one layer far to the vacuum. The binding energy was then calculated by the difference of total energies in the equilibrium position and at far interlayer distance. For cleavage energy calculations, we used a bulk unit cell as shown in Fig. \ref{fig:cleavagestruct}. One cell contained two K$_2$CuF$_4$\ layers, which ensured a distance between two fractures of more than 10 \AA\ and the fracture was simulated by increasing the c parameter of the cell. For calculations of the elastic energies, LDA+$U$ calculations were performed with the optimized lattice constant $a=b=5.561$ \AA. This lattice constant differs by about 5\% from the experimental value but also reproduces Jahn-Teller distortions and ferromagnetism. \textbf{Magnetism and electronic structure.} For electronic structure calculations and calculations of magnetic moments, we account for local Coulomb interactions in Cu $d$ orbitals by making use of the LDA+$U$ method. This also reproduced best the experimentally known magnetism and magnetic coupling of the bulk. In the simple rotationally invariant formulation of Dudarev \cite{Dudarev}, an effective Hubbard parameter $U_{\rm eff}= U_{\rm H} - J_{\rm H}$ is defined, thus only the difference between the average Hubbard Coulomb repulsion $U_{\rm H}$ and Hund's rule exchange $J_{\rm H}$ is important. We chose $U_{\rm eff}=7.03$ eV, a similar value was used to calculate KCuF$_3$ \cite{LichtensteinLDAU}. We emphasize here that the band gap strongly depends on $U_{\rm eff}$, for LDA without Coulomb interactions we find only 0.3 eV instead of 3.2 eV using LDA+U with $U_{\rm eff}=7.03$ eV. For calculations of the magnetic crystalline anisotropy, we included spin-orbit coupling in the calculations and used a denser k-mesh (18x18x1). With the parameter $U_{\rm eff}$, we avoid the problem of a strong anisotropy dependence of the chosen Hund's exchange $J_{\rm H}$ \cite{bousquet2010j}. The anisotropy parameter might slightly differ in reality, but note that the magnetic anisotropy $\eta$ only enters logarithmically in the formula for $T_{\rm KT}$. We also stress here that the shape anisotropy does not significantly contribute to the overall magnetic anisotropy: while the MCAE energy is on the order of $10^{-4}$ eV, we estimate shape anisotropy contributions to be much smaller ($\sim 10^{-6}$ eV). \section{Band Structure and orbital ordering} \begin{figure}[htb] \includegraphics[width=0.99\textwidth]{Fig2_sm.eps} \caption{a), b) Band structure of single layer K$_2$CuF$_4$\ in the spin up (a) and spin down (b) channel. Red "fat bands" depict the $d_{\rm{z^2-r^2}}$ contribution to the bands, whereby the thickness is proportional to the magnitude of the contribution. c) Contour plot of the hole density in the Cu-F$_2$ plane of the hole state at $\sim 3.5$ eV shown in b) in a window between 0.35--0.55e/unit cell. Red dots denote Cu atoms. } \label{fig:bands} \end{figure} In the manuscript, we state that the magnetism mainly originates from $d_{\rm z^2-r^2}$ orbitals. To learn more about this, we look at the band structure of K$_2$CuF$_4$\ which is depicted in Fig. \ref{fig:bands}a) and b). A comparison of the up and down channels shows that the band gap is direct. Many bands are present below the Fermi level because of the relatively large unit cell, but here we focus on Cu $d_{\rm{z^2-r^2}}$ states. The red fat bands visualize the Cu $d_{\rm{z^2-r^2}}$ character of the bands whereby the thickness of the red band depicts the magnitude of the contribution. One can see a different distribution of spectral weight in both channels. Most importantly, the bands in the spin up channel exhibit a pronounced $d_{\rm{z^2-r^2}}$ contribution at bands below -6 eV that is not present in the down channel. Here, on the contrary, the main spectral weight is located at the hole state. This hole state is responsible for a ferromagnetic ordering through the following mechanism: the Jahn-Teller distortions of F atoms lead to an orbital ordering effect at Cu sites with an alternating occupation of $d_{\rm{z^2-x^2}}$ and $d_{\rm{z^2-y^2}}$ orbitals of neighbored Cu atoms. Thus, the ferromagnetism in 2D K$_2$CuF$_4$\ is orbital-selective and related with the $d_{\rm{z^2-x^2}}$ and $d_{\rm{z^2-y^2}}$ subspace. The orbital ordering is visualized in Fig. \ref{fig:bands}c. We tested explicitly the importance of Jahn-Teller distortions. Indeed, in a fictitious undistorted geometry, we find from our calculations that an antiferromagnetic coupling of spins is energetically slightly more favorable than a ferromagnetic coupling. Only the distortions of F atoms lead to an alternating occupation of the $d_{\rm{z^2-x^2}}$ and $d_{\rm{z^2-y^2}}$ orbitals which finally allows for a ferromagnetic ordering. This is also in line with a scenario proposed by Khomskii and Kugel \cite{khomskii1973orbital} to explain ferromagnetic ordering in bulk K$_2$CuF$_4$. \section{Membrane properties} In Ref. \cite{booth2008macroscopic}, the peculiar elastic properties of graphene were discussed. The authors showed that these allow for fabrication of long graphene beams which need support from only one side due to their extraordinary in-plane stiffness. In the spirit of Ref. \cite{booth2008macroscopic}, we estimate the bending of a K$_2$CuF$_4$\ flake using elastic theory for large flakes of similar length $l$ and width $w$. The typical out-of-plane deformation $h$ of a flake can be estimated by the formula $h/l \approx \left(\rho g l / Y_{\rm 2D}\right)^{1/3}$, with $l$ being the length of the flake, and $\rho=2.1 \times 10^{-6} kg/m^2$ the density. For a large flake of length of the order $l \approx 100$ $\mu$m, we obtain $h/l \approx 10^{-3}-10^{-4}$. For graphene, $h/l \approx 10^{-4}$ for flakes of this size. Thus, even for large flakes, 2D K$_2$CuF$_4$\ is able to withstand its own weight, significantly large extra loads or vibration during handling.
\section{Introduction} The Euclidean minimum is a numerical invariant which measures how elements of a number field can be approximated by algebraic integers. Its study is a classical topic in algebraic number theory, going back to Minkowski and Hurwitz. Many mathematicians have studied its properties; among the others, we can cite Barnes, Swinnerton-Dyer, Cassels, Davenport and Fuchs, but this list is far from being exhaustive. This research area is still very active nowadays, and many problems are still open. We can mention for example the works of Van der Linden, Cerri and Bayer-Fluckiger. One of the central questions is to find upper bounds for Euclidean minima. We refer to~\cite{l} for a nice survey of the topic. The Euclidean minima can also be defined in the function field case, replacing the absolute value by the degree. Not much is known in this situation. The existence of Euclidean function fields was studied by Armitage, Markanda, Madam, Queen and Smith but Euclidean minima never explicitly appear in their work. The aim of the present paper is to investigate this question more closely, trying to obtain bounds analoguous to the known bounds in the number field case. The paper is organized as follows: section 2, which is of geometric flavour is mainly inspired from the guiding example of polynomials: it is well-known that, given a field $k$, the ring $k[T]$ is euclidean with respect to the degree. Now, from a geometric point of view, the field $k(T)$ is just the function field of the projective line, and $k[T]$ is the subring consisting of functions which are regular outside the point at infinity. Starting from this observation, we try to generalize the construction: we first of all consider a (smooth, proper) curve $\mathcal C$ over a field $k$, which is assumed to be algebraically closed. We then consider a finite subset $S$ of $\mathcal C$ and define a degree function, denoted by $\deg_S$, for rational functions on the curve. The ring $\mathcal O_S$ is the subset of the field $K=k(\mathcal C)$ consisting of rational functions which are regular outside $S$. In other words, the affine curve $\mathcal C\setminus S$ is just the spectrum of $\mathcal O_S$. For $\mathcal C=\Bbb P^1$ and $S=\{\infty\}$, we then find $K=k(T)$ and the ring $\mathcal O_S=k[T]$ is an univariate polynomial ring over $k$, the function $\deg_S$ coinciding with the usual degree for polynomials. The (logarithmic) Euclidean minimum of $K$ with respect to $S$, denoted by $M_S(K)$, is then defined as \[M_S(K)=\max_{x\in K}\min_{y\in\mathcal O_S}\{\deg_S(x-y)\}.\] This integer measures how well a rational function can be approximated by an element of $\mathcal O_S$. In particular, the ring $\mathcal O_S$ is Euclidean with respect to the function $\deg_S$ if and only if the inequality $M_S(K)<0$ holds. The main result of the section relates $M_S(K)$ to a geometric invariant we introduce here: the degree of speciality $\mu(S)$ of $S$, which, roughly speaking, describes the behaviour of differential forms on $\mathcal S$ having no poles outside $S$. In Theorem 2, we prove that we have the inequality $M_S(K)\leq\mu(S)$. The main ingredient of the proof is the Riemann-Roch theorem. It then follows that $M_K(S)$ is less than or equal to $2g-1$, where $g$ denotes the genus of $\mathcal C$ (this bound again follows from Riemann-Roch formula), providing an upper bound which can be considered, as we will show later, as the precise analogue of the results obtained in the number field case in~\cite{b}. The end of the section is devoted to the special case where the set $S$ is a singleton. In this situation, we can prove (cf. Proposition 3) that we have in fact the equality $M_S(K)=\mu(S)$ and, moreover, that $\mu(S)$ is the greatest integer appearing in the Weierstrass gap sequence of $S$. In particular, we obtain the lower bound $M_S(K)\geq g$ and we provide examples where $M_S(K)$ is explicitly computed (hyperelliptic curves, \'etale covers of the affine line, classical curves). As stated above, in section 3 we show that Theorem 2 can be considered as a (logarithmic) analogue of the general bounds obtained in the number field case. It is important to notice that there is a crucial difference between number fields an function fields. Indeed, a number field can be obtained in a unique way as an extension of the rationals. In particular, its degree and discriminant are uniquely defined global objects. For a function field $K$, things are quite different, since it can be realized in many (infinite) ways as an extension of the field $k(T)$. Its degree and discriminant are no more intrinsic invariants and strongly depend on this realization. This is the main reason why in section 2 we tried to obtain results avoiding this arbitrary choice. But now, if we want to find a parallel between the present results and the number field case couterpart, we have to consider the field $K$ as a finite extension of the field $k(T)$. In this setting, we slightly modify the definition of the Euclidian minimum, now simply denoted by $M(K)$, which turns out to be just a special case of the one defined before (for a particular choice of the set $S$). In theorem 8, we obtain an upper bound for $M(K)$ only depending on the degree and on the discriminant of the extension $\mathcal O_K/k[T]$ (here, the ring $\mathcal O_K$ is the integral closure of $k[T]$ in $K$, which, once again coincides withnthe ring $\mathcal O_S$ previously defined). In fact, Theorem 8 is just a weaker version (and a direct consequence) of Theorem 2, the main ingredient for its proof being the Hurwitz formula. The main limitation is that we have to suppose that the ramification above the point at infinity is tame. A more general, and sharper result could be obtained by considering the full ramification of the extension $K/k(T)$ (taking account of the behaviour above infinity) but we decided to state it in a form more similar to the number field case. Until now, we worked in a geometric setting, assuming that the base field $k$ is algebraically closed. In the last section of the paper, we drop this assumption and investigate the more general case of a perfect base field. Most of the constructions and results remain valid in this situation: the Euclidean minimum and the index of speciality are defined in the same way and the inequality of Theorem 2 still holds. However, it is no longer true that $\mu(S)$ is less than or equal to $2g-1$. Moreover, it actually turns out that the Euclidean minimum depends on the base field. Indeed, we explicitly treat the case of hyperelliptic curves and in Theorem 10 we show that in this case $M(K)$ can actually be equal to $2g$, which is impossible if $k$ is algebraically closed. \ We would like to thank Eva Bayer-Fluckiger, who originally motivated and then constantly encouraged this joint work. Her results in the number field case have been the main inspiration of the paper. The second and third authors also express their thanks for her invitation at the EPFL, where most of the results were obtained. \section{Geometric approach} \subsection{Settings and notation} In the following, $\mathcal C$ denotes a smooth, projective curve of genus $g$, defined over an algebraically closed field $k$. Set $K=k(\mathcal C)$ and let $v_P$ (resp. $\mathcal O_P$) be the valuation of $K$ (resp. the local ring) associated to a point $P$ of $\mathcal C$. For any finite, nonempty subset $S$ of $\mathcal C$, let $\mathcal O_S$ be the subring of $K$ consisting of rational functions regular outside $S$. For any rational function $x\in K^\times$, consider the integer $\deg_S(x)$ defined by \[\deg_S(x)=-\sum_{P\in S}v_P(x).\] For simplicity, we furthermore set $\deg_S(0)=-\infty$. It is easily checked that for any element $x\neq0$ of $\mathcal O_S$, we have the inequality $\deg_S(x)\geq0$, which is an equality if and only if $x$ belongs to $\mathcal O_S^\times$. By construction, for any $x,y\in K$, we have the relations \begin{itemize} \item $\deg_S(xy)=\deg_S(x)+\deg_S(y)$, \item $\deg_S(x+y)\leq\max\{\deg_S(x),\deg_S(y)\}$. \end{itemize} \begin{rem} For $\mathcal C=\Bbb P^1$ and $S=\{\infty\}$, the ring $\mathcal O_S$ can be identified with the polynomial ring $k[T]$ and, for any $x\in\mathcal O_S$, the integer $\deg_S(x)$ is the usual degree of a polynomial. \end{rem} The {\it Euclidean minimum} of $K$ (with respect to $S$) is the integer defined as \[M_S(K)=\max_{x\in K^\times}\min_{y\in\mathcal O_S}\{\deg_S(x-y)\}.\] The inequality $M_S(K)<0$ holds if and only if the ring $\mathcal O_S$ is Euclidean with respect to the Euclidean function $\deg_S$. \subsection{The index of speciality} Following the usual notation, for any divisor $D$ on $\mathcal C$, let $\Omega(D)$ be the $k$-vector space of differential forms $\omega$ on $\mathcal C$ satisfying the relations $v_P(\omega)+v_P(D)\geq0$ for any $P\in\mathcal C$ (in~\cite{s}, this $k$-vector space is denoted by $\Omega(-D)$, we nevertheless adopt the more recent and usual notation). Denote by $\Bbb Z[S]$ the free abelian group generated by the elements of $S$, i.e. the subgroup of $\mbox{Div}(\mathcal C)$ consisting of divisors whose support is contained in $S$. The {\it index of speciality} of $S$ is the integer $\mu(S)$ defined by \[\mu(S)=\min\{\deg(D)\,\,|\,\,D\in\Bbb Z[S]\mbox{ and }\Omega(-D)=0\}.\] \begin{rem} The dimension of the $k$-vector space $\Omega(D)$ only depending on the linear equivalence of the divisor $D$, we may replace $\Bbb Z[S]$ by the subgoup of $\mbox{Pic}(\mathcal C)$ generated by $S$. \end{rem} \begin{lemm} We have the inequalities \[g-1\leq\mu(S)\leq2g-1.\] \end{lemm} \begin{proof} Let $D$ be a divisor on $\mathcal C$ of degree greater than or equal to $2g-1$ and suppose that there exist an element $\omega\in\Omega(-D)$ different from $0$. In this case, we obtain the relations \[2g-2=\deg(\omega)\geq\deg(D)\geq2g-1,\] a contradiction. We therefore have $\Omega(-D)=0$. Since the field $k$ is algebraically closed, there exists an element $D\in\Bbb Z[S]$ of degree $2g-1$, which implies that $\mu(S)\leq2g-1$. Suppose now that $\deg(D)<g-1$ and denote by $L(D)$ the $k$-vector space of rational functions $x\in K$ for which $v_P(x)+v_P(D)\geq0$. The Riemann-Roch formula leads to the relations \[\dim_k\Omega(-D)=\dim_kL(D)-\deg(D)+g-1>0,\] and the second inequality follows. \end{proof} \subsection{An upper bound for Euclidean minima} We now prove the main result of the section. We will later see that (and explain why) it is analogous to the bound obtained in the number field case. \begin{theo} We have the inequality \[M_S(K)\leq\mu(S).\] \end{theo} \begin{proof} Let $A_K$ denote the adele ring of $K$, i.e. an element $r\in A_K$ is a collection $r=(r_P)_{P\in\mathcal C}$ of rational functions such that $r_P\in\mathcal O_P$ for almost all $P$ in $\mathcal C$ (see~\cite{s}, Chapter II, \S5 for general definitions and properties of adele rings on curves). For any divisor $D$ on $\mathcal C$, let $A_K(D)$ denote the $k$-vector space of adeles $r$ such that $v_P(r_P)+v_P(D)\geq0$ for all $P\in\mathcal C$. By Serre duality, the quotient $I(D)=A_K/(A_K(D)+K)$ is canonically isomorphic to the dual of the $k$-vector space $\Omega(-D)$. Let now $D\in\Bbb Z[S]$ be a divisor such that $\deg(D)=\mu(S)$. Fix a rational function $x\in K^\times$ and consider the adele $r=(r_P)$ defined by \[r_P=\begin{cases} x\quad\mbox{ for }P\in S,\\ 0\quad\mbox{ otherwise.} \end{cases}\] By construction, we have $\Omega(D)=0$ and therefore $I(D)=0$. In particular, there exist $y\in K$ and $r'\in A_K(D)$ such that $r=r'+y$. It then follows that the poles of the rational function $y$ are contained in $S$, and thus we have $y\in\mathcal O_S$. Furthermore, from the definition of $r$, we have the inequalities \[v_P(x-y)+v_P(D)\geq0\] for any $P\in S$, which finally leads to \[\deg_S(x-y)\leq\deg(D)=\mu(S).\] We therefore have the inequality $M_S(K)\leq\mu(S)$. \end{proof} \subsection{A special case. Weierstrass gaps} Though we are not able to prove it in full generality, it is reasonable to believe that the inequality of Theorem 2 is in fact an equality. We will now see that it is actually true if the set $S=\{P\}$ is reduced to a single point. Until the end of this section, we restrict to this situation. The study of the index of speciality, simply denoted by $\mu(P)$ in this case, is closely related to the {\it Weierstrass gap sequence} of $P$, we therefore briefly recall some basic facts of the theory: setting $\ell(nP)=\dim_kL(nP)$ and $i(nP)=\dim_k\Omega(nP)$, we have the relations \[\ell(0)=1,\ell(nP)\leq\ell((n+1)P)\leq\ell(nP)+1\mbox{ and }\ell((2g-1)P)=g.\] Furthermore, for $n\geq{2g-1}$, we have the identity $\ell(nP)=n-g+1$. In terms of differential forms, the Riemann-Roch formula leads to the relations \[i(0)=g,i(-nP)\geq i(-(n+1)P)\geq i(-nP)-1\mbox{ and }i((1-2g)P)=0.\] A {\it Weierstrass gap} of $P$ is an integer $n$ such that $l(nP)=l((n-1)P)$, which means that there does not exist a rational function $x\in K$ having a unique pole at $P$ of exact order $n$. Equivalently, we have the identity $i(-nP)=i((1-n)P)-1$, which implies that there exists a regular differential form having a zero at $P$ of order $n-1$. It then follows that there are exactly $g$ gaps $0<n_1<n_2\cdots<n_g<2g$. The gap sequence is just the collection $(n_1,\dots,n_g)$. Notice that, by definition, the index of speciality $\mu(P)$ is just the integer $n_g$, the greatest Weierstrass gap. Given the curve $\mathcal C$, there exist a sequence $N=(n_1,\dots,n_g)$ such that all but finitely many points of $\mathcal C$ have $N$ as Weierstrass gap sequence. The points having a Weierstrass gap sequence different from $N$ are called {\it Weierstrass points}, while the others are {\it ordinary points}. If $N=\{1,2,\dots,g\}$, we say that $\mathcal C$ is a {\it classical curve}. In characteristic $0$, any curve is classical, but this is false in positive characteristic (though it is the case if the characteristic of the base field is large enough with respect to the genus of the curve). \begin{prop} If $S=\{P\}$ is reduced to a single point, then we have the identity \[M_S(K)=\mu(S).\] In particular, we have the inequalities \[g\leq M_S(K)\leq2g-1.\] \end{prop} \begin{proof} Given a point $Q\neq P$ of $\mathcal C$, the Riemann-Roch theorem implies that there exists a rational function $x\in K$ having no poles outside the set $\{P,Q\}$ and such that $v_P(x)=-\mu(P)$, so that $\deg_S(x)=\mu(S)$. Now, the relation $\deg_S(x-y)<\mu(x)$, with $y\in\mathcal O_S$, would imply that the rational function $y$ has a unique pole at $P$ of order $\mu(P)$, which is impossible since $\mu(P)$ is a Weierstrass gap. The two inequalities again follow from the fact that $\mu(P)$ is a Weierstrass gap. \end{proof} \begin{exa} Once again, taking $\mathcal C=\Bbb P^1$ and $S=\{\infty\}$, so that $K=k(T)$ and $\mathcal O_S=k[T]$, we find the inequality $M_S(K)<-1$, recovering the well known fact that the ring $k[T]$ is Euclidean with respect to the degree. \end{exa} We close this section with three examples where we can actually compute the Euclidean minimum, and show that for fixed $g$, the bounds of Proposition 3 are attained. \begin{coro} Let $\mathcal C$ be an hyperelliptic curve and suppose that $P$ is a Weierstrass point on it. We then have the identity \[M_S(K)=2g-1.\] \end{coro} \begin{proof} A well-know result on hyperelliptic curves asserts that $P$ is a Weierstrass point if and only its gap sequence is $(1,3,\dots,2g-1)$ (see for example~\cite{f}, Exercice 8.37, where the characteristic of the base field is assumed to be $0$, but everything remains true in full generality), from which we deduce the relation $\mu(P)=2g-1$. The result is then a direct consequence of Proposition 3. \end{proof} \begin{rem} The above result is actually true for superelliptic curves, i.e. curves $\mathcal C$ defined by affine equations $Y^n=f$, where $f\in k[X]$ and the characterictic of $k$ does not divide $n$. More precisely, if $r$ denotes the degree of $f$, then there is a canonical degree $n$ cover $\mathcal C\to\Bbb P^1$ totally ramified at $r$ or $r+1$ points. If $P$ is one of them, it is then easily checked that $2g-1$ belongs to its gap sequence and is therefore the greatest Weierstrass gap. In particular, setting $S=\{P\}$, we find $\mu(S)=2g-1$ and Proposition 3 asserts that we have the identity $M_S(K)=2g-1$. \end{rem} \begin{coro} Suppose that there exists an \'etale cover $\pi:\mathcal C\setminus S\to\Bbb A^1$. We then have the identity \[M_S(K)=2g-1.\] \end{coro} \begin{proof} Consider a differential form $\omega$ on $\Bbb P^1$ having a double pole at $\infty$ (it is unique, up to multiplication by an element of $k^\times$). Since the canonical divisor of $\Bbb P^1$ has degree $-2$, it follows that $\omega$ is regular and non-vanishing outside $\infty$. In particular, the cover $\pi$ being \'etale outside $\infty$, the pull-back $\pi^*\omega$ is regular and nowhere vanishing outside $P=\pi^{-1}(\infty)$, so that $\omega$ has a unique zero at $P$, of order $2g-2$, and thus the integer $2g-1$ belongs to the gap sequence of $P$, which concludes the proof. \end{proof} \begin{coro} Suppose that the curve $\mathcal C$ is classical and that $P$ is not a Weierstrass point. We then have the identity \[M_S(K)=g.\] \end{coro} \begin{proof} It is again a consequence of Proposition 3, since in this case, the gap sequence of $P$ is $(1,2,\dots,g)$, leading to $\mu(P)=g$. \end{proof} \section{The analogy with the number field case} \subsection{Upper bounds for the Euclidean minima of number fields} As it is stated in the previous paragraph, Theorem 2 could seem quite far from the kind of results obtained in the number field case. We will now show that it actually leads to an upper bound for the Euclidean minima which is, {\it mutatis mutandi}, a complete analogy with the known results and conjectures. We first of all briefly summarize the number field case: let $K$ be a number field, i.e. a finite extension of the field $\Bbb Q$ of rational numbers, and denote by $\mathcal O_K$ its ring of integer (the integral closure of $\Bbb Z$ in $K$). The {\it Euclidean minimum} of $K$ is the real number $M(K)$ defined as \[M(K)=\sup_{x\in K}\inf_{y\in\mathcal O_K}\{|N_{K/\Bbb Q}(x-y)|\},\] where $|\cdot|$ denotes the usual (Archimedean) absolute value. The study of Euclidean minima is a hard and classical topic in algebric number theory. One of the main results, obtained by E. Bayer-Fluckiger in~\cite{b} in 2006, is the following general upper bound for Euclidean minima: \begin{theo}[Bayer-Fluckiger, 2006] Let $n$ be the degree of the extension $K/\Bbb Q$ and denote by $d_k$ its discriminant. We then have the inequality \[M(K)\leq2^{-n}|d_k|.\] \end{theo} \subsection{The function fields analogue} The aim of this section is to give a bound for the Euclidean minima of function fields, analogous to Theorem 7. We first of all fix the notation and settings: the field $\mathcal Q=k(T)$ will play the role of the field $\Bbb Q$, the ring $\mathcal Z=k[T]$ replacing $\Bbb Z$. We denote by $K$ a finite, separable extension of $\mathcal Q$ of degree $n$ and by $\mathcal O_K$ the integral closure of $\mathcal Z$ in $K$. The discriminant of the extension $\mathcal O_K/\mathcal Z$, denoted by $d_K$, considered as an ideal of $\mathcal Z$, is generated by a polynomial $f$ and we set $\deg(d_K)=\deg(f)$. The {\it Euclidean minimum} $M(K)$ is the integer defined as \[M(K)=\max_{x\in K}\min_{y\in\mathcal O_K}\{\deg(N_{K/\mathcal Q}(x-y))\},\] where we have set $\deg=\deg_\infty$ (the extention to $\mathcal Q$ of the usual degree of a polynomial, cf. the first remark of \S2). \begin{rem} Before continuing, we must stress on two important points: \begin{itemize} \item The rings $\Bbb Z$ and $\mathcal Z$ are both Euclidean. In the first case, the associated Euclidean function is the absolute value, which is multiplicative i.e. $|xy|=|x|\cdot|y|$, while in the latter case the Euclidean function is the degree, which is additive i.e. $\deg(xy)=\deg(x)+\deg(y)$. The Euclidean minimum is then defined in a similar way, but because of this difference, the bounds for function fields should be considered as a 'logarithmic' analogue of those obtained in the number field case. \item For function fields case, the discriminant $d_K$ does not take account of the ramification above $\infty$ (the valuation associated to the prime element $T^{-1}$); this is the main difference with number fields, where the discriminant completely describes the ramification of the extension. \end{itemize} \end{rem} We can now state and prove the main result of this section, which is a weaker form of Theorem 2 and can be considered as the analogue of Theorem 7. \begin{theo} The assumptions and hypothesis being as above, suppose that the extension $K/\mathcal Q$ is tamely ramified above $\infty$. We then have the inequality \[M(K)\leq\deg(d_K)-n.\] \end{theo} \begin{proof} From a geometric point of view, the extension $K/\mathcal Q$ corresponds to a finite cover $\pi:\mathcal C\to\Bbb P^1$. Setting $S=\pi^{-1}(\infty)$, the ring $\mathcal O_K$ then coincides with the ring $\mathcal O_S$ defined in the first section and, taking spectra, the extension $\mathcal O_K/\mathcal Z$ defines a finite cover $\mathcal C\setminus S\to\Bbb A^1$, which is just the restriction of $\pi$ to the open set $\Bbb A^1\subset\Bbb P^1$. For any rational function $x\in K$, a straightforward computation leads to the relation \[\deg(N_{K/\mathcal Q}(x))=\deg_S(x).\] In particular, we find the identity \[M(K)=M_S(K).\] Let $R$ be the ramification divisor of $\pi$ and $B=\pi_*R$ its branch divisor (see~\cite{h}, \S IV.2 for general definitions and properties). We then have the identity $\deg(R)=\deg(B)$ and $d_K$, considered as a divisor, is just the restriction of $B$ to $\Bbb A^1$. Moreover, if $g$ denotes the genus of $\mathcal C$, the Hurwitz formula (\cite{h}, Corollary 2.4 of Chapter IV) reads \[2g-2=\deg(B)-2n.\] Now, if the ramification above $\infty$ is tame, we then have the inequality \[\deg(B)\leq\deg(d_K)+n-1.\] In this case, combining Lemma 1 and Theorem 2, we finally obtain the relations \[M(K)\leq2g-1=\deg(B)-2n+1\leq\deg(d_K)-n,\] which concludes the proof. \end{proof} \begin{theo} Suppose that the extension $K/\mathcal Q$ is totally ramified above $\infty$ and that the characteristic of $k$ does not divide $n$. We then have the inequalities \[\frac12\left(\deg(d_K)-n-1\right)\leq M(K)\leq\deg(d_K)-n.\] \end{theo} \begin{proof} Following the notation of the proof of Theorem 8, the ramification above $\infty$ being tame, we have the identity \[\deg(B)=\deg(d_K)+n-1\] Now, since the set $S$ is reduced to a point, Proposition 3 asserts that we have the inequalities \[g-1\leq M(K)\leq2g-1\] and the Hurwitz formula finally leads to the relations \[\frac12\left(\deg(d_K)-n-1\right)\leq M(K)\leq\deg(d_K)-n,\] as desired. \end{proof} \section{The case of a general base field} \subsection{General base fields} Until now, we assumed that the base field $k$ is algebraically closed. We now drop this assumption and we only suppose that $k$ is perfect. Most of the constructions of the previous sections actually apply in this more general situation, with only minor changes: first of all, going back to section 2 and setting $G_k=\mbox{Gal}(\bar k/k)$, we must assume that the set $S\subset\mathcal C(\bar k)$ is stable under the action of $G_k$. The definition of the $S$-degree $\deg_S$ and of the Euclidean minima are exactly the same. Concerning the index of speciality, we just replace the group $\Bbb Z[S]$ by the subgroup $\Bbb Z[S]^{G_k}$ of $G_k$-invariants. In particular, it turns out that $\mu(S)$ can be strictly bigger than in the case of an algebraically closed field $k$. \begin{exa} Suppose that $\mathcal C$ is an elliptic curve and that $S=\{P,Q\}$ consists of two $G_k$-conjugated points. Setting $D=P+Q$, we then find $\Bbb Z[S]^{G_k}=\Bbb Z[D]$ an therefore $\mu(S)=\deg(D)=2$. On the other hand, over $\bar k$, setting $F=P-Q\in\Bbb Z[S]$, we find $\Omega(-F)=0$ and it actually turns out that $\mu(S)=\deg(F)=0$. \end{exa} The result in Theorem 2 still holds. However, in this case we no longer have the inequality $\mu(S)\leq2g-1$, which was of crucial importance in section 3. Nevertheless, if $S$ is reduced to a point, then everything works perfectly. \begin{rem} If the residual degrees (i.e. the degrees of the fields of definition) of the elements of $S$ are globally coprime, which is the same assumption made in~\cite{a}, then we still have the inequality $\mu(S)\leq2g-1$ and all the results of the previous sections are true. \end{rem} \subsection{An example: hyperelliptic curves} We close the paper with an example where the Euclidean minimum can be explicitly computed, showing that its behaviour is different than in the case of an algebraically closed field. In the following, we assume that the characteristic of $k$ is different from $2$. Let $g$ be a positive integer, fix a polynomial $f\in k[X]$ of degree $2g+2$ whose leading coefficient is not a square in $k^\times$ and consider the hyperelliptic curve $\mathcal C$ defined by the affine equation $Y^2=f$. By construction, its genus is equal to $g$ and there exist two points $P$ and $Q$ at infinity which are permuted by $G_k$. There is a canonical double cover $\mathcal C\to\Bbb P^1$ and, setting $S=\{P,Q\}$, the ring $\mathcal O_S=k[X,Y]$ is an extension of $k[X]$ of degree $2$ which coincides with the integral closure of $k[X]$ in $K$. Moreover, as noticed in the proof of Theorem 8, for any rational function $x\in K$, we have the identity \[\deg_S(f)=\deg(N_{K/\mathcal Q}(x)),\] where we have set $\mathcal Q=k(X)$. As in the previous section, we simply write $M(K)$ instead of $M_S(K)$. \begin{theo} The notation and assumptions being as above, we have the identity \[M(K)=2g.\] \end{theo} \begin{proof} Any element of $x\in K$ can be uniquely written as a sum $x=a+Yb$, with $a,b\in\mathcal Q$, and we have the relation \[N_{K/\mathcal Q}(x)=a^2-fb^2.\] We want to minimize the degree (meant as $\deg_{\infty}$) of the rational function \[N_{K/\mathcal Q}(x-y)=(a-c)^2-f(b-d)^2\in\mathcal Q,\] where $y=c+dY$ is an element of $\mathcal O_S$, with $c,d\in k[X]$. The assumption on the leading coefficient of $f$ leads to the relation \[\deg\left((a-c)^2-f(b-d)^2\right)=2\max\{\deg(a-c),g+1+\deg(b-d)\}.\] Once again, we have set $\deg=\deg_\infty$ (in particular, the degree of a rational function is not the number of its poles, counted with their multiplicity). Now, as remarked in the first section, the fact that the ring $k[X]$ is Euclidean with respect to the degree implies that there exist two elements $c,d\in k[X]$ (which are in fact unique) such that $\deg(a-c)\leq-1$ and $\deg(b-d)\leq-1$. We therefore find the inequality \[\min_{y\in k[X]}\{\deg(N_{K/\mathcal Q}(x-y))\}\leq2g\] and thus $M(K)\leq2g$. Finally, it is easily checked that for the rational function $x=YX^{-1}\in K$ we have the identity \[\min_{y\in k[X]}\{\deg(N_{K/\mathcal Q}(x-y))\}=2g,\] so that $M(K)=2g$, which concludes the proof. \end{proof}
\section*{Introduction}\label{sec:introduction} The goal of this paper is to prove the following theorem: \begin{thmi}\label{thmi:irreducible} Let $F$ be a finite-rank free group and let $\Phi:F\rightarrow F$ be an irreducible automorphism, and suppose that $G=F\rtimes_{\Phi}\ensuremath{\field{Z}}$ is word-hyperbolic. Then $G$ acts freely and cocompactly on a CAT(0) cube complex. \end{thmi} This result is a special case of Corollary~\ref{cor:irreducible_case}, which handles the more general case of a hyperbolic ascending HNN extension of a free group by an irreducible endomorphism. Theorem~\ref{thmi:irreducible} provides a widely-studied class of hyperbolic groups for which Gromov's question (see~\cite{Gromov87}) of whether hyperbolic groups are CAT(0) has a positive answer, but goes further, since nonpositively-curved cube complexes enjoy numerous useful properties beyond having universal covers that admit a CAT(0) metric. For example, combining Theorem~\ref{thmi:irreducible} with a result of~\cite{AgolVirtualHaken} shows that groups $G$ of the type described in Theorem~\ref{thmi:irreducible} are virtually special in the sense of~\cite{HaglundWiseSpecial} and therefore virtually embed in a right-angled Artin group. This implies that $G$ has several nice structural features, including $\ensuremath{\field{Z}}$-linearity. A group $G\cong F\rtimes_{\Phi}\ensuremath{\field{Z}}$ is word-hyperbolic exactly when $\Phi$ is atoroidal~\cite{BestvinaFeighnCombination,Brinkmann}, so that Theorem~\ref{thmi:irreducible} applies to all mapping tori of irreducible, atoroidal automorphisms of free groups. More generally, ascending HNN extensions are hyperbolic precisely if they have no Baumslag-Solitar subgroups~\cite{Kapovich:2000}. We actually prove the following more general statement: \begin{thmi}\label{thmi:general} Let $\phi:V\rightarrow V$ be a train track map of a finite graph $V$. Suppose that $\phi$ is $\pi_1$-injective and that each edge of $V$ is expanding. Moreover, suppose that the transition matrix $\mathfrak M$ of $\phi$ is irreducible and that the mapping torus $X$ of $\phi$ has word-hyperbolic fundamental group $G$. Then $G$ acts freely and cocompactly on a CAT(0) cube complex. \end{thmi} Our CAT(0) cube complex arises by applying Sageev's construction~\cite{Sageev:cubes_95} to a family of walls in the universal cover $\widetilde X$ of $X$. To ensure that the resulting action of $G$ on the dual cube complex is proper and cocompact, we show that there is a quasiconvex wall separating any two points in $\partial G$, thus verifying the cubulation criterion in~\cite{BergeronWise}. As train track maps are central to the proof that there are many walls in this sense, our results build upon the work of Bestvina, Feighn, and Handel in~\cite{BestvinaHandel,BestvinaFeighnHandel_lamination}. It appears likely that in the case where $\phi$ is $\pi_1$-surjective, the hypothesis that $\phi$ is irreducible can be removed, and we are currently working on developing the methodology in this paper to generalize Theorem~\ref{thmi:irreducible} to all hyperbolic mapping tori of free group automorphisms\footnote{We posted a tortuous generalization eight months after submitting this paper; see~\cite{HagenWise:general}.}. Moreover, for the construction of immersed walls in $X$, hyperbolicity of $G$ plays a minor role. It is therefore natural to wonder which free-by-cyclic groups admit free actions on CAT(0) cube complexes arising from immersed walls constructed essentially as in Section~\ref{sec:immersed_walls}. If $\Phi$ is fully irreducible and $G$ is not hyperbolic, then $\Phi$ is represented by a homeomorphism of a surface, by~\cite[Thm.~4.1]{BestvinaHandel}. Consequently, in this case $G$ acts freely on a locally finite, finite-dimensional CAT(0) cube complex~\cite{PrzytyckiWise:mixed}. It is reasonable to conjecture that in general, if $G=F\rtimes_{\Phi}\ensuremath{\field{Z}}$ is hyperbolic relative to virtually abelian subgroups, then $G$ acts freely on a locally finite, finite-dimensional CAT(0) cube complex. The techniques in this paper are largely portable to that context. However, one cannot expect to obtain cocompact cubulations for general free-by-cyclic groups. Indeed, Gersten's group $\langle a,b,c,t\,\mid\,a^t=a,\,b^t=ba,\,c^t=ca^2\rangle$ is free-by-cyclic but does not act metrically properly by semisimple isometries on a CAT(0) space~\cite{Gersten:automorphism}, and hence Gersten's group cannot act freely on a locally finite, finite-dimensional CAT(0) cube complex. Nevertheless, Gersten's group does act freely on an infinite-dimensional CAT(0) cube complex~\cite{WiseTubular}, so there is still much work to do in this direction. \subsection*{Summary of the paper} In Sections~\ref{sec:mapping_tori} and~\ref{sec:leaf_level_ladder}, we describe the mapping torus $X$ and introduce some features -- \emph{levels} and \emph{forward ladders} -- that play a role in the construction of immersed walls in $X$. In Section~\ref{sec:immersed_walls}, we describe \emph{immersed walls} $W\rightarrow X$ when $\phi:V\rightarrow V$ is an arbitrary $\pi_1$-injective map sending vertices to vertices and edges to combinatorial paths, under the additional assumptions that no power of $\phi$ maps an edge to itself and $\pi_1X$ is hyperbolic. The immersed wall $W$ is homeomorphic to a graph and has two parts, the \emph{nucleus} and the \emph{tunnels}, and is determined by a positive integer $L$ and a collection of sufficiently small intervals $d_i\subset V$, each contained in the interior of an edge. The nucleus is obtained by removing from $V$ each \emph{primary bust} $d_i$, along with its $\phi^L$-preimage. The tunnels are ``horizontal'' immersed trees joining endpoints of $d_i$ to endpoints of its preimage. Let $\widetilde W\rightarrow\widetilde X$ be a lift of the universal cover of $W$ and let $\overline W\subset\widetilde X$ be its image. Since $W\rightarrow X$ is not in general $\pi_1$-injective, $\widetilde W\rightarrow\overline W$ is not in general an isomorphism. However, under suitable conditions described in Section~\ref{sec:quasiconvexity}, $\overline W$ is a wall in $\widetilde X$ whose stabilizer is a quasiconvex free subgroup of $G$. The immersed walls in $X$ are analogous to the ``cross-cut surfaces'' introduced in~\cite{CooperLongReid}, and Dufour used these to cubulate hyperbolic mapping tori of self-homeomorphisms of surfaces~\cite{DufourThesis}. Section~\ref{sec:cutting_geodesics} and~\ref{sec:getting_lol} are devoted to the proof of Theorem~\ref{thmi:general}. We use a continuous surjection $\widetilde X\rightarrow\mathcal Y$ to an $\ensuremath{\field{R}}$-tree that arises in the case where $\phi$ is a train track representative of an irreducible automorphism (see~\cite{BestvinaFeighnHandel_lamination}). \subsection*{Acknowledgements}\label{subsec:acknowledgements} We thank the referees for their many helpful comments and corrections that improved this text. This is based upon work supported by the National Science Foundation under Grant Number NSF 1045119 and by NSERC. \section{Mapping tori}\label{sec:mapping_tori} Let $V$ be a finite connected graph based at a vertex $v$, and let $\phi:V\rightarrow V$ be a continuous, basepoint-preserving map such that $\phi(w)$ is a vertex for each vertex $w$ of $V$, and such that $\phi(e)$ is a \emph{combinatorial path} in $V$ for each edge $e$ of $V$. This means that there is a subdivision of $e$ such that vertices of the subdivision map to vertices and whose open edges map homeomorphically to open edges. We also assume that $\phi$ is parametrized so that these homeomorphisms are linear. Moreover, we require that the map $\Phi:F\rightarrow F$ induced by $\phi$ is injective, where $F\cong\pi_1V$ is a finite-rank free group. We note that any injective $\Phi:F\rightarrow F$ is represented by such a map $\phi$. The reader should have in mind the case where $\Phi$ is an irreducible automorphism of $F$ and $\phi$ is a train track map representing $\Phi$, in the sense of~\cite{BestvinaHandel}: \begin{defn}[Train track map]\label{defn:train_track_map} $\phi:V\rightarrow V$ is a \emph{train track map} if for all edges $e$ of $V$ and all $n\geq0$, the path $\phi^n(e)\rightarrow X$ is immersed. \end{defn} For an integer $L\geq 1$, let $X_L$ be obtained from $V\times[0,L]$ by identifying $(x,L)$ with $(\phi^L(x),0)$ for each $x\in V$, so that $X_L$ is the mapping torus of $\phi^L$, and let $X=X_1$. See Figure~\ref{fig:pic_of_x}. Let $G=\pi_1X$ and let $G_L=\pi_1X_L$ for each $L\geq 1$. Note that if $\Phi$ is surjective then $G_L\cong F\rtimes_{\Phi^L}\ensuremath{\field{Z}}$. \begin{figure}[ht] \includegraphics[width=0.5\textwidth]{pic_of_x.pdf}\\ \caption{The mapping torus $X_L$.}\label{fig:pic_of_x} \end{figure} We regard $V$ as a subspace of $X_L$, and we denote by $E$ the image in $X_L$ of $V\times\{\frac{1}{2}\}$; the space $E$ plays a role in Section~\ref{sec:immersed_walls}. We now describe a cell structure on $X_L$. Let $V\times[0,L]$ have the product cell structure: its vertices are $V^0\times\{0,L\}$, its \emph{vertical edges} are the edges of $V\times\{0,L\}$, and its \emph{horizontal edges} are of the form $\{w\}\times[0,L]$, where $w\in V^0$. We direct each horizontal edge $\{w\}\times[0,L]$ from $\{w\}\times\{0\}$ to $\{w\}\times\{L\}$, and horizontal edges of $X_L$ are directed accordingly. The $2$-cells of $X_L$ are images of the $2$-cells of $V\times[0,L]$, which have the form $e\times[0,L]$, where $e$ is an edge of $V$. For each vertex $w\in V^0\subset X_L^0$, we let $t_w$ denote the unique horizontal edge outgoing from $w$. When $L=1$, let $z\in G$ be the element represented by the loop $t_v$, where $v$ is the $\phi$-invariant basepoint of $V$. Note that conjugation by $z$ induces the monomorphism $\Phi:F\rightarrow F$. For each vertical edge $e$, joining vertices $a,b$, there is a $2$-cell $R_e$ with attaching map $t_b^{-1}e^{-1}t_a\phi^L(e)$, where $t_a,t_b$ are horizontal edges and $\phi^L(e)$ is a combinatorial path in $V$. Define a map $\varrho_L:X_L\rightarrow X$ as follows. First, $\varrho_L$ restricts to the identity on $V$. Each horizontal edge $t_w$ of $X_L$, joining $w$ to $\phi^L(w)\in V^0$, maps to the concatenation of $L$ horizontal edges of $X$ beginning at $w$. This determines $\varrho_L:X_L^1\rightarrow X$. This map extends to the 2-skeleton by mapping each 2-cell $R_e$ of $X_L$ to a disc diagram $D_e\rightarrow X$. Specifically, for $0\leq i\leq L$, the $i^{th}$ component $P_i$ of the vertical 1-skeleton of $D_e$ is the path $\phi^i(e)$, and there are strips of 2-cells between consecutive vertical components. The boundary path of $D_e$ consists of $P_0,P_L$, and the horizontal edges of $X$ joining the initial [terminal] point of $P_i$ to the initial [terminal] point of $P_{i+1}$ for $0\leq i\leq L-1$. Such a diagram $D_e$ is a \emph{long 2-cell} and is depicted in Figure~\ref{fig:long_2_cell}. Note that $D_e\rightarrow X$ is an immersion when $\phi$ is a train track map. Otherwise, the paths $P_i\rightarrow X$ are not necessarily immersed and the map $D_e\rightarrow X$ need not be locally injective. The map $G_L\rightarrow G$ induced by $\varrho_L$ embeds $G_L$ as an index-$L$ subgroup of $G$. \begin{figure}[ht] \includegraphics[width=0.4\textwidth]{long_2_cell.pdf} \caption{A 2-cell of $X_L$ is shown at left, its image in $X_L$ is at the center, and its image in $X$, which is the image of a long 2-cell in $X$, is shown at right. Here $L=3$. Horizontal edges have arrows and all non-arrowed edges are vertical.}\label{fig:long_2_cell} \end{figure} The universal cover $\widetilde X_L\rightarrow X_L$ inherits a cell structure from $X_L$. Let $\tilde v\in\widetilde X_L^0$ be a lift of the basepoint $v\in X_L^0$, and let $\widetilde V_0$ denote the smallest $F$-invariant subgraph containing the $\tilde v$ component of the preimage of $V$. Let $\widetilde V_{nL}=z^{nL}\widetilde V_0$ for $n\in\ensuremath{\field{Z}}$. There is a \emph{forward flow} map $\tilde\phi_L:\widetilde X_L\rightarrow\widetilde X_L$ defined as follows. For each $p\in V\times\{0\}$, let $S_p$ be the path $\{p\}\times[0,L]\rightarrow X_L$. The \emph{horizontal ray} $m_p\rightarrow X_L$ at $p$ is the concatenation $S_pS_{\phi^L(p)}S_{\phi^{2L}(p)}\cdots$. For $\tilde p\in\widetilde V_{nL}$ mapping to $p$, let $\tilde m_{\tilde p}$ be the lift of $m_p$ at $\tilde p$. For any $\tilde a\in\tilde m_{\tilde p}$, the point $\tilde\phi_L(\tilde a)$ is defined by translating $\tilde a$ a positive distance $L$ along $\tilde m_{\tilde p}$. When $L=1$, we denote $\tilde\phi_L$ by $\tilde\phi$. Let $\mathbf R_L$ denote the \emph{combinatorial line} with a vertex for each $nL\in L\ensuremath{\field{Z}}$ and an edge for each $[nL,nL+L]$ and let $\mathbf S_L$ be a circle with a single vertex and a single edge of length $L$. We define a map $q_L:\widetilde X_L\rightarrow\mathbf R_L$ as follows. There is a map $\bar q_L:X_L\rightarrow \mathbf S_L$ induced by the projection $V\times[0,L]\rightarrow[0,L]$. The map $\bar q_L$ lifts to the desired map $q_L$. Note that $q_L$ sends vertical edges to vertices and horizontal edges and 2-cells to edges of $\mathbf R_L$. We let $\mathbf R=\mathbf R_1$, $\mathbf S=\mathbf S_1$, and $q=q_1$. Let $\widetilde E_{nL}=q_L^{-1}(nL+\frac{1}{2})$. Each horizontal edge $t_w\cong\{w\}\times[0,L]\subset\widetilde X_L$ intersects $\widetilde E_{nL}$ at the point $\{w\}\times\{\frac{1}{2}\}$ for a unique $n\in\ensuremath{\field{Z}}$. \subsection{Metrics and subdivisions}\label{subsec:metrics_and_subdivisions} For each edge $e$ of $X$, let $|e|$ be a positive real number, with $|t_w|=1$ for each horizontal edge $t_w$. The assignment $e\mapsto |e|$ is a \emph{weighting} of $X^1$, and pulls back to a $G$-equivariant weighting of $\widetilde X^1$, with all horizontal edges having unit weight. Regarding $e$ as a copy of $[0,1]$, the subinterval $d\cong[a,b]\subset e$ has weight $|d|=(b-a)|e|$. Consider an embedded path $P\rightarrow\widetilde X^1$ (not necessarily combinatorial). The \emph{length} $|P|$ of $P$ is the sum of the weights of $P\cap e$, where $e$ varies over all edges. This yields a geodesic metric $\textup{\textsf{d}}$ on $\widetilde X^1$ such that $(\widetilde X^1,\textup{\textsf{d}})$ is quasi-isometric to $\widetilde X^1$ with the usual combinatorial path-metric in which edges have unit length. For each $L\geq 1$, let ${\widetilde X^\bullet}_L$ be the subdivision of $\widetilde X_L$ such that the lift $\tilde\varrho_L:\widetilde X_L\rightarrow\widetilde X$ of $\varrho_L$ sends open cells homeomorphically to open cells. The resulting map ${\widetilde X^\bullet}_L\rightarrow\widetilde X$ is an isomorphism on subspaces $\widetilde V_{nL}$ and sends 2-cells to long 2-cells. Note that 2-cells of $\widetilde X_L$ do not immerse in $\widetilde X$ unless $\phi$ is a train track map. Pulling back weights of edges in $\widetilde X$ to ${\widetilde X^\bullet}_L$ yields a metric $\textup{\textsf{d}}_L$ on $({\widetilde X^\bullet}_L)^1$ with respect to which $({\widetilde X^\bullet}_L)^1\rightarrow\widetilde X^1$ is a distance-nonincreasing quasi-isometry. We shall work mainly in $\widetilde X$, except in Section~\ref{sec:cutting_geodesics}, where it is essential to consider ${\widetilde X^\bullet}_L$. Beginning in Section~\ref{sec:immersed_walls}, we shall assume that $G$ is word-hyperbolic, so that there exists $\delta\geq 0$ such that $(\widetilde X^1,\textup{\textsf{d}})$ is $\delta$-hyperbolic. \section{Forward ladders and levels}\label{sec:leaf_level_ladder} In this section, we define various subspaces of $\widetilde X$ needed in the construction and analysis of quasiconvex walls in $\widetilde X$ and ${\widetilde X^\bullet}_L$. \begin{defn}[Midsegment]\label{defn:midsegment} Let $R_e\rightarrow\widetilde X$ be a 2-cell with boundary path $t_b^{-1}e^{-1}t_a\tilde\phi(e)$, where $e$ is a vertical edge joining vertices $a,b$. Regarding $R_e$ as a Euclidean trapezoid with parallel sides of length $|e|$ and $|\tilde\phi(e)|$, the \emph{midsegment in $R_e$} determined by $x\in e$ is the line segment joining $x$ to $\tilde\phi(x)$. The \emph{midsegment} in $\widetilde X$ determined by $x$ is the image of the midsegment in $R_e$ determined by $x$ under the map $R_e\rightarrow\widetilde X$, and is denoted $m_x$. Midsegments are directed so that $x$ is initial and $\tilde\phi(x)$ is terminal. The midsegment $m_x$ is \emph{singular} if $\tilde\phi(x)\in\widetilde X^0$ and \emph{regular} otherwise. In general, $R_e\rightarrow\widetilde X$ is not an embedding, and there may be distinct $x,y\in e$ with the property that the terminal points of $m_x$ and $m_y$ coincide. Note, however, that the intersection of two midsegments contains at most one point. See Figure~\ref{fig:midsegment}. \begin{figure}[ht] \includegraphics[width=0.3\textwidth]{midsegment}\\ \caption{Some midsegments in the image of a 2-cell in $\widetilde X$.}\label{fig:midsegment} \end{figure} \end{defn} \begin{defn}[Forward path, forward ladder]\label{defn:forward_ladder} Let $x\in\widetilde V_n$ for some $n\in\ensuremath{\field{Z}}$ and let $M\in\ensuremath{\field{Z}}$. The \emph{forward path} $\sigma_M(x)$ of length $M$ determined by $x$ is the embedded path that is the concatenation of midsegments starting at $x$ and ending at $\tilde\phi^M(x)$. In other words, $\sigma_M(x)$ is isomorphic to the combinatorial interval $[0,M]$, whose vertices are the points $\tilde\phi^i(x),\,0\leq i\leq M$ and whose edges are the midsegments joining $\tilde\phi^i(x)$ to $\tilde\phi^{i+1}(x)$. Any path $\sigma$ of this form is a \emph{forward path}. Note that $\sigma$ is a directed path with respect to the directions of midsegments in the sense that each internal point in which $\sigma$ intersects the vertical 1-skeleton of $\widetilde X$ has exactly one incoming and one outgoing midsegment. The \emph{forward ladder} $N(\sigma)$ associated to $\sigma$ is the smallest subcomplex of $\widetilde X$ containing $\sigma$. The $1$-skeleton $N(\sigma)^1$ plays an important role in many arguments. See Figure~\ref{fig:ladder}. \end{defn} \begin{figure}[ht] \includegraphics[width=0.3\textwidth]{ladder} \caption{A forward ladder. The forward path is labelled with arrows.}\label{fig:ladder} \end{figure} A subgraph $Y$ of $\widetilde X^1$ is \emph{$\lambda$-quasiconvex} if every geodesic of $\widetilde X^1$ starting and ending on $Y$ lies in $\mathcal N_\lambda(Y)$. We use the notation $\mathcal N_r(Y)$ to denote the closed $r$-neighborhood of $Y$. \begin{prop}[Quasiconvexity of forward ladders]\label{prop:forward_ladder_quasiconvex} There exist constants $\lambda_1\geq 1,\lambda_2\geq 0$ such that for each forward path $\sigma\rightarrow\widetilde X$, the inclusion $N(\sigma)^1\hookrightarrow\widetilde X^1$ is a $(\lambda_1,\lambda_2)$-quasi-isometric embedding. Hence, if $\widetilde X^1$ is $\delta$-hyperbolic, there exists $\lambda\geq 0$ such that each $N(\sigma)^1$ is $\lambda$-quasiconvex. \end{prop} \begin{proof} Let $\sigma$ join $x$ to $\tilde\phi^M(x)$, so that $\sigma=m_xm_{\tilde\phi(x)}\cdots m_{\tilde\phi^{M-1}(x)}$. Let $R_i$ be the 2-cell containing $m_{\tilde\phi^i(x)}$. Then a geodesic $P$ of $N(\sigma)^1$ joining $x$ to $\tilde\phi^M(x)$ has the form $P=Q_0t_0Q_1t_1Q_2\cdots t_{M-1}Q_M$, where each $t_i$ is a horizontal edge in $R_i$ and each $|Q_i|\leq\max_{e}\{|\phi(e)|\}$. Since each $m_i$ is a midsegment, $R_i\neq R_j$ for $i\neq j$, whence $q(P)$ is a combinatorial interval of length $M$, and the preimage in $N(\sigma)^1$ of each point in $q(P)$ is uniformly bounded. Hence $P$ is a uniform quasigeodesic in $\widetilde X^1$. \end{proof} In the case that $\widetilde X^1$ is $\delta$-hyperbolic, we denote by $\lambda$ the resulting quasiconvexity constant of the 1-skeleton of a forward ladder. \begin{defn}[Level]\label{defn:level} Let $x\in\widetilde V_n$ and let $L\geq 0$. Note that the preimage $(\tilde\phi_L)^{-1}(x)$ is a finite set $\{x_i\}$ in $\widetilde V_{n-L}$. Let $\sigma_L(x_i)$ be the forward path beginning at $x_i$ and ending at $x$. The \emph{level} $T^o_L(x)$ is the subspace $\cup_i\sigma_L(x_i)$. The point $x$ is the \emph{root} of $T_L^o(x)$ and $L$ is the \emph{length}. The \emph{carrier} $N(T^o_L(x))$ is the smallest subcomplex of $\widetilde X$ containing $T^o_L(x)$. Note that $N(T^o_L(x))=\cup_iN(\sigma_i)$, where $\sigma_i$ varies over the finitely many maximal forward paths in $T^o_L(x)$. Note that each level has a natural directed graph structure in which edges are midsegments. \end{defn} \begin{prop}[Properties of levels]\label{prop:properties_of_levels} Let $T^o_L(x)$ be a level. Then: \begin{enumerate} \item $T^o_L(x)$ is a directed tree in which each vertex has at most one outgoing edge. \item If $x\not\in\widetilde X^0$, then there exists a topological embedding $T^o_L(x)\times[-1,1]\rightarrow\widetilde X$ such that $T^o_L(x)\times\{0\}$ maps isomorphically to $T^o_L(x)$. \item If $L'\geq L$, then $T^o_L(x)\subseteq T^o_{L'}(x)$. \end{enumerate} \end{prop} \begin{proof} $T^o_L(x)$ is connected since it is the union of a collection of paths, each of which terminates at $x$. Each vertex of $T^o_L(x)$ has at most one outgoing edge. Hence any cycle in $T^o_L(x)$ is directed. The map $q:T^o_L(x)\rightarrow\mathbf R$ thus shows that there are no cycles in $T^o_L(x)$. This establishes assertion~$(1)$. Let $x_i\in(\tilde\phi_L)^{-1}(x)$ and let $\sigma_i\subset T^o_L(x)$ be the forward path joining $x_i$ to $x$. Then $\sigma_i$ is disjoint from $\widetilde X^0$, since $T^o_L(x)$ is regular. Hence there exists $\epsilon_i>0$ such that $N(\sigma_i)$ contains an embedded copy of $\sigma_i\times[-\epsilon_i,\epsilon_i]$ with $\sigma_i\times\{0\}=\sigma_i$, which we denote by $F_i$. Let $\epsilon=\min_i\epsilon_i$. For each $i$, let $F'_i\subset F_i$ be $\sigma_i\times[-\epsilon,\epsilon]\subseteq\sigma_i\times[-\epsilon_i,\epsilon_i]$, and let $F=\cup_iF'_i$. Since $\sigma_i\cap\sigma_j$ is a forward path for all $i,j$, the subspace $F\cong T^o_L(x)\times[-\epsilon,\epsilon]$. See Figure~\ref{fig:product_neighborhoods}. \begin{figure}[ht] \includegraphics[width=0.6\textwidth]{product_neighborhoods.pdf} \caption{The product neighborhood of a regular level in $\widetilde X$ is shown at right; the corresponding level in $\widetilde X_L$ appears at left. In general, the product neighborhood may contain several subintervals of each vertical edge since $\phi$ is not in general an immersion on edges.}\label{fig:product_neighborhoods} \end{figure} Assertion~$(3)$ follows from the fact that $\tilde\phi^{L'}:\widetilde X\rightarrow\widetilde X$ factors as $\widetilde X\stackrel{\tilde\phi^{L}}{\longrightarrow}\widetilde X\stackrel{\tilde\phi^{L'-L}}{\longrightarrow}\widetilde X$. \end{proof} For each $L\geq 1$, forward paths and levels are defined in precisely the same way in $\widetilde X_L$. A level of $\widetilde X_L$ is subdivided when we formed ${\widetilde X^\bullet}_L$ in Section~\ref{subsec:metrics_and_subdivisions}. Accordingly, each length-$L$ level in ${\widetilde X^\bullet}_L$ is isomorphic to a star whose edges are subdivided into length-$L$ paths. The map $\tilde\varrho_L:{\widetilde X^\bullet}_L\rightarrow\widetilde X$ sends each length-$n$ level of $\widetilde X_L$, each of whose maximal forward paths contains $nL$ midsegments of ${\widetilde X^\bullet}_L$, to a length-$nL$ level in $\widetilde X$. Thus $\varrho_L$ maps subdivided stars to rooted trees, as shown in Figure~\ref{fig:product_neighborhoods}. The image in $X_L$ of a level from ${\widetilde X^\bullet}_L$ is also referred to as a level; it will be clear from the context whether we are working in the base space or the universal cover. The following observation about forward ladders is required in several places in Section~\ref{sec:quasiconvexity} and Section~\ref{sec:cutting_geodesics}. \begin{lem}\label{lem:vert_horiz_bound} Let $\sigma$ be a forward path. Then for each $R\geq 0$, there exists $\Theta_R\geq 0$, independent of $\sigma$ and $n$, such that $\diam(\mathcal N_R(N(\sigma)^1)\cap\mathcal N_R(\widetilde V_n))\leq\Theta_R$ for all $n\in\ensuremath{\field{Z}}$. \end{lem} \begin{proof} This follows from the fact that $q(\widetilde V_n)=n$, while the image of $q|_{N(\sigma)}$ is an interval, each of whose points has uniformly bounded preimage in $N(\sigma)^1$. \end{proof} \section{Immersed walls, walls, and approximations}\label{sec:immersed_walls} In this section, we will describe immersed walls $W\rightarrow X$, which are determined by two parameters. The first parameter is a collection $\{d_i\}$ of subintervals of edges in $V$, called \emph{primary busts}. The second parameter is an integer $L\geq 1$ called the \emph{tunnel length}. The graph $W$ consists of $V-\cup_id_i-\cup_i(\phi^L)^{-1}(d_i)$ together with a collection of rooted trees called \emph{tunnels}, and is immersed in $X_L$. We shall show that when $L$ is sufficiently large, $W\rightarrow X$ corresponds to a quasiconvex codimension-1 subgroup of $G$. \subsection{Primary busts}\label{subsec:bust_choose} Let $\{e_1,\ldots,e_k\}$ be edges of $V\subset X$. For each $i$, let $e'_i$ be the image of $e_i$ under the isomorphism $V\rightarrow E$ given by $(x,0)\mapsto(x,\frac{1}{2})$. The subspaces $e'_i$, regarded as edges of $E$, are \emph{primary busted edges}. We will choose closed nontrivial intervals $d'_i\subset\interior{e'_i}$, whose distinct endpoints we denote by $p_i^{\pm}$. The corresponding subinterval of $e_i$ is denoted $d_i$, and its endpoints $q_i^{\pm}$ correspond to $p_i^{\pm}$. Let $E^{\flat}=E-\cup_{i=1}^k\interior{d'_i}$ and let $V^{\flat}$ denote its preimage in $V$ under the above isomorphism $V\rightarrow E$. The subspace $E^{\flat}$ is the \emph{primary busted space}, and each $d_i$ (or $d'_i$) is a \emph{primary bust}; $\interior{d_i}$ (or $\interior{d'_i}$) is an \emph{open primary bust}. Let $C$ be a component of $V^{\flat}$ and let $\widetilde C$ be a lift of its universal cover to some $\widetilde V_n$. Since $C\hookrightarrow V\hookrightarrow X$ is $\pi_1$-injective, $\widetilde C$ embeds in $\widetilde V_n$. Its parallel copy $\widetilde C'\subset\widetilde E_n$ is a \emph{primary nucleus}, and likewise, each component of $E^{\flat}$ is a \emph{primary nucleus} in $X$. \begin{rem}[Quasiconvexity of $\widetilde C$ under various conditions]\label{rem:nucleus_subgroup_conditions} In our applications, we will require $\widetilde C$ to be quasiconvex in $\widetilde X^1$. This is achievable in several ways. Clearly, if $\{e_i\}$ contains enough edges that $E^{\flat}$ is a forest, then the subspaces $\widetilde C\subset\widetilde X^1$ are finite trees and therefore quasiconvex. Quasiconvexity of $\widetilde C$ occurs under other circumstances. For example, suppose that $\Phi:F\rightarrow F$ is an automorphism and $\phi$ is a train track map that is \emph{aperiodic} in the sense of~\cite{MitraQuasiconvex}, i.e. $\phi^n(e)$ traverses $f$ for all edges $e,f$ and all sufficiently large $n$. Then, provided $\{e_i\}$ contains at least one edge corresponding to a nontrivial splitting of $F$, the following theorem of Mitra (see~\cite[Prop.~3.4]{MitraQuasiconvex}), which is an analog of a result of Scott and Swarup~\cite{Scott_Swarup}, ensures that each $\widetilde C$ is quasiconvex \begin{thm}\label{thm:mitra} Let $\Phi:F\rightarrow F$ be an aperiodic automorphism of the finite-rank free group $F$. If $H\leq F$ is a finitely generated, infinite-index subgroup, then $H$ is quasiconvex in $F\rtimes_{\Phi}\ensuremath{\field{Z}}$. \end{thm} \end{rem} \subsection{Constructing immersed walls}\label{subsec:immersed_wall_construction} We now assume that $\widetilde X^1$ is $\delta$-hyperbolic. Let $L\geq 1$ be an integer, called the \emph{tunnel length}. For any set $\{d_i\}$ of nontrivial primary busts, the spaces $E^{\flat}$ and $V^{\flat}$ embed in $X_L$ by maps factoring through $E\hookrightarrow X_L$ and $V\hookrightarrow X_L$ respectively. For each $i$, let $\{d_{ij}\}_{j}$ denote the finite set of components of $(\phi^L)^{-1}(d_i)$. For each $i,j$, let $d'_{ij}$ be the parallel copy of $d_{ij}$ in $E$. Each $d_{ij}$ or $d_{ij}'$ is a \emph{secondary bust}. In order to choose busts, we will assume that each edge $e$ of $V$ is \emph{expanding} in the sense that $\phi^k(e)\neq e$ for all $k>0$. This assumption is justified by the following lemma (see also~\cite{BestvinaHandel}). \begin{defn} We say $x\in V$ is \emph{periodic} if $\phi^n(x)=x$ for some $n\geq 1$. A point $x\in V$ has \emph{period~$m$} if $\phi^m(x)=x$ and $\phi^k(x)\neq x$ for $0<k<m$. We then refer to $x$ as being \emph{$m$-periodic}. A forward path $\sigma\rightarrow\widetilde X$ is \emph{periodic} if it is a subpath of a bi-infinite forward path whose stabilizer in $G$ is nontrivial. Note that this holds exactly when each point of $\widetilde X^1\cap\sigma$ projects to a periodic point of $V$. \end{defn} Recall that the map $\phi:V\rightarrow V$ is \emph{irreducible} if for all edges $e,f$, there exists $n\geq0$ such that $\phi^n(e)$ traverses $f$ \begin{lem}\label{lem:edges_expanding} Let $F\rtimes_{\Phi}\ensuremath{\field{Z}}$ be hyperbolic. Then $\Phi:F\rightarrow F$ can be represented by a map $\phi:V\rightarrow V$ with respect to which each edge of $V$ is expanding and no edge is mapped to a point. Moreover, if $\Phi$ has an irreducible train track representative, then $\phi:V\rightarrow V$ can be chosen to be an irreducible train track map with respect to which each edge is expanding. \end{lem} \begin{proof} We begin with a representative $\phi:V\rightarrow V$, which we will adjust by contracting subtrees of $V$. Let $U\subset V$ be the union of all vertices and all closed edges $e$ such that $|\phi^k(e)|$ is bounded as $k\rightarrow\infty$. First, note that $\phi(U)\subseteq U$. Second, each component of $U$ is contractible, since otherwise either $\phi$ is not $\pi_1$-injective or $X$ would contain an immersed torus, contradicting hyperbolicity. We now collapse the $\phi$-invariant forest $U$ as in~\cite[Page~7]{BestvinaHandel}, resulting in a graph $\overline V$ and a map $\bar\phi:\overline V\rightarrow\overline V$ representing $\Phi$ (by reparametrizing, we can assume that the restriction of $\bar\phi$ to each edge is a combinatorial path). Note that either $U$ contained no edges (so all edges were expanding and did not map to points), or $\overline V$ has strictly fewer edges than $V$. We repeat the above procedure finitely many times to obtain a graph $\overline V$ and a map $\bar\phi:\overline V\rightarrow\overline V$ such that edges map to nontrivial paths and all edges are expanding. The collapse of $U$ preserves the property of being a train track map. Indeed, let $\bar e$ be an edge of $\overline V=V/U$ that is the image of an edge $e$ of $V$. Let $n>0$, and consider the restriction of $\bar\phi^n$ to $\bar e$. The path $\bar\phi^n(\bar e)$ is obtained from the immersed path $\phi^n(e)$ by collapsing each edge that maps to $U$. Let $\bar u,\bar v$ be consecutive edges of $\bar\phi^n(\bar e)$ that fold. Then there is a subpath $u^{-1}fv\subset\phi^n(e)$, where $u\mapsto\bar u,v\mapsto\bar v$ and $f$ is an immersed path in $U$. Observe that $f$ is a closed path since $u,v$ have the same initial point. This contradicts the fact that $U$ is a forest. Finally, the property of irreducibility is preserved by collapsing invariant forests. Indeed, let $\bar e,\bar f$ be edges of $\overline V$ that are images of edges $e,f$ of $V$. Then by irreducibility of $\phi$, there exists $m>0$ such that $\phi^m(e)$ passes through $f$, and hence $\bar\phi^m(\bar e)$ passes through $\bar f$. \end{proof} A point $y\in V$ is \emph{singular} if $\phi^k(y)\in V^0$ for some $k$. \begin{lem}\label{lem:choosing_busts_1} Let $L\in\ensuremath{\field{N}}$. Let $\{e_i\}_{i=1}^k$ be a set of expanding edges in $V$, let $x_i\in\interior{e_i}$ for each $i$, and let $\epsilon>0$. Then there exists a collection $\{d_i\}_{i=1}^k$ of closed subintervals, with each $d_i\subset\interior{e_i}$, such that: \begin{enumerate} \item \label{item:disjoint_from_preimage}$\cup_id_i$ is disjoint from $\cup_{ij}d_{ij}$. \item \label{item:neighborhood}$\cup_jd_{ij}$ lies in the $\epsilon$-neighborhood of $(\phi^L)^{-1}(x_i)$ for each $i$. \item \label{item:regular_endpoints}The endpoints $p_i^{\pm}$ of $d_i$ are nonsingular. \item \label{item:anywhere}If $x_i$ is nonsingular and $\phi^L(x_i)\neq x_i$ then we can choose $d_i$ such that $x_i\in\{p_i^{\pm}\}$ is an endpoint of $d_i$. \item \label{item:bustsembed}$\phi^L$ restricts to an embedding on $d_i$, for each $i$. \item \label{item:disjoint}Suppose that $\phi^L(x_i)\neq\phi^L(x_j)$ for all $i\neq j$. Then $\phi^L(d_i)\cap\phi^L(d_j)=\emptyset$. \end{enumerate} \end{lem} \begin{proof} We first establish the finiteness of the set $\mathcal S$ consisting of points $s\in e_i$ such that $\phi^L(s)=s$. Each component $b$ of $e_i\cap(\phi^L)^{-1}(e_i)$ is the concatenation of one or more subintervals of $e_i$, each of which maps homeomorphically to $e_i$. Since $e_i$ is expanding, Brouwer's fixed point theorem implies that each such subinterval contains a unique point $s$ with $\phi^L(s)=s$. As there are finitely many such $b$, we conclude that $\mathcal S$ is finite. Let $z_i \in \interior{e_i} -\mathcal S$. There exists a nonempty closed interval $h_i$ containing $z_i$ such that $h_i\cap (\phi^L)^{-1}(h_i) =\emptyset$. Indeed, if $h_i\cap (\phi^L)^{-1}(h_i) \neq \emptyset$ for each closed interval containing $z_i$ then there would be a sequence of points converging to $z_i$ whose $\phi^L$-images also converge to $z_i$, and so $z_i\in S$. Property~(1) holds whenever $d_i\subset h_i$. By continuity of $\phi^L$, there exists $\delta>0$ such that $\mathcal N_\delta((\phi^L)^{-1}(x_i)) \subset (\phi^L)^{-1}(\mathcal N_\epsilon(x_i))$. Property~(2) holds by choosing $z_i\in\mathcal N_\delta((\phi^L)^{-1}(x_i))$ and letting $d_i$ be a nontrivial component of $h_i\cap\closure{\mathcal N_\delta((\phi^L)^{-1}(x_i))}$. As there are countably many singular points, Property~(3) holds since we can assume that neither endpoint of $h_i$ is singular. Property~(4) holds by letting $z_i=x_i$, and then choosing $h_i$ above so that it has $x_i$ as an endpoint. To prove~(5), note that $e_i$ has a subdivision where the vertices are points of $(\phi^L)^{-1}(V^0)$. If $d_i$ is properly contained in a single closed edge in this subdivision, then $\phi^L$ restricts to an embedding on $d_i$. This can be arranged by choosing $d_i$ sufficiently small (fixing $x_i$). We prove~(6) by induction on $|\{e_i\}|$. The base case, where $k=0$, is vacuous. Suppose that $d_1,\ldots,d_{k-1}$ have been chosen to satisfy~(1)-(6), with each $d_i$ satisfying $x_k\not\in\phi^L(d_i)$ for $1\leq i<k$. Choose $d_k$ with properties~(1)-(5) small enough to avoid the finitely many $\phi^L(d_i),\,1\leq i<k$. \end{proof} Clearly $d_i\cap d_{i'}=\emptyset$ for $i\neq i'$, since $d_i,d_{i'}$ are contained in distinct open edges. Consequently $d_{ij}\cap d_{i'j'}=\emptyset$ unless $i=i'$ and $j=j'$. The subspace $\mathbf N$ of $E^{\flat}$ obtained by removing the image under $V\stackrel{\sim}{\rightarrow}E$ of each open secondary bust is the \emph{nucleus}. Observe that $\mathbf N$ need not be connected. For each $i,j$, let $q_{ij}^{\pm}$ be the endpoints of $d_{ij}$, which map to $q_i^{\pm}\in V$, and let $p^{\pm}_{ij}$ be the corresponding points of $d'_{ij}\subset E$. See Figure~\ref{fig:nucleus_attaching_points}. \begin{figure}[ht] \includegraphics[width=0.6\textwidth]{nucleus_attaching_points} \caption{Constructing a wall in $X_L$.}\label{fig:nucleus_attaching_points} \end{figure} For each $i$, let $\check T_{i}^{o\pm}$ be the image in $X_L$ of the level $T_L^o(\tilde q_i^{\pm})\subset\widetilde X_L$, where $\tilde q_i^{\pm}$ is an arbitrary lift of $q_i^{\pm}\in d_i$. Recall that $\check T_i^{o\pm}$ is an embedded star of length $L$ rooted at $q_i^{\pm}$ with leaves at the various $q_{ij}^{\pm}$. Let $S_i^{\pm}$ be a segment in the 2-cell $R_{e_i}$ of $X_L$ that joins $q_i^{\pm}$ to $p_i^{\mp}\in E$. (Note that $S_i^+$ joins $p_i^+$ to $q_i^-$ and $S_i^-$ joins $p_i^-$ to $q_i^+$.) The arcs $S_i^{\pm}$ are \emph{slopes}. The \emph{level-part} $T_i^{o\pm}$ is the rooted subtree of $\check T_i^{o\pm}$ with leaves at $p_{ij}^{\pm}$. The subspace $T_i^{\pm}=T_i^{o\pm}\cup S_i^{\pm}$ obtained by joining the level-part $T_i^{\pm}$ and the slope along the common point $q_i^{\pm}$ is a \emph{tunnel}. The space $\widehat W^{\bullet}$ determined by the primary busts $\{d_i\}$ and the tunnel length $L$ is the graph obtained by joining each tunnel $T_i^{\pm}$ to $\mathbf N$ along $\{p_{ij}^{\pm}\}\cup\{p_i^{\mp}\}$. The inclusion $\mathbf N\hookrightarrow X_L$ and the inclusions $T_i^{\pm}\hookrightarrow X_L$ induce a (non-combinatorial) immersion $\widehat W^\bullet\rightarrow X_L$. Note that $T^{\pm}_i\cap T^{\pm}_j=\emptyset$ when $i\neq j$ since $d_i\cap d_j=\emptyset$. Note that for each $i$, the tunnels $T_i^+$ and $T_i^-$ intersect in the single point $S_i^+\cap S_i^-$. Composing with the map $X_L\rightarrow X$ gives an immersion $\widehat W^{\bullet}\rightarrow X$. This extends to a local homeomorphism $\widehat W^{\bullet}\times[-1,1]\rightarrow X$ with $\widehat W^{\bullet}$ identified with $\widehat W^{\bullet}\times\{0\}$. Indeed, we described a map $T_i^{\pm}\times[-1,1]\rightarrow X$ earlier, and $\mathbf N\times[-1,1]\rightarrow X$ is an embedding since $\mathbf N\subset E$, and each $S^{\pm}_i$ lies in a 2-cell. Appropriately chosen neighborhoods $T_i^{\pm}\times[-1,1]$, and $S_i^{\pm}$, and $\mathbf N$ can be glued to form $\widehat W^{\bullet}\times[-1,1]\rightarrow X_L$. These gluings can be chosen to preserve a ``normal vector'' at each point of the tunnel, and hence the result is a trivial $[-1,1]$ bundle. The map $\widehat W^{\bullet}\rightarrow X$ factors through an immersion $W^{\bullet}\rightarrow X$, where $W^{\bullet}$ is obtained from $\widehat W^{\bullet}$ by folding the levels according to the map $\varrho_L:X_L\rightarrow X$ illustrated in Figure~\ref{fig:product_neighborhoods}. The spaces $\widehat W^\bullet$ and $W^\bullet$ are shown in Figure~\ref{fig:wall_cartoon}. \begin{figure}[ht] \includegraphics[width=0.35\textwidth]{wallcartoon} \caption{At left is $\widehat W^{\bullet}$; at right is $W^{\bullet}$. Light lines indicate the nucleus, while heavy lines are tunnels. The emphasized points in each picture are the interior points of a primary bust edge of $V$ where the slopes and levels meet.}\label{fig:wall_cartoon} \end{figure} \begin{defn}\label{defn:wall_connected} A component $W$ of $W^{\bullet}$ is an \emph{immersed wall}. \end{defn} \subsection{Description of $\overline W$}\label{subsec:wall_description} The map $W\rightarrow X$ lifts to a map $\widetilde W\rightarrow\widetilde X$ of universal covers. For each component $C$ of $\mathbf N$, the universal cover $\widetilde C$ of $C$ lifts to $\widetilde W$, and the restriction of $\widetilde W\rightarrow\widetilde X$ to each such $\widetilde C$ is an embedding. Moreover, each tunnel lifts to $\widetilde W$, and the map $\widetilde W\rightarrow\widetilde X$ restricts to an embedding on each tunnel $T_i\subset\widetilde W$. Let $\overline W=\image(\widetilde W\rightarrow\widetilde X)$ and let $H_W=\stabilizer_G(\overline W)$. We conclude that: \begin{rem}\label{rem:invariant_halfspace} When $\overline W$ is locally isomorphic to $W$, the trivial $[-1,1]$-bundle discussed above ensures that there are exactly two components of $\widetilde X-\overline W$, each of which is $H_W$-invariant. \end{rem} \begin{rem}[Future shape of $\overline W$]\label{rem:wall_structure} We now describe the structure of $\overline W$ in the situation in which distinct tunnels are disjoint. Note that tunnels $T_i^{\pm}$ and $T_j^{\pm}$ in $\overline W$ are disjoint when $i\neq j$, since they map to disjoint tunnels in $\image(W\rightarrow X)$. Moreover, we shall show below that, under certain conditions, tunnels $T,T'\subset\overline W$, mapping to $T_i^+,T_i^-$ respectively, are disjoint when $L$ is large. In this situation, $\overline W$ will be shown to have the structure of a tree of spaces, whose underlying vertices are equipped with a 2-coloring. Red vertices correspond to slopes, while green vertices correspond to subspaces that are maximal connected unions of universal covers of nuclei and lifts of level-parts. Note that $\overline W$ may still fail to be simply connected -- i.e. $\widetilde W$ may still fail to embed -- since subspaces corresponding to green vertices may not be simply-connected. If $\overline W$ contains a nucleus in $\widetilde E_n$, then all nuclei lie in $\widetilde E_{n+kL},\,k\in\ensuremath{\field{Z}}$, and any two nuclei contained in a common vertex space lie in the same space $\widetilde E_{n+kL}$. A heuristic picture of $\overline W$ is shown in Figure~\ref{fig:heuristic_wall}, and Figure~\ref{fig:small_wall} shows a part of $\overline W$ inside $\widetilde X$. \end{rem} \begin{figure}[ht] \begin{overpic}[width=0.35\textwidth]{new_fig_8.pdf} \put(25,10){\scriptsize{level}} \put(75,30){\scriptsize{level}} \put(55,23){\scriptsize{slope}} \put(-4,45){\scriptsize{nucleus}} \end{overpic} \caption{A heuristic picture of part of a wall in $\widetilde X$. The two nuclei at left, and the levels at left, belong to the same knockout. This knockout does not contain the slope or the nucleus and level at right.} \label{fig:heuristic_wall} \end{figure} \begin{com}add the replacement figure 8\end{com} \begin{figure}[ht] \includegraphics[width=0.4\textwidth]{small_wall} \caption{Part of a wall $\widetilde W\rightarrow\widetilde X$. The single-arrowed segments belong to a nucleus, while the double-arrowed segments are tunnels.}\label{fig:small_wall} \end{figure} \subsection{The approximation}\label{subsec:approximation} Let $N(\overline W)$ denote the union of all closed 2-cells of $\widetilde X$ that intersect $\overline W$. We will show that $N(\overline W)^1$ is quasiconvex in $\widetilde X^1$ under certain conditions, notably sufficiently large tunnel length. However, the quasiconvexity constant will depend on the tunnel length. This is partly because levels are not uniformly quasiconvex and partly because distinct levels emanating from very close secondary busts may contain long forward paths that closely fellow-travel. To achieve uniform quasiconvexity we define the \emph{approximation} of $\overline W$, which also has the key feature that it lifts to a geometric wall in ${\widetilde X^\bullet}_L$. \newcommand{\comappn}[1]{N(\mathbf A(#1))^1} \begin{defn}[Approximation]\label{defn:approximation} Let $W\rightarrow X$ be an immersed wall with tunnel length $L$ and primary busted edges $\{e_i\}$. Let $\overline W$ be the image of a lift $\widetilde W\rightarrow\widetilde X$ of the universal cover of $W$ to $\widetilde X$. We define a map $\comapp:\overline W\rightarrow\widetilde X$ as follows. First, suppose that $\widetilde C\subset\overline W$ is the universal cover of a component of the nucleus of $W$. Let $n\in\ensuremath{\field{Z}}$ be such that $\widetilde C\subset\widetilde E_n$, and let $\widetilde C'\subset\widetilde V_n$ be the parallel copy of $\widetilde C$. For each $c\in\widetilde C$, let $c'$ denote the corresponding point of $\widetilde C'$. Then $\comapp:\widetilde C\rightarrow\widetilde X$ is defined by $\comapp(c)=\tilde\phi^L(c')$. For each level-part $T^o$ of $\overline W$, let $q$ be the root of $T^o$. Then $\comapp(t)=q$ for each $t\in T_o$. Finally, let $S\subset\overline W$ be a slope, beginning at $q$ and ending on a point $p$ in a nucleus $\widetilde C$. Then $p$ is an endpoint of a primary bust $d_i\subset\widetilde E_n$. The map $\comapp$ sends the slope $S$ homeomorphically to the path $d_i'P$, where $d_i'$ is the parallel copy of $d_i$ in $\widetilde V_n$ that joins $q$ to $p'$ and $P$ is the forward path joining $p'$ to $\tilde\phi^L(p')$. See Figure~\ref{fig:comapp_def}. The \emph{approximation} of $\overline W$ is the subspace $\comapp(\overline W)\subset\widetilde X$. Note that $\comapp(\overline W)$ is the union of length-$L$ forward paths together with subspaces of $\widetilde V_{nL}$ for each $n\in\ensuremath{\field{Z}}$. Let $\comappn{\overline W}$ be the 1-skeleton of the smallest subcomplex of $\widetilde X$ containing $\comapp(\overline W)$. \end{defn} \begin{figure}[ht] \includegraphics[width=0.4\textwidth]{comapp_def}\\ \caption{Part of a wall and its approximation. The arrowed paths are the approximations of the slopes intersecting them.}\label{fig:comapp_def} \end{figure} \begin{rem} If $\widetilde C_1,\widetilde C_2$ are nuclei of $\overline W$ intersecting a level-part of a tunnel of $\overline W$, then $\comapp(\widetilde C_1)\cap\comapp(\widetilde C_2)\neq\emptyset$. For each $n\in\ensuremath{\field{Z}}$, each component of $\comapp(\overline W)\cap\widetilde V_n$ is formed as follows. A \emph{knockout} $\widetilde K$ is a maximal connected subspace of $\overline W$ that does not contain an interior point of a slope. The knockout $\widetilde K$ is \emph{at position $n$} if it is the union of nuclei in $\overline W\cap\widetilde E_{n-L}$ together with level-parts traveling from $\widetilde E_{n-L}$ to $\widetilde V_n$. To each position-$n$ knockout $\widetilde K$, we associate a component of $\comapp(\overline W)\cap\widetilde V_n$, namely the one obtained from the connected subspace $\comapp(\widetilde K)\subset\widetilde V_n$ by adding all (closed) primary bust intervals that intersect $\comapp(\widetilde K)$. See Figure~\ref{fig:pushforward}. \end{rem} \begin{figure}[ht] \includegraphics[width=0.3\textwidth]{pushforward}\\ \caption{Two nuclei intersecting a common level-part have intersecting approximations whose union avoids primary busts.}\label{fig:pushforward} \end{figure} \begin{rem}\label{rem:distinct_slopes_disjoint_approxes} If $S_1,S_2$ are distinct slopes, rooted at primary busts $d_1,d_2$, then $\comapp(S_1)\cap\comapp(S_2)=\emptyset$ by Lemma~\ref{lem:choosing_busts_1}.(5)-(6). \end{rem} \subsection{Hypotheses on $\comapp(\overline W)$ enabling quasiconvexity}\label{sec:lol} We recall that we are assuming, for the rest of the paper, that there exists $\delta\geq0$ such that $\widetilde X^1$ is $\delta$-hyperbolic. The following statements are instrumental in proving that, provided $L$ is sufficiently large, $\comappn{\overline W}\hookrightarrow\widetilde X$ is quasi-isometrically embedded, and the quasi-isometry constants are independent of $L$. \begin{defn}[Sub-quasiconvex]\label{defn:bust_quasiconvex} $W$ is \emph{sub-quasiconvex} if there exist constants $\mu_1',\mu_2'$ such that each component in $\widetilde X$ of the preimage of $V-\cup_i\interior{d_i}$ is $(\mu_1',\mu_2')$-quasi-isometrically embedded. For example, as noted above, $W$ is sub-quasiconvex if $V^{\flat}$ is a forest or if $\Phi$ is an aperiodic isomorphism and there is at least one bust. \end{defn} \begin{lem}[Quasiconvexity of approximations of nuclei]\label{lem:avoids_bust} Approximations have the following properties when $W$ is sub-quasiconvex: \begin{enumerate} \item For each nucleus $\widetilde C$ and each open primary bust $\tilde d$, we have $\comapp(\widetilde C)\cap\tilde d=\emptyset$. \item Let $\widetilde K$ be a knockout. Then $\comapp(\widetilde K)\cap\tilde d=\emptyset$ for each open primary bust $\tilde d$. \item \label{nucleus_approx_quasiconvex} Hence there exist $\mu_1\geq1,\mu_2\geq 0$ such that, for each $n\in\ensuremath{\field{Z}}$ and each component $\mathbf K$ of $\widetilde V_n\cap\comapp(\overline W)$, the inclusion $\mathbf K\rightarrow\widetilde X^1$ is a $(\mu_1,\mu_2)$-quasi-isometric embedding. Moreover, $\mu_1$ and $\mu_2$ are independent of $\{d_i\}$ and $L$. \end{enumerate} \end{lem} \begin{proof} \begin{enumerate} \item $\widetilde C$ has empty intersection with the set of open secondary busts in $\widetilde V_n$, and hence maps to the complement of the set of open primary busts in $\widetilde V_{n+L}$. \item This follows immediately from $(1)$ because level-parts map to points. \item Since $W$ is sub-quasiconvex, Statement~(2) implies that $\mathbf K$ is a subtree of a uniform neighborhood in $\widetilde V_n$ of some $\comapp(\widetilde K)$, and the claim follows. Since there are finitely many possible sets of primary busted edges, the constants $\mu_1,\mu_2$ can be chosen independently not only of $L$ and $\{d_i\}$, but also of $\{e_i\}$. Indeed, each set $\{e_i\}$ of edges yielding a sub-quasiconvex immersed wall gives rise to a pair of quasi-isometry constants, and we take $\mu_1,\mu_2$ to be the maximal such constants.\qedhere \end{enumerate} \end{proof} Lemma~\ref{lem:avoids_bust} provides uniform quasiconvexity of nucleus approximations, and Proposition~\ref{prop:forward_ladder_quasiconvex} provides uniform quasiconvexity of forward ladders. Lemma~\ref{lem:vert_horiz_bound} provides a bound on the diameters of coarse intersections of nucleus approximations and carriers of approximations of slopes. To prove quasiconvexity of $\comappn{\overline W}$ requires the following additional property. \begin{defn}[Ladder overlap property]\label{defn:ladder_overlap_property} The family of immersed walls $\{W_i\rightarrow X\}$ has the \emph{ladder overlap property} if there exists $B\geq0$ such that for all $i$ and all distinct tunnels $T_1,T_2\subset\overline W_i$ intersecting a common nucleus, $$\diam\left(\mathcal N_{3\delta+2\lambda}(N(\comapp(T_1)))\cap\mathcal N_{3\delta+2\lambda}(N(\comapp(T_2)))\right)\leq B,$$ where $\lambda$ is the constant from Proposition~\ref{prop:forward_ladder_quasiconvex}. \end{defn} \begin{rem}\label{rem:lol_and_wall} The purpose of the ladder overlap property is to guarantee that, when $L$ is large and $W$ is sub-quasiconvex, paths of the form $\beta\alpha\beta'$ are uniform quasigeodesics, where $\beta,\beta'$ are geodesics of carriers of slope-approximations and $\alpha$ is a vertical geodesic in $\comapp(\overline W)$. If the interiors of $\beta,\beta'$ have disjoint images in $\mathbf R$, Lemma~\ref{lem:vert_horiz_bound} ensures that $\beta\alpha\beta'$ is a uniform quasigeodesic. The interesting situations are those in which $\beta,\beta'$ are both incoming or both outgoing with respect to the vertical part of $\comapp(\overline W)$ containing $\alpha$. A thin quadrilateral argument shows that in either case, the ladder overlap property ensures that $\beta,\beta'$ have uniformly bounded coarse intersection, from which one concludes that $\beta\alpha\beta'$ is a uniform quasigeodesic (see Lemma~\ref{lem:RBRquasi} below). \end{rem} \section{Quasiconvex codimension-1 subgroups from immersed walls}\label{sec:quasiconvexity} In this section, we determine conditions ensuring that $\comappn{\overline W}$ is quasiconvex and $\overline W$ is a wall in $\widetilde X$. \subsection{Uniform quasiconvexity}\label{subsec:quasiconvexity_of_approximation} A collection $\{W\rightarrow X\}$ of immersed walls is \emph{uniformly sub-quasiconvex} if there exist constants $\mu_1,\mu_2$ such that $\comapp(\widetilde K)\hookrightarrow\widetilde X^1$ is a $(\mu_1,\mu_2)$-quasi-isometric embedding for each $W$ and each knockout $\widetilde K$ of $\overline W$. The first goal of this section is to prove: \begin{prop}\label{prop:quasiconvexity} Let $\mathbb W=\{W\rightarrow X\}$ be a uniformly sub-quasiconvex set of immersed walls with the ladder-overlap property. Then there exists $L_0,\kappa_1,\kappa_2$ such that for all $W\in\mathbb W$ with tunnel length at least $L_0$, the inclusion $\comappn{\overline W}\hookrightarrow\widetilde X^1$ is a $(\kappa_1,\kappa_2)$-quasi-isometric embedding. \end{prop} The constants are $\kappa_1=4\lambda_1\mu_1$ and $\kappa_2=\frac{\mu_2}{2}+2L_0(1+\frac{1}{4\lambda_1\mu_1})$. Here $\mu_1,\mu_2$ are the quasi-isometry constants from uniform sub-quasiconvexity, and $\lambda_1$ is the multiplicative quasi-isometry constant for 1-skeleta of forward ladders. We emphasize that these are independent of $L$ and of the collection of primary busts. \begin{proof}[Proof of Proposition~\ref{prop:quasiconvexity}] For a path $P$ in $\widetilde X^1$, as usual $\|P\|$ denotes the distance in $\widetilde X^1$ between the endpoints of $P$. If $P$ is a geodesic of $\comappn{\overline W}$, then its edge-length $|P|$ equals the distance in $\comappn{\overline W}$ between the endpoints of $P$. We will show that when $L\geq L_0$, there are constants $\kappa_1,\kappa_2$ such that $\|P\|\geq\kappa_1^{-1}|P|-\kappa_2$. \textbf{Alternating geodesics:} Let $P'$ be a geodesic in the graph $\comappn{\overline W}$. Suppose $P'$ \emph{alternates}, in the sense that $P'=\alpha_0\beta_1\alpha_1\cdots\beta_k\alpha_k$, where each $\alpha_i$ is a vertical geodesic path, and each $\beta_i$ is a geodesic of the 1-skeleton of a length-$L$ forward ladder (and thus a $(\lambda_1,\lambda_2)$-quasigeodesic). We allow the possibility that $\alpha_0$ or $\alpha_{k}$ has length~0. Each $\alpha_i$ is a $(\mu_1,\mu_2)$-quasigeodesic by our hypothesis that knockout-approximations are quasi-isometrically embedded. Since $W$ has the ladder overlap property, $\diam(\mathcal N_{3\delta+2\lambda}(\beta_i)\cap\mathcal N_{3\delta+2\lambda}(\beta_{i+1}))\leq B$. Let $B_0=\max(B,\Theta_{3\delta+2\lambda})$, where $\Theta_{3\delta+2\lambda}$ is as in Lemma~\ref{lem:vert_horiz_bound}. Applying Lemma~\ref{lem:RBRquasi} yields a constant $L_0$ such that, if $L\geq L_0$, then $\|P'\|\geq\frac{1}{4\lambda_1\mu_1}|P'|-\frac{\mu_2}{2}$. \textbf{$\comapp(\overline W)$ quasi-isometrically embeds:} Let $P$ be a geodesic of $\comappn{\overline W}$. By construction $P=\beta_0'P'\beta_{k+1}'$ where $P'$ is alternating and $\beta_0',\beta_{k+1}'$ are (possibly trivial) paths in forward ladders. If $|\beta_0'|,|\beta_{k+1}'|\geq L_0$, then $\|P\|\geq\frac{1}{4\lambda_1\mu_1}|P|-\frac{\mu_2}{2}$ by Lemma~\ref{lem:RBRquasi}. If $|\beta_0'|,|\beta_{k+1}'|\leq L_0$, then since $P'$ is alternating, $$ \|P\|\geq\frac{1}{4\lambda_1\mu_1}|P'|-\frac{\mu_2}{2}-2L_0\geq\frac{1}{4\lambda_1\mu_1}|P|-\frac{\mu_2}{2}-2L_0(1+\frac{1}{4\lambda_1\mu_1}). $$ In the remaining case, without loss of generality, $P=\beta_0'P''$, where $|\beta_0'|\leq L_0$ and $P''$ satisfies $\|P''\|\geq\frac{1}{4\lambda_1\mu_1}|P''|-\frac{\mu_2}{2}$ by Lemma~\ref{lem:RBRquasi}. The proof is thus complete with $\kappa_1=4\lambda_1\mu_1$ and $\kappa_2=\frac{\mu_2}{2}+2L_0(1+\frac{1}{4\lambda_1\mu_1})$.\end{proof} \begin{lem}\label{lem:hsuwiseRBR} Let $Z$ be $\delta$-hyperbolic, and let $P=\alpha_0\beta_1\alpha_1\cdots\beta_k\alpha_k$ be a path in $Z$ with all $\alpha_i$ and $\beta_i$ geodesic. Suppose there exists $B\geq 0$ such that for all $i$, each intersection below has diameter $\leq B$: $$\mathcal N_{3\delta}(\beta_i)\cap\beta_{i+1},\,\,\,\,\,\,\,\mathcal N_{3\delta}(\beta_i)\cap\alpha_i,\,\,\,\,\,\,\,\mathcal N_{3\delta}(\beta_i)\cap\alpha_{i-1}.$$ Then if $|\beta_i|\geq12(B+\delta)$ for each $i$, then $\|P\|\geq\frac{1}{2}\left(\sum_{i=0}^{k}|\alpha_i|+\sum_{i=1}^k|\beta_i|\right)$. \end{lem} \begin{proof} This is a standard argument. We refer, for instance, to~\cite[Thm~2.3]{HsuWiseCubulatingMalnormal}. \end{proof} We now promote Lemma~\ref{lem:hsuwiseRBR} to a statement about piecewise-quasigeodesics. \begin{lem}\label{lem:RBRquasi} Let $Z$ be $\delta$-hyperbolic and let $P=\alpha_0\beta_1\alpha_1\cdots\beta_k\alpha_k$ be a path in $Z$ such that each $\beta_i$ is a $(\lambda_1,\lambda_2)$-quasigeodesic and each $\alpha_i$ is a $(\mu_1,\mu_2)$-quasigeodesic. Suppose that for each $R\geq 0$ there exists $B_R\geq 0$ such that for all $i$, each intersection below has diameter $\leq B_R$: $$\mathcal N_{3\delta+R}(\beta_i)\cap\beta_{i+1},\,\,\,\,\,\,\,\mathcal N_{3\delta+R}(\beta_i)\cap\alpha_i,\,\,\,\,\,\,\,\mathcal N_{3\delta+R}(\beta_i)\cap\alpha_{i-1}.$$ Then there exists $L_0$ such that, if $|\beta_i|\geq L_0$ for each $i$, then $\|P\|\geq\frac{1}{4\lambda_1\mu_1}|P|-\frac{\mu_2}{2}$. \end{lem} \begin{proof} For each $i$, let $\bar\alpha_i$ [respectively, $\bar\beta_i$] be a geodesic with the same endpoints as $\alpha_i$ [respectively, $\beta_i$], and let $\overline P=\bar\alpha_0\bar\beta_1\bar\alpha_1\cdots\bar\beta_k\bar\alpha_k$ be a piecewise-geodesic with the same endpoints as $P$. Since $Z$ is $\delta$-hyperbolic, there exists $\mu=\mu(\mu_1,\mu_2,\delta)$ such that $\alpha_i$ and $\bar\alpha_i$ lie at Hausdorff distance at most $\mu$, and there exists $\lambda=\lambda(\lambda_1,\lambda_2,\delta)$ such that $\bar\beta_i$ and $\beta_i$ lie at Hausdorff distance at most $\lambda$. Note that if $R_1\leq R_2$, then we may assume $B_{R_1}\leq B_{R_2}$. By hypothesis, $\mathcal N_{3\delta+2\lambda}(\beta_i)\cap\beta_{i+1}$ has diameter $\leq B_{2\lambda}$. Moreover, if $\bar\beta'\subset\bar\beta_i$ is a subpath that $3\delta$-fellowtravels with a subpath $\bar\alpha'$ of $\bar\alpha_i$ or $\bar\alpha_{i-1}$, then $\beta'$ fellowtravels at distance $3\delta+\mu+\lambda$ with a subpath $\alpha''$ of $\alpha_i$ or $\alpha_{i-1}$, whence $|\beta'|\leq B_{\mu+\lambda}$ by hypothesis. Letting $L_0\geq12(\delta+B_{\mu+2\lambda})$ and applying Lemma~\ref{lem:hsuwiseRBR} shows that $\overline P$ is a $(\frac{1}{2},0)$-quasigeodesic, and we have: \begin{equation}\label{eqn:piecewisegeodesic} \|P\|=\left\|\overline P\right\|\geq\frac{1}{2}\left|\overline P\right|. \end{equation} Since $\mu_1,\lambda_1\geq 1$, we can bound $|\overline P|$ as follows: \begin{eqnarray*} \left|\overline P\right|=\sum_{i=1}^k|\bar\beta_i|+\sum_{i=0}^k|\bar\alpha_i|&\geq&\sum_{i=1}^k(\lambda_1^{-1}|\beta_i|-\lambda_2)+\sum_{i=0}^k(\mu_1^{-1}|\alpha_i|-\mu_2)\\ &=&\left[\lambda_1^{-1}\sum_{i=1}^k\left(|\beta_i|-\lambda_1(\lambda_2+\mu_2)\right)+\mu_1^{-1}\sum_{i=0}^k|\alpha_i|\right]-\mu_2\\ &\geq&(\lambda_1\mu_1)^{-1}\left[\sum_{i=1}^k(|\beta_i|-\lambda_1(\lambda_2+\mu_2))+\sum_{i=0}^k|\alpha_i|\right]-\mu_2.\\ \end{eqnarray*} If $L_0\geq 2\lambda_1(\lambda_2+\mu_2)+1$, then, provided that $|\beta_i|\geq L\geq L_0$, we have: \[\left|\overline P\right|\geq\frac{1}{2\lambda_1\mu_1}\left[\sum_{i=1}^k|\beta_i|+\sum_{i=0}^k|\alpha_i|\right]-\mu_2.\] Combining this with Equation~\eqref{eqn:piecewisegeodesic} yields $\|P\|\geq\frac{1}{4\lambda_1\mu_1}|P|-\frac{\mu_2}{2}$. \end{proof} \subsection{$\overline W$ is a wall when tunnels are long}\label{subsec:W_is_wall} A subspace $Y\subset\widetilde X$ is a \emph{wall} if $\widetilde X-Y$ has exactly two components, each of which is stabilized by $\stabilizer(Y)$. Note that this definition is stricter than usual. For more about wallspaces and the various definitions, background, and references, see~\cite{HruskaWise}. Our goal is now to show that if $W\rightarrow X$ is an immersed wall with sufficiently long tunnels, then $\overline W$ is a wall. We need the following useful consequence of quasiconvexity. \begin{prop}\label{prop:approximation_is_tree} Let $\mathbb W$ satisfy the hypotheses of Proposition~\ref{prop:quasiconvexity}. There exists $L_1\geq L_0$ such that $\comapp(\overline W)$ is a tree for each $W\in\mathbb W$ with tunnel length $L\geq L_1$. \end{prop} \begin{proof} Let $Q$ be an immersed path in $\comapp(\overline W)$, and let $Q'$ be a geodesic of $\comappn{\overline W} $ with the same endpoints as $Q$. Proposition~\ref{prop:quasiconvexity} implies that $\|Q'\|\geq\kappa_1^{-1}|Q|-\kappa_2$. Hence if $|Q'|\geq L_1=\max(L_0,\kappa_1\kappa_2+\kappa_1)$, then $Q$ is not closed. Any immersed path $Q$ in $\comapp(\overline W)$ either lies in a single vertex space and is thus not closed, or contains a slope approximation and thus $Q'$ has length at least $L$. \end{proof} \begin{rem}[Tree of spaces structure on $\overline W$]\label{rem:wall_is_tree_of_spaces} Proposition~\ref{prop:approximation_is_tree} justifies our claim in Remark~\ref{rem:wall_structure} that $\overline W$ is a tree of spaces when $L$ is sufficiently large, assuming that $W$ is sub-quasiconvex and has the ladder overlap property. Indeed, any cycle in $\overline W$ that is not contained in a knockout will map to a cycle in $\comapp(\overline W)$, contradicting Proposition~\ref{prop:approximation_is_tree}. \end{rem} \begin{prop}\label{prop:tunneliswall} Let $W\rightarrow X$ be an immersed wall in a collection $\mathbb W$ satisfying the hypotheses of Proposition~\ref{prop:quasiconvexity}. There exists $L_1\geq L_0$ such that the image $\overline W\subset\widetilde X$ of $\widetilde W\rightarrow\widetilde X$ is a wall provided that $W$ has tunnel length $L\geq L_1$. \end{prop} \begin{proof} Since $\ensuremath{{\sf{H}}}^1(\widetilde X)=0$, it suffices to show that $\overline W$ has an open neighborhood homeomorphic to $\overline W \times [-1,1]$ with $\overline W$ identified with $\overline W\times \{0\}$. The local homeomorphism $W\times[-1,1]\rightarrow X$ lifts to a map $\widetilde W\times[-1,1]\rightarrow \widetilde X$. We note as well that $W\rightarrow X$ is locally two-sided by construction. The image of $\widetilde W\times[-1,1]\rightarrow\widetilde X$ would provide the desired neighborhood $\overline W \times [-1,1]$ provided that this map is a covering map onto its image. By choosing the image of $W\times[-1,1]$ to be sufficiently narrow, the only place where this could fail is where distinct slopes of $\overline W$ intersect. To exclude this possibility, we will show that distinct tunnels $T_0,T_k$ of $\overline W$ are disjoint. Suppose that $T_0\neq T_k$ and $T_0\cap T_k\neq\emptyset$. Let $e$ be the primary busted edge dual to $T_0$ and $T_k$. Since $T_0,T_k\subset\overline W$, there exists a path $P\rightarrow\overline W$ that starts on $T_0$, ends on $T_k$, and which is disjoint from the interiors of $T_0$ and $T_k$. Indeed, let $\widetilde P\rightarrow\widetilde W$ be a path joining lifts of $\widetilde T_0,\widetilde T_k$ and let $P$ be the image of $\widetilde P$ in $\overline W$. Moreover, we assume that $\widetilde P$ is minimal in the sense that it is disjoint from intervening lifts of $T_0,T_k$. The minimality of $\widetilde P$ ensures that $P$ has the desired property. There are three cases. The first is where $P$ starts and ends on the levels of $T_0,T_k$. The second is where $P$ starts and ends on the slopes of $T_0,T_k$. The third case is where $P$ starts on the level of (say) $T_0$ and ends on the slope of $T_k$. In the first case, the approximation $\comapp(P)$ of the image of $P$ is a connected subspace of $\comapp(\overline W)$ that contains the endpoints of $e$ but does not contain the entire edge $e$. Hence $\comapp(P)\cup e$ contains a cycle. Since $\comapp(P)\subset\comapp(\overline W)$ and $e\subset\comapp(\overline W)$, there is a cycle in $\comapp(\overline W)$, contradicting Proposition~\ref{prop:approximation_is_tree}. Similarly, in the second case, $\comapp(P)$ is disjoint from $e$, so that $\comapp(P)\cup e\cup\comapp(T_0)\cup\comapp(T_k)$ is a subspace of $\comapp(\overline W)$ that contains a cycle. In the third case, the contradictory subspace is $\comapp(P)\cup\comapp(T_k)\cup e$. \end{proof} We note the following corollary: \begin{cor}\label{cor:quasiconvexsubgroup} Let $\mathbb W$ be a set of sub-quasiconvex immersed walls such that $\mathbb W$ has the ladder overlap property. Then there exists $L_1$ such that for all $W\in\mathbb W$ with tunnel length $L\geq L_1$, the stabilizer $H_W\leq G$ of $\overline W$ is a quasiconvex, codimension-1 free subgroup. \end{cor} \section{Cutting geodesics}\label{sec:cutting_geodesics} In this section, we recall the criterion for cocompact cubulation of hyperbolic groups given in~\cite{BergeronWise} and describe how a sufficiently rich collection of quasiconvex walls in $\widetilde X$ ensures that this criterion is satisfied. \newcommand{\moustache}[1]{\overline{{#1}}_L} \subsection{Separating points on $\partial\widetilde X$}\label{subsec:cutting_definition} Let $\partial\widetilde X$ denote the Gromov boundary of $\widetilde X^1$. Let $W\rightarrow X$ be an immersed wall with the property that $\comappn{\overline W}$ is quasiconvex in $\widetilde X^1$ and $\overline W$ is a wall. Let $\overleftarrow{W}$ and $\overrightarrow{W}$ be the components of $\widetilde X-\overline W$, and let $N(\overleftarrow{W}),N(\overrightarrow{W})$ be the smallest subcomplexes containing $\overleftarrow{W},\overrightarrow{W}$ respectively. Then $N(\overleftarrow{W})^1\cap N(\overrightarrow{W})^1=N(\overline W)^1$, which is coarsely equal to $\comappn{\overline W}$. Let $\partial\overline W$ denote $\partial N(\overline W)^1=\partial\comappn{\overline W}$, which is a closed subset of $\partial\widetilde X$ since $\comappn{\overline W}$ is quasiconvex in $\widetilde X^1$. Let $\partial\overleftarrow{W}=\partial N(\overleftarrow{W})^1-\partial\overline W$ and let $\partial\overrightarrow{W}=\partial N(\overrightarrow{W})^1-\partial\overline W$, so that $\partial\overleftarrow{W}$ and $\partial\overrightarrow{W}$ are disjoint open subsets of $\partial\widetilde X$. Note that $\partial\overleftarrow{W}$ and $\partial\overrightarrow{W}$ are $H_W$-invariant, since $N(\overleftarrow{W})$ and $N(\overrightarrow{W})$ are $H_W$-invariant by Remark~\ref{rem:invariant_halfspace}. Let $p,q\in\partial\widetilde X$ be the endpoints of a bi-infinite geodesic $\gamma:\mathbf R\rightarrow\widetilde X^1$. Then $\gamma$ is \emph{cut} by $\overline W$ if $p\in\partial\overleftarrow{W}$ and $q\in\partial\overrightarrow{W}$ or vice versa. The following holds by~\cite[Thm~1.4]{BergeronWise}: \begin{prop}\label{prop:cutting_criterion} Suppose that for every geodesic $\gamma:\mathbf R\rightarrow\widetilde X^1$, there exists a wall $\overline W$ of the type described in Section~\ref{sec:immersed_walls}, such that $N(\overline W)$ is quasiconvex and such that $\overline W$ cuts $\gamma$. Then there exists a $G$-finite collection $\{\overline W\}$ of walls in $\widetilde X$ such that $G$ acts freely and cocompactly on the dual CAT(0) cube complex. \end{prop} \subsection{A method for cutting the two types of geodesics}\label{subsec:5.2} \begin{defn}[Ladderlike, deviating]\label{defn:level_like_deviating} Let $M\geq 0$ and let $\gamma\subset\widetilde X^1$ be an embedded infinite or bi-infinite path whose image is $\xi$-quasiconvex, for some $\xi\geq 0$. Then $\gamma$ is \emph{$M$-ladderlike} if there exists a forward ladder $N(\sigma)$, where $\sigma$ is a forward path of length $M$, such that a geodesic of $N(\sigma)$ joining the endpoints of $\sigma$ fellow-travels with a subpath of $\gamma$ at distance $2\delta+\lambda+\xi$. Here, $\widetilde X^1$ is $\delta$-hyperbolic and $\lambda$ is the constant from Proposition~\ref{prop:forward_ladder_quasiconvex}. Otherwise, $\gamma$ is \emph{$M$-deviating} \end{defn} Note that if $\gamma$ is $M$-deviating, for each $R\geq 0$ there exists $M_R$ depending only on $\xi,M,R$ such that for all forward paths $\sigma$, we have $\diam(\gamma\cap\mathcal N_R(N(\sigma)^1))\leq M_R$. Moreover, if $\gamma$ is $2M$-deviating, the same bound holds with $\sigma$ replaced by any level, since any geodesic in a level decomposes as the concatenation of two (possibly trivial) forward paths. \begin{defn}[Many effective walls]\label{defn:many_effective_walls} A set $\mathbb W$ of immersed walls in $X$ is \emph{spreading} if: \begin{itemize} \item For arbitrarily large $L$, there exists $W\in\mathbb W$ with tunnel length $L$. \item $\mathbb W$ has the ladder overlap property of Definition~\ref{defn:ladder_overlap_property}. \end{itemize} $\widetilde X$ has \emph{many effective walls} if Conditions~\eqref{item:bust_point} and~\eqref{item:w_a_periodic} below hold. \begin{enumerate} \item \label{item:bust_point} For each regular $y\in V$, there exists a spreading set $\mathbb W$ such that for each $\epsilon>0$ and each $m\in\ensuremath{\field{N}}$, there exists $L>m$ and $W\in\mathbb W$ with tunnel length $L$, a primary bust in each edge of $V$, and a primary bust in the $\epsilon$-neighborhood of $y$. \item \label{item:w_a_periodic} For each $a\in\widetilde V_0$, whose image in $V$ is periodic and whose corresponding point in $\widetilde E_0$ is denoted by $a'$, there exists $k=k(a)\geq 0$ such that the following set $\mathbb W_a$ of immersed walls in $X$ is uniformly sub-quasiconvex and spreading: $\mathbb W_a$ is the set of $W$ such that each edge of $V$ contains a primary bust of $W$, and such that for each primary bust $d'$ of $\overline W\cap\widetilde E_0$ (corresponding to an interval $d\subset\widetilde V_0$) that is joined to $a'\in\widetilde E_0$ by a path in a knockout of $\overline W$, we have $\textup{\textsf{d}}_{\widetilde X^1}(\tilde\phi^n(a),\tilde\phi^n(d))\geq3\delta+2\lambda$ for all $n\geq k$. See Figure~\ref{fig:mew}. \end{enumerate} \end{defn} \begin{figure} \begin{overpic}[scale=0.10]{mew} \put(1,85){$d$} \put(1,2){$a$} \put(20,85){$d'$} \put(19,2){$a'$} \end{overpic} \caption{Definition~\ref{defn:many_effective_walls}.\eqref{item:w_a_periodic}.} \label{fig:mew} \end{figure} \begin{rem}\label{rem:uniform_k} The constant $k$ in Definition~\ref{defn:many_effective_walls} can be chosen independently of the point $a$. For each $a$, let $k'(a)$ be chosen so that for each bust $d$ and each $n\geq k'(a)$, we have $\textup{\textsf{d}}_{\widetilde X^1}(\tilde\phi^n(a),\tilde\phi^n(d))\geq3\delta+2\lambda+1$. This is possible since the existence of $k(a)$ implies that the forward rays emanating from $a$ and any point of $d$ diverge. Fix $\epsilon\in(0,1)$ and let $U_a=(\tilde\phi^{k'(a)})^{-1}(\mathcal N_{\epsilon}(\tilde\phi^{k'(a)}(a))$. Then for each $b\in U_a$, we have $\textup{\textsf{d}}_{\widetilde X^1}(\tilde\phi^n(b),\tilde\phi^n(d))\geq3\delta+2\lambda+1-\epsilon>3\delta+2\lambda$ for each $n\geq k'(a)$. Hence we have $k(b)\leq k'(a)$. The closure of the subset $V^{\circlearrowright}\subseteq V$ consisting of periodic points is compact since $V$ is compact. Hence the claim follows since the open covering of $V^{\circlearrowright}$ produced above has a finite subcovering. Similarly, since the collection $\mathbb W_a$ has the ladder overlap property, there is likewise a constant $B_a$ such that for all $W\in\mathbb W_a$, any two tunnels $T,T'$ of $\overline W$ that are joined by a path in $\overline W$ not traversing a slope have the property that the $(3\delta+2\lambda)$-neighborhoods of $\comapp(T),\comapp(T')$ intersect in a set of diameter at most $B_a$. Let $V_a$ be the set of images in $V$ of points $b\in \widetilde V$ such that for all $W\in\mathbb W_a$, the point $b$ lies in a nucleus of some $\overline W$ and for all primary busts $d$ of $\overline W$, we have $\textup{\textsf{d}}_{\widetilde X^1}(\tilde\phi^n(a),\tilde\phi^n(d))\geq3\delta+2\lambda$ for all $n\geq k$. The previous argument showed that, with $k$ chosen appropriately, the set $V_a$ is open. It follows that if $\widetilde X$ has many effective walls, the ladder overlap constant $B_a$ can be chosen independently of $a$. This is used in the proof of Proposition~\ref{prop:cutting_ladderlike}. \end{rem} \begin{defn}[Separating level]\label{defn:separating_level} $\widetilde X$ is \emph{$(M,K)$-separated} if for each $M$-deviating geodesic $\gamma$ there exists $y\in\widetilde X$ such that the following holds for all sufficiently large $n$: the set $\gamma\cap T^o_n(\tilde\phi^n(y))$ has odd cardinality, and the distance in $T^o_n(\tilde\phi^n(y))$ from $\gamma\cap T^o_n(\tilde\phi^n(y))$ to the root or to any leaf of $T^0_n(\tilde\phi^n(y))$ exceeds $M+K$. We say $\widetilde X$ is \emph{level-separated} if it is $(M,K)$-separated for all $M>0,K\geq0$. \end{defn} \begin{rem}\label{rem:nearly_fixed} If the level $T_n^o(\tilde\phi^n(y))$ separates $\gamma$ in the above sense, then we can choose $y$ so that the image $\bar y\in V$ of $y$ is not periodic. Indeed, if $y',y$ are sufficiently close, then $T_n^o(\tilde\phi^n(y))$ and $T_n^o(\tilde\phi^n(y'))$ both separate $\gamma$. There are points $y'$ arbitrarily close to $y$ whose images in $V$ are not periodic since there are only countably many periodic points. \end{rem} \begin{defn}[Bounded level-intersection]\label{defn:bounded_level_intersection} $\widetilde X$ has \emph{bounded level-intersection} if for each $z\in\widetilde X^1$ and each vertical edge $e\subset\widetilde X^1$, there exists $K=K(z,e)$ such that for every level $T$ with a leaf at $z$, we have $|T\cap e|\leq K$. \end{defn} \begin{rem}\label{rem:bounded_level_intersection} In the case of greatest interest, where $X$ is the mapping torus of a train track map, each level intersects each vertical edge in at most a single point, and hence $\widetilde X$ has bounded level-intersection. This holds in particular for the complexes $\widetilde X$ considered in Theorem~\ref{thm:irreducible}. More generally, this holds whenever there is a continuous map from $\widetilde X$ to an $\ensuremath{\field{R}}$-tree that is constant on levels and sends edges to concatenations of finitely many arcs. \end{rem} \begin{defn}[Exponentially expanding]\label{defn:exponentially_expanding} The train track map $\phi:V\rightarrow V$ is \emph{exponentially expanding} if there exists an \emph{expansion constant} $\varpi>1$ such that for all edges $e$ of $V$ and all arcs $\alpha\subset e$, and all $L\geq 0$, we have $|\phi^L(\alpha)|\geq\varpi^L|\alpha|$. Note that if $\phi$ is an irreducible train track map and edges are expanding, then $\phi$ is exponentially expanding, as can be seen by taking $\varpi$ to be the Perron-Frobenius eigenvalue of the transition matrix of $\phi$. See Section~\ref{sec:forward_space_in_train_track_case} for more on the eigenvalues of the transition matrix. \end{defn} The main result of this section is: \begin{prop}\label{prop:everything_cut} Suppose that $\phi:V\rightarrow V$ is a $\pi_1$-injective train track map. Let $X$ be the mapping torus of $\phi$. Suppose that $\pi_1X$ is word-hyperbolic and that $\widetilde X$ satisfies: \begin{enumerate} \item $\widetilde X$ is level-separated. \item $\widetilde X$ has many effective walls. \item Every finite forward path fellow-travels at uniformly bounded distance with a periodic forward path. \item $\phi$ is exponentially expanding. \end{enumerate} Then $G$ acts freely and cocompactly on a CAT(0) cube complex. \end{prop} \begin{proof} Proposition~\ref{prop:cutting_ladderlike} shows that there exists $M$ such that every $M$-ladderlike geodesic is cut by a wall. Proposition~\ref{prop:cutting_deviating} shows that each $M$-deviating geodesic is cut by a wall; Proposition~\ref{prop:cutting_deviating} requires $\widetilde X$ to have bounded level-intersection, which is the case since $\phi$ is a train track map. The claim then follows from Proposition~\ref{prop:cutting_criterion} since each geodesic that is not $M$-ladderlike is by definition $M$-deviating. \end{proof} \begin{conv} In the remainder of this section, $\phi:V\rightarrow V$ is assumed to satisfy the initial hypotheses of Proposition~\ref{prop:everything_cut}, except that the enumerated hypotheses will be invoked as needed. \end{conv} \subsection{Walls in $\widetilde X_L$}\label{subsec:walls_in_long_space} Let $W\rightarrow X$ be an immersed wall with tunnel length $L\geq 1$, and suppose that $\overline W$ is a wall and $\comappn{\overline W}$ is $(\kappa_1,\kappa_2)$-quasi-isometrically embedded and $\kappa$-quasiconvex. Each primary bust has regular endpoints, by Lemma~\ref{lem:choosing_busts_1}.\eqref{item:regular_endpoints}, so that each level-part of $\overline W$ is disjoint from $\widetilde X^0$. Similarly, $\widetilde X^0$ is disjoint from $\comapp(S)$ for each slope $S$ of $\overline W$. Recall that $\widetilde X^{\bullet}_L$ denotes the subdivision of $\widetilde X_L$ obtained by pulling back the 1-skeleton of $\widetilde X$. For each $n\in\ensuremath{\field{Z}}$, the inclusion $\widetilde V_{nL}\hookrightarrow\widetilde X$ lifts to an embedding $\widetilde V_{nL}\hookrightarrow({\widetilde X^\bullet}_L)^1$, and we continue to use the notation $\widetilde V_{nL}$ for this subspace. We make the same observation and convention about $\widetilde E_{nL}$. By translating, we can assume that $\overline W$ has a primary bust in $\widetilde V_0$, and hence all primary busts in $\overline W$ lie in the various $\widetilde V_{nL}$ and the map $\widetilde W\rightarrow\widetilde X$ lifts to $\widetilde W\rightarrow\widetilde X^{\bullet}_L$. Let $\overline W_L$ be the image of $\widetilde W\rightarrow\widetilde X^{\bullet}_L$, so that we have the commutative diagram: \begin{center} $ \begin{diagram} \node{\moustache{W}}\arrow{e}\arrow{s}\node{{\widetilde X^\bullet}_L}\arrow{s}\\ \node{\overline W}\arrow{e}\node{\widetilde X} \end{diagram} $ \end{center} \renewcommand{\overleftarrow{W}}{\overleftarrow W_L} \renewcommand{\overrightarrow{W}}{\overrightarrow W_L} Note that $\moustache{W}$ and $\overline W$ are very similar: each tunnel $\moustache T$ of $\moustache{W}$ consists of a slope and a level-part that is a (subdivided) star, and $\overline W$ is obtained from $\moustache{W}$ by folding each such subdivided star into a tree. The halfspaces $\overleftarrow{W},\overrightarrow{W}$ in ${\widetilde X^\bullet}_L$ associated to $\overline W_L$ respectively map to the halfspaces $\overleftarrow W,\overrightarrow W$ in $\widetilde X$. The approximation map $\comapp$ is defined in $\widetilde X_L$ just as it is in $\widetilde X=\widetilde X_1$. Consider $\comapp:\moustache{W}\rightarrow\widetilde X_L$, which is a lift of $\comapp:\overline W\rightarrow\widetilde X$. There is a corresponding commutative diagram: \begin{center} $ \begin{diagram} \node{\comapp(\moustache{W})}\arrow{e}\arrow{s}{}\node{{\widetilde X^\bullet}_L}\arrow{s}\\ \node{\comapp(\overline W)}\arrow{e}\node{\widetilde X} \end{diagram} $ \end{center} in which the map $\comapp(\overline W_L)\rightarrow\comapp(\overline W)$ is an isomorphism. Thus $\comapp(\overline W)\rightarrow\widetilde X$ lifts to an embedding $\comapp(\overline W)\rightarrow\widetilde X^{\bullet}_L$ whose image is $\comapp(\overline W_L)$. Figure~\ref{fig:juxtaposition} depicts $\moustache{W}$ and $\comapp(\overline W_L)$. There is also a lift $\comappn{\overline W}\hookrightarrow{\widetilde X^\bullet}_L$. Since $\comappn{\overline W}\hookrightarrow\widetilde X^1$ factors as $\comappn{\overline W}\hookrightarrow({\widetilde X^\bullet}_L)^1\rightarrow\widetilde X^1$ and since $({\widetilde X^\bullet}_L)^1\rightarrow\widetilde X^1$ is distance nonincreasing, $\comappn{\overline W}\rightarrow({\widetilde X^\bullet}_L)^1$ is a $(\kappa_1,\kappa_2)$-quasi-isometric embedding. Thus $\partial\comappn{\overline W}$ embeds in $\partial{\widetilde X^\bullet}_L$ as a closed subset. The following proposition explains that the tree $\comapp(\overline W_L)$ determines a wall in ${\widetilde X^\bullet}_L$, and therefore determines a coarse wall in $\widetilde X$ that coarsely agrees with $\overline W$. \begin{figure} \includegraphics[width=0.35\textwidth]{juxtaposition}\\ \caption{$\moustache{W}$ and $\comapp(\overline W)$ inside ${\widetilde X^\bullet}_L$.}\label{fig:juxtaposition} \end{figure} \begin{prop}\label{prop:approximation_separates} ${\widetilde X^\bullet}_L$ contains subspaces $\overleftarrow A,\overrightarrow A$ such that $\overleftarrow A\cup\overrightarrow A={\widetilde X^\bullet}_L$ and $\overleftarrow A\cap\overrightarrow A=\comapp(\overline W_L)$. Both $\overleftarrow A-\comapp(\overline W_L)$ and $\overrightarrow A-\comapp(\overline W_L)$ are connected. Moreover, the images of $\overleftarrow A$ and $\overrightarrow A$ under the map ${\widetilde X^\bullet}_L\rightarrow\widetilde X$ are coarsely equal to $\overleftarrow{W}$ and $\overrightarrow{W}$. \end{prop} \begin{proof} It suffices to produce the subspaces $\overleftarrow A,\overrightarrow A$ so that each is coarsely equal to a component of ${\widetilde X^\bullet}_L-\moustache{W}$. Let $\overleftarrow{W},\overrightarrow{W}$ be the closures of the components of ${\widetilde X^\bullet}_L-\moustache{W}$. The halfspaces $\overleftarrow A$ and $\overrightarrow A$ will be obtained from $\overleftarrow{W}$ and $\overrightarrow{W}$ by adding and subtracting ``discrepancy zones'', which are subspaces between $\moustache{W}$ and $\comapp(\moustache{W})$ suggested by Figure~\ref{fig:juxtaposition}. \textbf{Discrepancy zones:} Let $e\subset\widetilde V_{nL}$ be a primary busted edge with outgoing long 2-cell $R_e\subset{\widetilde X^\bullet}_L$. Let $d\subset e$ be the closed primary bust with endpoints $p,q$. Let $p',q'$ be the points at distance $\frac{1}{2}$ to the right of $p,q$ within $R_e$. The slope $S$ travels from $p$ to $q'$, as shown in Figure~\ref{fig:upward_intermediate_zone}. Let $Z^{\uparrow}$ be the 2-simplex in $R_e$ bounded by $S$ and the part of $\comapp(S)$ between $p$ and $q'$. The disc $Z^{\uparrow}$ is an \emph{upward discrepancy zone}. Let $\widetilde C\subset\widetilde E_{nL}$ be a nucleus in $\overline W_L$ and let $\comapp(\widetilde C)\subset\widetilde V_{nL+L}$ be its approximation. Consider the map $\widetilde C\times[\frac{1}{2},L]\rightarrow{\widetilde X^\bullet}_L$ that restricts to the inclusion $\widetilde C\times\{t\}\hookrightarrow\widetilde V_{nL}\times\{t\}\subset{\widetilde X^\bullet}_L$ for $t<L$ and acts as the map $\tilde\phi^L:\widetilde C\rightarrow\widetilde V_{nL+L}$ on $\widetilde C\times\{L\}$. The image of this map is a \emph{downward discrepancy zone} $Z^{\downarrow}$. In other words, $Z^{\downarrow}$ is the closure of $\widetilde C\times[\frac{1}{2},L)$ in $\widetilde X_L^\bullet$. See Figure~\ref{fig:downward_intermediate_zone}. \begin{figure}[ht] \begin{overpic}[scale=0.15]{upward_intermediate_zone} \put(6,6){$p$} \put(6,26){$q$} \put(31,26){$q'$} \end{overpic} \caption{An upward discrepancy zone.}\label{fig:upward_intermediate_zone} \end{figure} \begin{figure}[ht] \includegraphics[width=0.5\textwidth]{downward_intermediate_zone} \caption{A downward discrepancy zone is shaded.}\label{fig:downward_intermediate_zone} \end{figure} \textbf{The halfspaces $\overleftarrow A$ and $\overrightarrow A$:} Let $\mathfrak Z^{\uparrow}$ be the union of all upward discrepancy zones associated to $\comapp(\overline W_L)$, and likewise let $\mathfrak Z^{\downarrow}$ be the union of all downward discrepancy zones. Let $$\overleftarrow A=\closure{\left(\overleftarrow{W}-\mathfrak Z^{\uparrow}\right)\cup\mathfrak Z^{\downarrow}}\text{\,\,\,\,\,\,and\,\,\,\,\,\,}\overrightarrow A=\closure{\left(\overrightarrow{W}-\mathfrak Z^{\downarrow}\right)\cup\mathfrak Z^{\uparrow}}.$$ Since each discrepancy zone lies at distance less than $L$ from $\overline W_L$, we see that $\overleftarrow A$ coarsely equals $\overleftarrow{W}$. By construction, $\overleftarrow A\cup\overrightarrow A={\widetilde X^\bullet}_L$. Finally, suppose that $x\in\overleftarrow A\cap\overrightarrow A$. Then $x$ must lie on the boundary of an discrepancy zone. If $x\in\overline W_L$, and $x\in\mathfrak Z^{\uparrow}$, then $x\not\in\overleftarrow A$ unless $x\in\comapp(\overline W_L)\cap\overline W_L$. Similarly, if $x\in\overline W_L$ and $x\in\mathfrak Z^{\downarrow}$, then $x\not\in\overrightarrow A$ unless $x\in\comapp(\overline W_L)\cap\overline W_L$. Hence $\overleftarrow A\cap\overrightarrow A\subseteq\comapp(\overline W_L)$. On the other hand, every point in $\comapp(\overline W_L)$ lies in the boundary of an discrepancy zone, and thus $\comapp(\overline W_L)\subseteq\overleftarrow A\cap\overrightarrow A$. Observe that $\overleftarrow A-\comapp(\overline W_L)$ is homeomorphic to $\overleftarrow{W}-\overline W_L$, which is connected. Likewise $\left(\overrightarrow A-\comapp(\overline W_L)\right)\cong\left(\overrightarrow{W}-\overline W_L\right)$. Hence $\overleftarrow A-\comapp(\overline W_L)$ and $\overrightarrow A-\comapp(\overline W_L)$ are connected. \end{proof} \renewcommand{\overleftarrow{W}}{\overleftarrow W} \renewcommand{\overrightarrow{W}}{\overrightarrow W} \subsection{Lifting and cutting geodesics in ${\widetilde X^\bullet}_L$}\label{subsec:lift_and_cut} We now describe a criterion ensuring that a given geodesic in $\widetilde X$ is cut by a wall, in terms of quasigeodesics and walls $(\overleftarrow A, \overrightarrow A)$ in ${\widetilde X^\bullet}_L$. \subsubsection{Lifted augmentations of geodesics}\label{subsubsec:lifted_augmentations} The following construction adjusts a bi-infinite quasigeodesic $\gamma\rightarrow\widetilde X^1$ so that it can be lifted to a bi-infinite quasigeodesic $\widehat{\gamma_{_{\succ}}}\rightarrow{\widetilde X^\bullet}_L$ such that $\gamma$ and $\widehat{\gamma_{_{\succ}}}$ determine the same pair of points in $\partial\widetilde X\cong\partial{\widetilde X^\bullet}_L$. \begin{cons}[Lifted augmentations of quasigeodesics]\label{cons:LA} Let $\gamma:\mathbf R\rightarrow\widetilde X^1$ be an embedded quasigeodesic. The \emph{augmentation} $\gamma_{_{\succ}}$ of $\gamma$ is defined as follows. For each (possibly trivial) bounded maximal horizontal subpath $P\subset\gamma$, with endpoints $p,p'\in\widetilde V_n,\widetilde V_{n'}$, let $n''$ be the smallest multiple of $L$ greater than or equal to $\max\{n,n'\}$ and let $p''=\tilde\phi^{n''-n}(p)=\tilde\phi^{n''-n'}(p')$. Let $Q'$ be the horizontal path $pp''p'$, and replace $P$ by $Q'$. Performing this replacement for each such $P$ yields $\gamma_{_{\succ}}$. Note that $\gamma_{_{\succ}}$ is a quasigeodesic that $L$-fellowtravels with $\gamma$, so that $\partial\gamma_{_{\succ}}=\partial\gamma$. We use the following notation. First, $P=P_1P_2$, where $P_1$ and $P_2^{-1}$ are forward horizontal paths, one of which is trivial. Then $Q'=P_1QP_2$, where $Q=Q_1Q_1^{-1}$, with $Q_1$ a forward path. The terminal point $p''$ of $Q_1$ is the \emph{apex} of $Q$, and $Q=Q_1Q_1^{-1}$ is an \emph{augmentation} of $\gamma$. The path $\gamma_{_{\succ}}$ lifts to a quasigeodesic $\widehat{\gamma_{_{\succ}}}\rightarrow {\widetilde X^\bullet}_L$. More specifically, each lift of the union of the vertical edges of $\gamma_{_{\succ}}$ determines a unique lift of $\gamma_{_{\succ}}$ to a quasigeodesic. Indeed, we can write $\gamma_{_{\succ}}$ in one of the following four forms: \begin{enumerate} \item $\cdots A_{-1}e_{-1}B_{-1}A_0e_0B_0A_1e_1B_1A_2e_2B_2\cdots$, where $e_{\pm i}$ are present for all $i\in\ensuremath{\field{N}}$; \item $A_0e_0B_0A_1\cdots$, where $A_0$ is unbounded; \item $\cdots A_0e_0B_0$, where $B_0$ is unbounded; \item $A_{s}e_sB_s\cdots A_{t}e_tB_t$ with $A_s,B_t$ unbounded. (This includes the case $B_0A_1$ in which $\gamma=\gamma_{_{\succ}}$ is horizontal.) \end{enumerate} Each $A_i$ starts at an apex and each $B_i$ ends at an apex. Observe that each lift of $e_i$ determines a lifts of $A_i$ and $B_i$ to $\widetilde X^{\bullet}_L$. Since the apexes lift uniquely, any lift of $B_i$ is concatenable with any lift of $A_{i+1}$, and we conclude that a lift of $\{e_i\}$ induces a lift of $\gamma_{_{\succ}}$. In the case where $\gamma$ is horizontal, $\gamma=\gamma_{_{\succ}}$ lifts uniquely since any horizontal path starting and ending in $\cup_{k}\widetilde V_{kL}$ lifts uniquely. Under the quasi-isometry $(\widetilde X^\bullet_L)^1\rightarrow\widetilde X^1$, the quasigeodesic $\widehat{\gamma_{_{\succ}}}$ is sent to $\gamma_{_{\succ}}$, and thus $\partial\widehat{\gamma_{_{\succ}}}=\partial\gamma$. Finally, if some augmentation of $\gamma$ has a subpath that lifts to a backtrack in $\widehat\gamma_{_{\succ}}$, then we truncate $\gamma_{_{\succ}}$ accordingly and define $\widehat\gamma_{_{\succ}}$ to be the lift of the truncated augmentation. An augmentation where this truncation is nontrivial is a \emph{truncated augmentation}. We call $\widehat{\gamma_{_{\succ}}}$ a \emph{lifted augmentation} of $\gamma$. \end{cons} \subsubsection{Cutting in ${\widetilde X^\bullet}_L$}\label{sec:general_cutting} We now establish a criterion, in terms of lifted augmentations, ensuring that a wall $\overline W$ cuts a given quasigeodesic in $\widetilde X^1$. We first require a classification of discrepancy zones. \begin{defn}[Exceptional zone, narrow exceptional zones]\label{defn:narrow_intermediate_zones} Let $W\rightarrow X$ be an immersed wall with tunnel-length $L$. An \emph{exceptional zone} is a downward discrepancy zone in $\widetilde X_L^{\bullet}$ whose boundary path intersects the interior of a slope approximation. The downward discrepancy zone shown in Figure~\ref{fig:downward_intermediate_zone} is exceptional. We say that $W$ has \emph{narrow exceptional zones} if for each exceptional zone $Z\subset\widetilde X^{\bullet}_L$, associated to a nucleus $\widetilde C$ of $\overline W_L$, the image in $Z\subset\widetilde X^\bullet_L$ of $\widetilde C\times[\frac{1}{2},\frac{3L}{4}]$ does not contain a vertex. \end{defn} \begin{lem}\label{lem:no_orange_orange} Suppose that $\phi:V\rightarrow V$ is a train track map with expanding edges. Suppose that $W\rightarrow X$ is an immersed wall such that every edge of $V$ contains a primary bust of $W$. Then if the tunnel length $L$ of $W$ is sufficiently large, each exceptional discrepancy zone $Z$ lies in the interior of a single long 2-cell of $\widetilde X^{\bullet}_L$, and hence $Z$ intersects a single slope-approximation. \end{lem} \begin{proof} Let $\comapp(S)$ be a slope-approximation in a long 2-cell $R$ based at the vertical edge $e\subset\widetilde V_n$. We will show that for $L$ sufficiently large, the nucleus $\widetilde C$ incident to $S$ is the copy in $\widetilde E_n$ of a subinterval of the interior of $e$. Let $\alpha$ be a component of $e-\interior{d}$, where $d$ is the primary bust associated to $S$. Then for all sufficiently large $L$, the path $\phi^L(\alpha)$ traverses an entire edge, and therefore contains a primary bust. \end{proof} \begin{lem}\label{lem:exceptional_no_vertex} Suppose that $\phi:V\rightarrow V$ is a train track map with exponentially expanding edges. Let $y_1,\ldots,y_s\in V$ be regular points such that each edge of $V$ contains exactly one $y_i$, and let $\epsilon>0$. Then for all sufficiently large $L$, there exists an immersed wall $W\rightarrow X$ with tunnel-length $L$, such that each primary bust is in the $\epsilon$-neighborhood of some $y_i$, and $W$ has narrow exceptional zones. \end{lem} \begin{proof} Let $\varpi>1$ be the expansion constant of $\phi$. For each $i$, let $y'_i\in V$ be a periodic regular point in the edge $e_i$ containing $y_i$ with $\textup{\textsf{d}}_{e_i}(y'_i,y_i)<\frac{\epsilon}{2}$. Let $\chi_i=\min\{\textup{\textsf{d}}_V(\phi^k(y'_i),V^0)\colon k\geq 0\},$ which is positive since $y'_i$ is periodic and regular. Let $\chi=\max_i\chi_i$. Let $$L>4(\log_{\varpi}\max_i|e_i|-\log_{\varpi}(\chi)).$$ For each $i$, let $d_i\subset\interior{e_i}$ be a primary bust chosen in the $\frac{\epsilon}{2\varpi^L}$-neighborhood of $y'_i$, and therefore in the $\epsilon$-neighborhood of $y_i$. Lemma~\ref{lem:choosing_busts_1} ensures that this can be done in such a way as to yield an immersed wall $W\rightarrow X$. \begin{figure}[ht] \begin{overpic}[scale=0.15]{exceptional_no_vertex} \put(30,18){$\comapp(S)$} \put(3,20){$e_i$} \put(23,24){$v$} \put(48,24.7){$e'$} \put(51, 24.5){$\comapp(\widetilde C)$} \end{overpic} \caption{The exceptional zone corresponding to $\comapp(S)$ cannot contain the vertex $v$ when $L$ is sufficiently large. Such a vertex $v$ could only be contained in a non-exceptional downward discrepancy zone, as shown at right.}\label{fig:exceptional_no_vertex} \end{figure} Let $Z$ be the image in $\widetilde X$ of an exceptional zone between $\overline W$ and $\comapp(\overline W)$. By Lemma~\ref{lem:no_orange_orange}, there is a unique slope $S$ such that the forward part of $\comapp(S)$ forms part of the boundary path of $Z$. See Figure~\ref{fig:exceptional_no_vertex}. If $v\in Z$ is a vertex at horizontal distince more than $\frac{L}{4}$ from the nucleus-approximation $\comapp(\widetilde C)$ on the right of $Z$, then our choice of $L$ would ensure that the right boundary path of $Z$ contains a complete edge $e'$, and thus a primary bust, which is impossible. \end{proof} \begin{prop}\label{prop:general_cutting} Let $\gamma:\mathbf R\rightarrow\widetilde X^1$ be an embedded quasigeodesic, and let $\widehat{\gamma_{_{\succ}}}\rightarrow{\widetilde X^\bullet}_L$ be a lifted augmentation. Let $C_o$ be a bounded subset of $\widehat{\gamma_{_{\succ}}}\cap\comapp(\overline W_L)$, let $C$ be the smallest subgraph containing $C_o$. Let $\widehat{\gamma_{_{\succ}}}\vee_C\comappn{\overline W_L}\rightarrow(\widetilde X^\bullet_L)^1$ be the graph obtained by wedging $\widehat{\gamma_{_{\succ}}}\rightarrow\widetilde X$ and $\comappn{\overline W_L}\rightarrow\widetilde X$ along the common subgraph $C$. Suppose that: \begin{enumerate} \item \label{item:wedgeqie}$\widehat{\gamma_{_{\succ}}}\vee_C\comappn{\overline W_L}\rightarrow(\widetilde X^\bullet_L)^1$ is a quasi-isometric embedding. \item \label{item:localsep}There are nontrivial intervals $f,f'\subset\widehat{\gamma_{_{\succ}}}$, immediately preceding and succeeding $C_o$ within $\widehat{\gamma_{_{\succ}}}$, that lie in $\overleftarrow A$ and $\overrightarrow A$ respectively. \item \label{item:nosurprise}For every component $D$ of $\widehat{\gamma_{_{\succ}}}\cap\comapp(\overline W_L)$ disjoint from $C_o$, the 1-neighborhood in $\widehat{\gamma_{_{\succ}}}$ of $D$ lies entirely in $\overleftarrow A$ or $\overrightarrow A$. \end{enumerate} Then $\overline W$ cuts $\gamma$. \end{prop} \begin{proof} Hypotheses~(2)~and~(3) together imply that $\widehat{\gamma_{_{\succ}}}$ decomposes as a concatenation $\overleftarrow\gamma\bar\gamma\overrightarrow\gamma$, where $\bar\gamma$ is a bounded path containing $C_o$ and $\overleftarrow\gamma,\overrightarrow\gamma$ are rays contained in $\overleftarrow A,\overrightarrow A$ respectively. The image of $\widehat{\gamma_{_{\succ}}}\vee_C\comappn{\overline W_L}\rightarrow{\widetilde X^\bullet}_L$ is $\widehat{\gamma_{_{\succ}}}\cup\comappn{\overline W_L}$, which is quasi-isometrically embedded in $({\widetilde X^\bullet}_L)^1$ by hypothesis~(1). The inclusion $\widehat{\gamma_{_{\succ}}}\cup\comappn{\overline W_L}\hookrightarrow({\widetilde X^\bullet}_L)^1$ thus induces an embedding $\partial\widehat{\gamma_{_{\succ}}}\sqcup\partial\comappn{\overline W_L}\rightarrow\partial{\widetilde X^\bullet}_L$. The two points of $\partial\widehat{\gamma_{_{\succ}}}$ are $\partial\overleftarrow\gamma\in\partial\overleftarrow A$ and $\partial\overrightarrow\gamma\in\partial\overrightarrow A$, and neither of these points lies in $\partial\comappn{\overline W_L}$ since hypothesis~(1) implies that no sub-ray of $\widehat{\gamma_{_{\succ}}}$ lies in a bounded neighborhood of $\comappn{\overline W_L}$. Applying the quasi-isometry ${\widetilde X^\bullet}_L\rightarrow\widetilde X$ shows that the points of $\partial\gamma\subset\partial\widetilde X$ lie in $\partial N(\overleftarrow{W})-\partial\overline W$ and $\partial N(\overrightarrow{W})-\partial\overline W$, whence $\overline W$ cuts $\gamma$. \end{proof} \subsection{Cutting deviating geodesics}\label{sec:deviating} \begin{prop}\label{prop:cutting_deviating} Let $\widetilde X$ satisfy the hypotheses of Proposition~\ref{prop:everything_cut}, let $M\geq0$, and let $\gamma:\mathbf R\rightarrow\widetilde X^1$ be an $M$-deviating geodesic. Then there exists a wall $\overline W$ such that $\comappn{\overline W}$ is quasiconvex in $\widetilde X^1$ and $\overline W$ cuts $\gamma$. \end{prop} \begin{proof} We will find a wall $\overline W\rightarrow\widetilde X$ satisfying the hypotheses of Proposition~\ref{prop:general_cutting}. \textbf{An oddly-intersecting forward path:} Since $\widetilde X$ is level-separated, there exists $z\in\widetilde X$ such that for all sufficiently large $n$, there is a regular level $\mathcal T_n=T_n^o(\tilde\phi^n(z))$ with a leaf at $z$, such that $\mathcal T_n$ has odd intersection with $\gamma$ and the distance in $\mathcal T_n$ from $\gamma\cap\mathcal T_n$ to the root or to any leaf of $\mathcal T_n$ exceeds $12(M+\delta)$. The fact that $\widetilde X$ has bounded level intersection and $\gamma$ is $M$-deviating implies that there exists $N$ and a finite, odd-cardinality set $C_o'\subset\gamma$ such that $\mathcal T_n\cap\gamma=C_o'$ for all $n\geq N$. Each $\mathcal T_n$ is the union of finitely many maximal forward paths emanating from leaves. For each $n\geq N$, we wish to choose a leaf $y$ of $\mathcal T_n$ such that the maximal forward path $\sigma_n\subset\mathcal T_n$ emanating from $y$ has the property that $\sigma_n\cap\gamma$ is a fixed odd-cardinality subset $C_o\subseteq C'_o$. However, to achieve this, we shall slightly modify $\gamma$ as follows by replacing it with an embedded deviating uniform quasigeodesic that coincides with the original $\gamma$ outside a diameter-$2M$ subset. We now describe the modification of $\gamma$. Let $e_1,\ldots,e_{|C_o'|}$ be the edges of $\gamma$ intersecting $\mathcal T_n$ for $n\geq N$. Index these so that $e_i$ precedes $e_j$ in the geodesic $\gamma$ if and only if $i<j$. The set $\{e_1,\ldots,e_{|C_o'|}\}$ is partially ordered as follows: $e_i\preceq e_j$ if for some $\ell\geq0$, the path $\tilde\phi^\ell(e_i)$ traverses $e_j$. The edges $e_i,e_j$ are \emph{confluent} if there exists $k$ such that $e_i,e_j\preceq e_k$. Confluence is an equivalence relation, and there is exactly one confluence class for each $\preceq$-maximal edge. Since $|C_o'|$ is odd, there exists an odd-cardinality confluence class, and we let $e_k$ be its $\preceq$-maximal element. Let $e_i,e_j$ be the elements of the confluence class of $e_k$ such that the indices $i,j$ are respectively minimal and maximal. Let $\alpha_i,\alpha_j$ be forward paths in $\mathcal T_n$ joining $e_i,e_j$ to $e_k$. Let $A_i$ be an embedded combinatorial path in the forward ladder $N(\alpha_i)$ that joins the terminal vertex $v_i$ of $e_i$ to a vertex $v_k$ of $e_k$ and does not intersect $\alpha_i$. The edge $e_j$ contains a vertex $v_j$ on the same side of $\mathcal T_n$ as the terminal vertex of $e_i$. Let $A_j$ be an embedded combinatorial path in $N(\alpha_j)$ joining $v_j$ to $v_k$ and not intersecting $\alpha_j$. Since $\gamma$ is deviating, $\textup{\textsf{d}}(e_i,e_k)$ and $\textup{\textsf{d}}(e_j,e_k)$ are uniformly bounded. Hence, since $N(\alpha_i)^1,N(\alpha_j)^1$ are uniformly quasiconvex, the paths $A_i,A_j$ have uniformly bounded length. Let $A$ be the path obtained from $A_iA_j^{-1}$ by removing backtracks. We replace the subpath of $\gamma$ between $v_i$ and $v_j$ by $A$, and finally replace $\gamma$ by a bi-infinite embedded in the image of this path. By construction, for all $n\geq N$, there is a forward path $\sigma_n$ of $\mathcal T_n$, that intersects the modified path exactly once, namely in a point of $e_i$. The argument proceeds using the new $\gamma$, which is an embedded quasigeodesic that is $M$-deviating, with $M$ a new, larger constant. However, since the quasi-isometry constants of $\gamma$ play no essential role in the argument, we will assume for simplicity that $\gamma$ remains a geodesic. There exists $\epsilon>0$ such that for all $x\in\mathcal N_{\epsilon}(y)$, any forward path $\sigma_x$ of length $n\geq N$ emanating from $x$ intersects $\gamma$ in a set $C^x_o$ of interior points of edges that has the same cardinality as $C_o$ and has the property that the smallest subcomplex $C$ containing $C_o$ is exactly the smallest subcomplex containing $C^x_o$. The wall we will choose will contain a slope $S$ such that $\comapp(S)$ contains such a $\sigma_x$ as its forward part. \textbf{Quasi-isometric embedding of $\gamma\vee_C\comappn{\overline W}\rightarrow\widetilde X^1$:} Let $W\rightarrow X$ be an immersed wall such that every edge of $V$ contains a primary bust, and suppose $\overline W\subset\widetilde X$ is the image of a lift $\widetilde W\rightarrow\widetilde X$ such that $\overline W$ contains a slope $S$ with the forward part of $\comapp(S)$ equal to a path $\sigma_x$, emanating from some $x\in\mathcal N_\epsilon(y)$, as above. Suppose moreover that $W$ was drawn from a set of immersed walls with uniformly bounded ladder-overlap. Since every edge contains a primary bust, Proposition~\ref{prop:quasiconvexity} provides constants $L_0,\kappa_1,\kappa_2$, depending only on $\widetilde X$, such that if the tunnel length of $W$ is at least $L_0$, then $\comappn{\overline W}$ is $(\kappa_1,\kappa_2)$ quasi-isometrically embedded. Recall also that $\overline W$ is a genuine wall if the tunnel-length exceeds a uniform constant $L_1$, by Proposition~\ref{prop:tunneliswall}. There exist constants $L_2\geq L_1,\kappa'_1,\kappa'_2$, depending on $\widetilde X$ and $M$ such that if $W$ has tunnel-length at least $L_2$, then $\gamma\vee_C\comappn{\overline W}\rightarrow\widetilde X^1$ is a $(\kappa_1',\kappa_2')$-quasi-isometric embedding. Indeed, this follows from an application of Lemma~\ref{lem:RBRquasi}, since $\gamma$ is $M$-deviating and hence has uniformly bounded $(3\delta+2\lambda)$-overlap with $\comapp(S)$. \textbf{Verification that $\gamma\vee_{C_o}\comapp(\overline W)$ embeds:} By construction, $\gamma$ does not intersect any point of $\comapp(S)$ outside of $C_o$. Hence suppose that $\tau\beta_1\alpha_1\cdots\beta_k\alpha_k\beta_{k+1}$ is a path in $\comappn{\overline W}\cup\gamma$ that begins and ends in $C_o$, such that: $\tau$ is a subpath of $\gamma$, and each $\beta_i$ lies in the carrier of a slope-approximation, and each $\alpha_i$ lies in a nucleus approximation, and $|\beta_i|\geq L$ except when $i=k+1$. If $L$ is sufficiently large and $|\beta_2|=L$, then the existence of such a closed path contradicts the above conclusion that $\comappn{\overline W}\vee_C\gamma$ uniformly quasi-isometrically embeds. The remaining possibility is that a path of the form $\tau\beta_1\alpha_1\beta_2$ or $\tau\beta_1\alpha_1$ is closed in $\widetilde X$. In either case, when $L$ is sufficiently large, a thin quadrilateral argument shows that $\gamma$ is forced to $(2\delta+\lambda)$-fellow-travel with $\beta_1$ or $\beta_2$ for distance exceeding $M$, since the fellow-traveling between $\alpha_1$ and $\beta_i$ is controlled by Lemma~\ref{lem:vert_horiz_bound}. Hence $\gamma\vee_{C_o}\comapp(\overline W)$ embeds in $\widetilde X$. \textbf{Preventing short backtracks from crossing $\comapp(\overline W_L)$ at an apex:} We now compute the tunnel-length $L_3\geq L_2$ necessary to ensure that each augmentation $QQ^{-1}$ in $\gamma_{_{\succ}}$ either fails to intersect $\comapp(\overline W)$ or has length at least $\frac{L}{4}$, where $L\geq L_3$ is the tunnel-length of $W$. Note that if $QQ^{-1}$ is a truncated augmentation in the sense of Construction~\ref{cons:LA}, then the apex lies in some $\widetilde V_{n}$ with $n\not\in L\ensuremath{\field{Z}}$, and hence $QQ^{-1}\cap\comapp(\overline W)=\emptyset$, so we only need consider non-truncated augmentations. Let $W$ have tunnel-length $L\geq L_2$ and let $QQ^{-1}$ be a augmentation whose apex $p$ lies in $\comapp(\overline W)$, and hence in a nucleus-approximation. Suppose that $|Q|\leq\frac{L}{4}$. Let $\gamma'$ be the subpath of $\gamma$ between $C$ and the initial point of $Q$, let $\beta$ be a geodesic of $\comappn{\overline W}$ joining $p$ to the terminal point of $\comapp(S)$, and let $\tau$ be a geodesic of $\comappn{S}$ joining the initial point of $\gamma'$ to the terminal point of $\beta$. Since $\gamma$ is deviating, the path $\gamma'Q$ is a quasigeodesic with constants depending only on $M$ and $\lambda$. Meanwhile, since $\overline W$ has uniform ladder-overlap and $L\geq L_2\geq L_0$, the path $\beta\tau^{-1}$ is a $(\kappa_1,\kappa_2)$-quasigeodesic. Hence $\gamma'Q$ fellow-travels with $\tau\beta^{-1}$ at distance depending only on $M$ and $\widetilde X$. This is impossible for sufficiently large $L$, since $\gamma',\tau$ have $(2\delta+2\lambda)$-overlap of length at most $M$. \textbf{Choosing $\overline W$:} Since $\widetilde X$ has many effective walls, there exists an immersed wall $W\rightarrow X$ with tunnel length $L\geq L_3$, involving a primary bust in every edge of $V$, such that the image $\overline W$ of a lift $\widetilde W\rightarrow\widetilde X$ satisfies the following: \begin{enumerate} \item $\overline W$ is a wall (since $L_3\geq L_2$). \item $\comapp(\overline W)$ is $(\kappa_1,\kappa_2)$-quasiconvex. \item $\gamma\cap\comapp(\overline W)=C$, which is contained in the carrier of a slope-approximation $\comapp(S)$. \item $\comappn{\overline W}\vee_C\gamma\rightarrow\widetilde X^1$ is a quasi-isometric embedding. \item Any augmentation $QQ^{-1}$ of $\gamma$ that intersects $\comapp(\overline W)$ has the property that $|Q|>\frac{L}{4}$ (since $L\geq L_3$). \end{enumerate} $W$ is chosen from the spreading set $\mathbb W$ given in Definition~\ref{defn:many_effective_walls}.(1). \textbf{An arbitrary lifted augmentation:} Let $\liftapp{\gamma}\rightarrow\widetilde X^{\bullet}_L$ be a lifted augmentation of $\gamma$. Since the map $\widetilde X^{\bullet}_L\rightarrow\widetilde X$ is a quasi-isometry and restricts to the identity on $\comapp(\overline W_L)$ and sends $\liftapp{\gamma}$ to $\gamma_{_{\succ}}$, the intersection $\widehat C=\liftapp{\gamma}\cap\comappn{\overline W_L}$ is bounded and $\liftapp{\gamma}\vee_{\widehat C}\comappn{\overline W_L}\rightarrow\widetilde X^{\bullet}_L$ is a quasi-isometric embedding. (We could have chosen a specific lifted augmentation to make $\widehat C\neq\emptyset$, but it follows from the discussion below that this holds for any lifted augmentation.) Thus any $\liftapp{\gamma}$, together with $\comapp(\overline W_L)$, satisfies Hypothesis~\eqref{item:wedgeqie} of Proposition~\ref{prop:general_cutting}. We now verify that $\liftapp{\gamma}$ satisfies the remaining two hypotheses of Proposition~\ref{prop:general_cutting}. To this end, let $\hat\eta$ be an embedded quasigeodesic in $\widetilde X^{\bullet}_L$ obtained from $\tilde\phi_L\circ\liftapp{\gamma}$ by removing backtracks, and let $\eta$ be the image in $\widetilde X$ of $\hat\eta$. Note that $\hat\eta$ is independent of the choice of lifted augmentation of $\gamma$. \textbf{Intersection of $\hat\eta$ with $\comapp(\overline W)$:} We first work in $\widetilde X$. Recall that $\gamma\cap\comapp(\overline W)$ is the odd-cardinality set $C_o$ of points in $\comapp(S)$ for some slope $S$ of $\overline W$. Consider the nucleus $\widetilde M\subset\overline W$ that intersects $S$ and $\comapp(S)$. Then, since we can assume that $L$ is sufficiently large to ensure that each primary bust is separated from each vertex by a secondary bust, $\widetilde M$ corresponds to a subinterval containing no vertex, and hence $\comapp(\widetilde M)$ maps to a subspace of the star of a vertex in $V$. By Lemma~\ref{lem:exceptional_no_vertex} and our above choice of $L$, the exceptional zone determined by $\widetilde M$ and $\comapp(\widetilde M)$ contains no vertex of a vertical edge containing a point of $C_o$. It follows that the tunnel $T'$ attached to the unique secondary bust of $\widetilde M$ intersects $\gamma$ in a set of points corresponding bijectively to $C_o$. Let $S'$ be the slope of $T'$. Then, since the primary busts can be chosen arbitrarily small, we can assume that $\comapp(S')\cap\eta$ is an odd-cardinality set $E_o'$. Hence $\comapp(S')\cap\hat\eta\subset\widetilde X^\bullet_L$ is an odd-cardinality set $E_o$ mapping bijectively to $E_o'$. Since the endpoints of primary busts are regular, the path $\hat\eta$ contains a nontrivial interval $I$ ending at a point of $E_o$ and lying in the image of $S'$ under the forward flow; hence $I\subset\overrightarrow{W}_L$ and, since $I$ does not lie in a discrepancy zone, $I\subset\overrightarrow A$. There is likewise a nontrivial interval $I'$ in $\hat\eta$ beginning on $E_o$ and lying in an exceptional zone determined by the nucleus intersecting $\comapp(S')$. Moreover, $I'$ and $I$ can be chosen to be separated in $\hat\eta$ by $E_o$. See Figure~\ref{fig:repair}. \begin{figure}[h] \begin{overpic}[width=0.6\textwidth]{repair.pdf} \put(12,4){$S$} \put(-5,3){$\overline W$} \put(3,11){$\widetilde M$} \put(34,4){$\comapp(S)$} \put(84,9){$\comapp(S')$} \put(30,22){$\gamma$} \put(77,22){$\eta$} \put(82,33){$p$} \put(50,35){$\upsilon$} \put(32,32){$p'$} \end{overpic} \caption{The relationship between $\gamma,\eta,\overline W,\comapp(\overline W)$ in $\widetilde X$. The intervals $I,I'$ in $\widetilde X^\bullet_L$ map to the bold intervals. The path $\theta$ contains the terminal part of $\comapp(S)$ and the initial part of $\comapp(S')$.}\label{fig:repair} \end{figure} We claim that $\comapp(\overline W)\cap\hat\eta=E_o$. Otherwise, applying the map $\widetilde X^\bullet_L\rightarrow\widetilde X$ would show that $\comapp(\overline W)\cap\eta$ contains some point $p\not\in E'_o$, since $\widetilde X^\bullet_L\rightarrow\widetilde X$ restricts to a bijection on $\comapp(\overline W)$. Then there is a forward path $\upsilon$ of length $L$ emanating from a point $p'\in\gamma$ and terminating at $p$. Let $\theta$ be a geodesic of $\comapp(\overline W)$ joining $p$ to a closest point $a$ of $C_o$, and let $\gamma_1$ be the subpath of $\gamma$ joining $a$ to $p'$. Then $\gamma_1\upsilon$ is a quasigeodesic with quasi-isometry constants depending only on the deviation constant of $\gamma$, while $\theta$ is a $(\kappa_1,\kappa_2)$-quasigeodesic. Hence $\gamma_1\upsilon$ fellowtravels with $\theta$ at distance depending only on $\delta,M,\kappa_1,\kappa_2$ and not on $L$. It follows that there is a uniform upper bound on $|\gamma_1|$ that is independent of $L$. Hence, if $L$ is sufficiently large, then since $C_o$ lies at distance at least $\frac{L}{4}$ from all nuclei, $\min_n|q(p)-nL|\geq\frac{L}{4}$, so that $p$ lies at horizontal distance at least $\frac{L}{4}$ from any nucleus approximation. Suppose that $\comapp(S')$ lies in $\theta$. Then $\theta$ contains a point at distance at least $\frac{L}{4}$ from $\gamma_1\upsilon$, and hence $\gamma_1\upsilon$ and $\theta$ cannot uniformly fellow-travel when $L$ is sufficiently large. Similarly, if $\theta$ enters some other slope-approximation attached to $\comapp(\widetilde M)$, we find that $\theta$ and $\gamma_1\upsilon$ cannot fellow-travel. The remaining possibility is that there is a path in $\comapp(\widetilde M)$ joining the endpoint of $\comapp(S)$ to a point of $\upsilon$. This is impossible since, by Remark~\ref{rem:nearly_fixed}, the level part of $T'$ has odd-cardinality intersection with $\gamma$ and $p\not\in E'_o$. \textbf{Conclusion:} It follows from the above discussion that $\hat\eta$ contains two quasigeodesic rays, one in each of the halfspaces of $\widetilde X^\bullet_L$ associated to $\comapp(\overline W)$. Since $\hat\eta$ fellow-travels with $\liftapp{\gamma}$, we see that $\liftapp{\gamma}$ satisfies all hypotheses of Proposition~\ref{prop:general_cutting}, whence $\overline W$ cuts $\gamma$. \end{proof} \iffalse \begin{figure}[ht] \includegraphics[width=0.75\textwidth]{five_cases2}\\ \caption{$\widehat{QQ^{-1}}$ is shown with arrows. The first picture is impossible. In every other case, $\widehat{QQ^{-1}}$ must remain in the left halfspace as it passes through the vertical part of $\comapp(\overline W_L)$.}\label{fig:five_cases} \end{figure} \fi \subsection{Cutting ladderlike geodesics}\label{subsec:ladderlike_case} \begin{prop}\label{prop:cutting_ladderlike} Suppose that $\widetilde X$ has many effective walls and for each bounded forward path $\alpha$ there exists a periodic regular forward path $\alpha'$ such that $N(\alpha)=N(\alpha')$. Then for each geodesic $\gamma:\mathbf R\rightarrow\widetilde X^1$ that is not $M$-deviating for any $M$, there exists an immersed wall $W\rightarrow X$ such that $\overline W$ is a wall, $\comappn{\overline W}$ is quasiconvex, and $\overline W$ cuts $\gamma$. \end{prop} \begin{proof} Suppose that $\gamma$ contains a path $\gamma'$ such that for some regular $x\in\widetilde X^1$ and some $M$ to be determined, the path $\gamma'$ fellowtravels at distance $(2\delta+2\lambda)$ with the sequence $x,\tilde\phi(x),\ldots,\tilde\phi^M(x)$, where $x$ is a periodic point. Such $\gamma'$ exists for arbitrarily large $M$ by combining the fact that $\gamma$ is $M$-ladderlike for arbitrarily large $M$ with the first hypothesis. We shall show that if $M$ is sufficiently large, then there exists a wall $\overline W$ that has the desired properties and cuts $\gamma$ and separates $x$ and $\tilde\phi^M(x)$. \textbf{Choosing $W$ using many effective walls:} Without loss of generality, $M$ is an even integer, and we let $a=\tilde\phi^{M/2}(x)$. Note that $a$ is periodic. Let $\{e_i\}$ be the collection of edges of $V$, and let $W\rightarrow X$ be an immersed wall busting each $e_i$, with tunnel-length $L$ to be determined. Let $e_1$ be the edge whose interior contains $a$. By Remark~\ref{rem:uniform_k} and the fact that $\widetilde X$ has many effective walls, there exist $\kappa_1,\kappa_2, L_1$ depending only on $\widetilde X$ such that we can choose $W$ with tunnel length $L\geq L_1$ so that $\overline W$ is a wall and $\comappn{\overline W}$ is $(\kappa_1,\kappa_2)$-quasi-isometrically embedded. Moreover, we choose $W$ from the the collection $\mathbb W_a$ of Definition~\ref{defn:many_effective_walls}.\eqref{item:w_a_periodic}, which guarantees that $\overline W$ can be chosen with the following properties: \begin{enumerate} \item There exists $k\geq0$ such that for each primary bust $d$ with an endpoint in $\overline W$ in the same knockout as $a$, we have $\textup{\textsf{d}}(\tilde\phi^n(a),\tilde\phi^n(d))\geq3\delta+2\lambda$ for all $n\geq k$.\label{eq:k} \item $W$ has tunnel length $L>\max\{12(\delta+k),L_1\}$, independent of $M$. \item The image of $a$ in $V$ lies in the interior of a nucleus of $W$ and so $\comapp(\overline W)$ contains $\tilde\phi^L(a)$. \end{enumerate} We assume that $M>JL$, where $J\geq 4$ will be chosen below. Let $\sigma$ be the uniform quasigeodesic in $\widetilde X^1$ obtained from $\gamma$ by removing $\gamma'$ and replacing it by the sequence $x,\ldots,\tilde\phi^M(x)$. \textbf{Verifying that $\sigma\vee_{\phi^L(a)}\comappn{\overline W}$ quasi-isometrically embeds:} Consider paths of the form $\alpha_0\beta_0\cdots\beta_{s-1}\alpha_s\tau$, where $\beta_i$ is a geodesic of the carrier of a slope-approximation, $\alpha_i$ is a vertical geodesic of $\comappn{\overline W}$, and $\alpha_s$ terminates at $\tilde\phi^L(a)$, and $\tau$ is a subpath of $\sigma$ beginning at $\tilde\phi^L(a)$ (actually, the initial part of $\tau$ is a subsequence of $x,\tilde\phi(x),\ldots\tilde\phi^{\frac{M}{2}+L}(x)$ or of $\tilde\phi^{\frac{M}{2}+L}(x),\ldots\tilde\phi^M(x)$). The $(3\delta+2\lambda)$-overlap between $\alpha_s$ and $\tau$ and between $\alpha_i$ and $\beta_{i}$ and between $\alpha_i$ and $\beta_{i-1}$ is controlled by Lemma~\ref{lem:vert_horiz_bound}, and Condition~\eqref{eq:k} on $\overline W$ ensures that the $(3\delta+2\lambda)$ overlap between $\tau$ and $\beta_{s-1}$ has length at most $k$. The choice of $L$ now allows us to invoke Lemma~\ref{lem:RBRquasi} to conclude that $\sigma\vee_{\tilde\phi^L(a)}\comappn{\overline W}$ is quasi-isometrically embedded in $\widetilde X^1$, with constants $(\kappa_1',\kappa_2')$ depending only on $\kappa_1,\kappa_2,\lambda$. \textbf{Verifying that $\sigma\vee_{\tilde\phi^L(a)}\comapp(\overline W)$ embeds:} We will show that there is no path $\tau\subset\sigma$ beginning at $\tilde\phi^L(a)$ and joining the endpoints of a path $\alpha_0\beta_0\cdots\alpha_{m}$ or $\alpha_0\beta_0\cdots\alpha_m\beta_m$ in $\comappn{\overline W}$ with each $\alpha_i$ vertical, and each $\beta_i$ a path in the carrier of a slope approximation. Each $\beta_i$ has length $L$ except for the $\beta_m$ in the path of the second form. Since $\sigma\vee_{\tilde\phi^L(a)}\comappn{\overline W}$ is quasi-isometrically embedded, it suffices to examine the case where $m\leq 1$. The quadrilateral $\alpha_0\beta_0\alpha_1\tau^{-1}$ is approximated by a quasigeodesic quadrilateral $\bar\alpha_0\beta_0\bar\alpha_1\tau^{-1}$, where each $\bar\alpha_i$ is a geodesic of length exceeding $3\delta+2\lambda$. This quadrilateral is $2\delta+2\lambda$ thin, and $\beta_0,\tau$ fellow travel at distance $2\delta+2\lambda$ for length at most $k$. Hence, without loss of generality, $\bar\alpha_0$ must fellow-travel with $\beta_0$ at distance $2\delta+\lambda$ for distance at least $\frac{11L}{24}$, whence $\alpha_0$ must $(2\delta+2\lambda+\mu)$-fellow-travel with $\beta_0$ for distance at least $\frac{11L}{24\mu_1}-\mu_2$, which contradicts Lemma~\ref{lem:vert_horiz_bound} when $L$ is sufficiently large. (Recall that the quasiconvexity constant $\mu$ and the quasi-isometry constants $(\mu_1,\mu_2)$ of the nucleus approximations are independent of $L$.) Hence $\sigma\vee_{\tilde\phi^L(a)}\comapp(\overline W)$ embeds. \textbf{Applying Proposition~\ref{prop:general_cutting}:} Let $\liftapp{\sigma}$ be a lifted augmentation of $\sigma$ induced by a lift of the forward path joining $x$ to $\phi^M(x)$. Let $\hat x_i$ denote the lift of $\tilde\phi^i(x)$, so that $\hat x_{M/2}$ is a lift of $a$ and $\hat x_{M/2+L}$ is a lift of $\tilde\phi^L(a)$. The wall $\overline W$ is the image of a wall $\overline W_L$ such that a nucleus of $\overline W_L$ separates $\hat x_{M/2}$ from $\hat x_{M/2+1}$ and thus $\hat x_{M/2+L}$ lies in a nucleus approximation of $\comapp(\overline W_L)$. (Recall that $\comapp(\overline W_L)$ maps isomorphically to $\comapp(\overline W)$.) Thus the points $\hat x_{M/2+L\pm1}$ lie in distinct halfspaces associated to $\comapp(\overline W_L)$. Indeed, $\hat x_{M/2+L-1}$ lies in a downward discrepancy zone and hence in $\overleftarrow A$, while $\hat x_{M/2+L+1}\in\overrightarrow A$. See Figure~\ref{fig:ladderlike_cut}. This verifies Hypothesis~\eqref{item:localsep} of Proposition~\ref{prop:general_cutting}. \begin{figure} \begin{overpic}[scale=0.25]{ladderlike_cut} \put(9.2,20){$\hat x_{_{0}}$} \put(60.3,20){$\hat x_{_{\frac{M}{2}+L}}$} \put(80.5,20){$\hat x_{_M}$} \put(42,33){$\overline W_L$} \put(61,33){$\comapp(\overline W_L)$} \end{overpic} \caption{Notation in the proof of Proposition~\ref{prop:cutting_ladderlike}.}\label{fig:ladderlike_cut} \end{figure} As in the proof of Proposition~\ref{prop:cutting_deviating}, the fact that $\sigma\vee_{\tilde\phi^L(a)}\comappn{\overline W}$ is quasi-isometrically embedded, together with the fact that $\widetilde X^{\bullet}_L\rightarrow\widetilde X$ is a quasi-isometry, shows that $\liftapp{\sigma}\vee_{\hat x_{M/2+L}}\comappn{\overline W_L}\rightarrow\widetilde X^{\bullet}_L$ is a quasi-isometric embedding. This verifies Hypothesis~\eqref{item:wedgeqie} of Proposition~\ref{prop:general_cutting}. Let $y\in\liftapp{\sigma}\cap\comapp(\overline W_L)$. Then either $y$ maps to a point of $\sigma\cap\comapp(\overline W)$, in which case $y=\tilde\phi^L(a)$ since $\sigma\vee_{\tilde\phi^L(a)}\comapp(\overline W)$ embeds in $\widetilde X$, or $y$ is an apex of $\liftapp{\sigma}$. The latter is impossible provided $J$ is sufficiently large compared to $\kappa_1',\kappa_2'$. Indeed, suppose $QQ^{-1}$ is an augmentation beginning on $\sigma$ and having an apex $p\in\comapp(\overline W)$. Let $\beta\rightarrow\comappn{\overline W}$ join $p$ to $\tilde\phi^L(a)$, let $\tau\rightarrow N(\sigma'')$ join $\tilde\phi^L(a)$ to $x$, and let $P$ be the subpath of $\sigma$ subtended by $x$ and the initial point of $Q$. Then the concatenation $P\tau^{-1}\beta^{-1}$ is a $(\kappa_1',\kappa_2')$-quasigeodesic containing a subpath of length at least $(J/2-1)L$, namely $\tau$. Hence if $J>2(\kappa_1'(L+\kappa_2')L^{-1}+1)$, then the offending apex $p$ cannot exist since $|Q|\leq L$. This verifies Hypothesis~\eqref{item:nosurprise} of Proposition~\ref{prop:general_cutting}, and the proof is complete. \end{proof} \section{Leaf-separation and many effective walls in the irreducible case}\label{sec:getting_lol} In this section, we describe conditions on $\phi$ ensuring that $\widetilde X^1$ satisfies the hypotheses of Proposition~\ref{prop:everything_cut}. \subsection{Leaves}\label{subsec:leaves_and_train_tracks} \begin{defn}[Leaf]\label{defn:leaf} Let $x,y\in\widetilde X$. Then $x,y$ are \emph{equivalent} if there exist forward paths $\sigma_x,\sigma_y$ such that $x\in\sigma_x,y\in\sigma_y$ and $\sigma_x\cap\sigma_y\neq\emptyset$. An equivalence class is a \emph{leaf}. We denote the leaf containing $x$ by $\mathcal L_x$. The leaf $\mathcal L_x$ is \emph{singular} if it contains a 0-cell, and otherwise $\mathcal L_x$ is \emph{regular}. \end{defn} Observe that $\mathcal L_x$ is $\tilde\phi$-invariant. Moreover, observe that $\mathcal L_x$ has a natural directed graph structure: vertices are points of $\mathcal L_x\cap\widetilde X^1$, and edges are midsegments. From Proposition~\ref{prop:leaf_properties}.\eqref{leaf_neighborhood} and Proposition~\ref{prop:properties_of_levels}, it follows that this subdivision makes $\mathcal L_x$ a directed tree in which each vertex has exactly one outgoing edge and finitely many incoming edges. \begin{prop}[Properties of leaves]\label{prop:leaf_properties} Leaves have the following properties: \begin{enumerate} \item If $\mathcal L_x$ is a regular leaf and $\phi$ is a train track map, then $\mathcal L_x$ has a neighborhood homeomorphic to $\mathcal L_x\times[-1,1]$ with $\mathcal L_x$ identified with $\mathcal L_x\times\{0\}$.\label{leaf_neighborhood} \item Each level is contained in a unique leaf, and $\mathcal L_x$ is an increasing union of levels.\label{leaf_level} \end{enumerate} \end{prop} \begin{proof} \textbf{Proof of~$\eqref{leaf_neighborhood}$:} This uses Lemma~\ref{lem:finite_intersection_leaf_edge} below. For each vertex $v_{\tilde e}=\mathcal L_x\cap\tilde e$ of $\mathcal L_x$, let $U_{\tilde e}$ be an open interval in $\tilde e$ about $v_{\tilde e}$. For each edge $f_{\tilde c}=\mathcal L_x\cap R_{\tilde c}$ of $\mathcal L_x$, with vertices at $v_{\tilde c}$ and $v_{\tilde d}$, let $U(f_{\tilde c})$ be the open trapezoid in $R_{\tilde c}$ joining $U_{\tilde c}$ to $U_{\tilde d}$. The desired open neighborhood of $\mathcal L_x$ is $\bigcup_{f_{\tilde c}\in\text{Edges}(\mathcal L_x)}U(f_{\tilde c})$, as shown in Figure~\ref{fig:product_neighborhood}. \begin{figure}[ht] \includegraphics[width=0.25\textwidth]{product_neighborhood} \caption{A product neighborhood of a regular leaf.}\label{fig:product_neighborhood} \end{figure} \textbf{Proof of~$\eqref{leaf_level}$:} This follows immediately from the definitions of levels and leaves. \end{proof} We denote by $\mathcal Y_0$ the set of leaves of $\widetilde X$ and define a surjection $\rho_0:\widetilde X\rightarrow\mathcal Y_0$ by $\rho_0(x)=\mathcal L_x$. \begin{lem}\label{lem:finite_intersection_leaf_edge} Let $\tilde e$ be a vertical edge of $\widetilde X$ and let $\mathcal L_x$ be a leaf. If $\phi$ is a train track map, then $|\tilde e\cap\mathcal L_x|\leq 1$. \end{lem} \begin{proof} When $\phi$ is a train track map, distinct points in each vertical edge $e$ lie on distinct leaves, i.e. the map $\rho_0:e\rightarrow\mathcal Y_0$ is injective. Note that in this case, each 2-cell of $\widetilde X$ is foliated by a family of distinct fibers of $\rho$, each of which is a midsegment. \end{proof} \subsection{Forward space in the train track case}\label{sec:forward_space_in_train_track_case} Suppose that $\phi$ is a train track map. We now describe an $\ensuremath{\field{R}}$-tree $\mathcal Y$ whose points are equivalence classes of leaves, and a $G$-action on $\mathcal Y$, and use this to establish that $\widetilde X$ is level-separated. This construction mimics the stable tree discussed in~\cite{BestvinaFeighnHandel_lamination}, although the underlying set is defined differently. The referee explains that it is a special case of a construction in~\cite{GJLL}. Let $\mathcal E=\ensuremath{\field{R}}[\text{Edges}(V)]$ and denote by $\vec e_i$ the basis element of $\mathcal E$ corresponding to $e_i$. Let $\mathfrak M:\mathcal E\rightarrow\mathcal E$ be the linear map whose matrix with respect to the basis $\{\vec e_i\}$ has $ij$-entry the number of times the path $\phi(e_i)$ traverses $e_j$, ignoring orientation. Note that this \emph{transition matrix}, which we also denote by $\mathfrak M$, is a nonnegative matrix. We further assume that $\mathfrak M$ is irreducible. Let $\varpi$ be the Perron-Frobenius eigenvalue of $\mathfrak M$. As shown in~\cite{BestvinaHandel}, $\varpi>1$ since $\phi$ is irreducible and has infinite order. Let $\mathbf v$ be a $\varpi$-eigenvector, all of whose entries are positive. For each $i$, let $c_i$ be the magnitude of the Perron projection of $\vec e_i$ onto $\ensuremath{\field{R}}[\mathbf v]$. As is made precise in Definition~\ref{defn:exponentially_expanding}, the map $\phi$ expands edges of $V$ by a factor of $\varpi$. We now choose an equivariant weighting of vertical edges in $\widetilde X$ by letting $|e_i|=c_i$ for each edge $e_i$ of $V$, letting each horizontal edge of $X$ have unit weight, and pulling back these weights to $\widetilde X$. This determines the metric $\textup{\textsf{d}}$ on $\widetilde X^1$. For each $e_i$ and each $n\in\ensuremath{\field{Z}}$, we define the \emph{scaled length} of a lift $\tilde e_i$ of $e_i$ to $\widetilde V_n$ to be $\varpi^{-n}|\tilde e_i|=\varpi^{-n}c_i$. Let $\textup{\textsf{d}}_n:\widetilde V_n\times\widetilde V_n\rightarrow[0,\infty]$ be the resulting path-metric. Given leaves $\mathcal L_x,\mathcal L_y$, with $x,y\in\widetilde V_k$ for some $k$, let $$\textup{\textsf{d}}_{\infty}(\mathcal L_x,\mathcal L_y)=\lim_{n\rightarrow\infty}\textup{\textsf{d}}_{n}(\tilde\phi^n(x),\tilde\phi^n(y)).$$ This limit exists and is finite because $\textup{\textsf{d}}_n(\tilde\phi^n(x),\tilde\phi^n(y))$ is non-increasing and bounded. Moreover, $\textup{\textsf{d}}_{\infty}(\mathcal L_x,\mathcal L_y)$ is well-defined since for other choices $x'\in\mathcal L_x\cap\widetilde V_{k'}$ and $y'\in\mathcal L_y\cap\widetilde V_{k'}$, for all but finitely many $n$, we have $\phi^{n'}(x')=\phi^n(x)$ and $\phi^{n'}(y')=\phi^n(y)$ for some $n'$. \begin{lem}\label{lem:metrizing} Let $\mathcal Y$ be the quotient of $\mathcal Y_0$ obtained by identifying points $\rho_0(x),\rho_0(y)$ for which $\textup{\textsf{d}}_{\infty}(\rho_0(x),\rho_0(y))=0$. Then the induced pseudometric $\textup{\textsf{d}}_{\infty}:\mathcal Y\rightarrow [0,\infty)$ is a metric. Let $\rho:\widetilde X\rightarrow\mathcal Y$ be the composition $\widetilde X\stackrel{\rho_0}{\longrightarrow}\mathcal Y_0\rightarrow\mathcal Y$. Then the restriction of $\rho$ to each vertical edge is an isometric embedding. \end{lem} \begin{proof} $\textup{\textsf{d}}_{\infty}$ is symmetric and satisfies the triangle inequality. Hence $\textup{\textsf{d}}_{\infty}:\mathcal Y\rightarrow[0,\infty)$ is a metric. Let $e_i$ be a vertical edge with endpoints $x,y$. Then the distance in $\widetilde V$ between the endpoints of $\tilde\phi^n(e)$ is $\varpi^n|e_i|$, whence $\textup{\textsf{d}}_{\infty}(\rho(x),\rho(y))=c_i$. Our assumption that $\tilde\phi$ has a constant-speed parametrization on each edge implies that the same equality holds for any subinterval of $e_i$. \end{proof} \begin{prop}\label{prop:R_tree_base case} Suppose that every edge of $V$ is expanding with respect to $\phi$. Then: \begin{enumerate} \item The map $\rho:\widetilde X\rightarrow\mathcal Y$ is continuous. \item $(\mathcal Y,\textup{\textsf{d}}_{\infty})$ is a 0-hyperbolic geodesic metric space, i.e. $\mathcal Y$ is an $\ensuremath{\field{R}}$-tree. \item $\mathcal Y$ admits a $G$-action by homeomorphisms with respect to which $\rho$ is $G$-equivariant. \item The restriction of the $G$-action on $\mathcal Y$ to $F$ is an action by isometries. \item The stabilizer in $F$ of $\rho(\tilde x)$ is trivial whenever $\tilde x$ is a lift to $\widetilde X$ of a periodic point in $V$. \end{enumerate} \end{prop} \begin{proof} \textbf{Continuity of $\rho$:} The restriction of $\rho$ to each vertical edge $e$ is continuous since it is an isometric embedding, and $\rho$ is continuous on each closed 2-cell since $\rho$ is constant on each midsegment and each 2-cell is therefore foliated by fibers of $\rho$ since $\phi$ is a train track map. Since $\widetilde X$ is locally finite, the pasting lemma implies that $\rho$ is continuous on $\widetilde X$. \textbf{$\ensuremath{\field{R}}$-tree:} Let $x,y\in\widetilde V_n$ and let $P\rightarrow\widetilde V_0$ be a path joining $x$ to $y$. Then since $\rho$ is continuous, $\rho(P)$ is a path joining $\rho(x)$ to $\rho(y)$, whence $\mathcal Y$ is path-connected. Since $\mathcal Y$ is a path-connected subspace of an asymptotic cone of the simplicial tree $(\widetilde V,\textup{\textsf{d}}_0)$, the space $\mathcal Y$ is an $\ensuremath{\field{R}}$-tree~\cite[Prop.~3.6]{KapovichLeeb_asymptotic_cone}. (The asymptotic cone in question is built using any non-principal ultrafilter on $\ensuremath{\field{N}}$, the observation point $(\tilde v,\tilde\phi(\tilde v),\ldots)$, and the scaled metrics $\textup{\textsf{d}}_n$ on $\widetilde V_0$.) \textbf{The $G$-action:} For $g\in G$ and $x\in\widetilde X$, let $g\rho(x)=\rho(gx)$. This defines an action since $G$ takes leaves in $\widetilde X$ to leaves. The action is by homeomorphisms since $\rho$ is continuous and $G$ acts by homeomorphisms on $\widetilde X$. \textbf{The $F$-action is isometric:} Let $x,y\in\widetilde X$. Since $F$ acts by isometries on each $\widetilde V_n$, for each $f\in F$, we have \begin{eqnarray*} \textup{\textsf{d}}_{\infty}(f\rho(x),f\rho(y))&=&\lim_n\textup{\textsf{d}}_n(\phi^n(fx),\phi^n(fy))\\ &=&\lim_n\textup{\textsf{d}}_{n}(\Phi^n(f)\phi^n(x),\Phi^n(f)\phi^n(y))=\lim_n\textup{\textsf{d}}_n(\phi^n(x),\phi^n(y))=\textup{\textsf{d}}_{\infty}(\rho(x),\rho(y)). \end{eqnarray*} \textbf{The $F$-action is free on periodic points:} Let $x\in V$ be a periodic point and let $\tilde x$ be a lift of $x$ to $\widetilde X$. By Corollary~\ref{cor:bh_separation}, either $\rho(\tilde x)\neq f\rho(\tilde x)$, and we are done, or the forward rays $\sigma_{\tilde x}$ and $f\sigma_{\tilde x}$ emanating from $\tilde x$ and $f\tilde x$ respectively lie at finite Hausdorff distance. It follows that the immersed vertical path $\widetilde P$ joining $\tilde x$ to $f\tilde x$ projects to an essential closed path $P\rightarrow V$, based at $x$, such that $\phi^k(P)$ is a periodic Nielsen path for some $k\geq 0$. This contradicts hyperbolicity of $G$. \end{proof} \begin{rem} When $\phi$ is a $\pi$-isomorphism, and $G$ is hyperbolic, the action of $F$ on $\mathcal Y$ can be shown to be free using Lemma~\ref{lem:splitting_lemma_tt} and the fact that there are no nontrivial periodic Nielsen paths. We expect that this is true for a general hyperbolic monomorphism, but a free action on the set of periodic points suffices for our purposes. \end{rem} \subsection{Level-separation in the train track case}\label{subsec:tt_cube} The purpose of this subsection is to prove Lemma~\ref{lem:tt_level_separated}. \begin{defn}[Transverse]\label{defn:transverse} Let $\mathcal T$ be an $\ensuremath{\field{R}}$-tree. The map $\theta:\mathbf R\rightarrow\mathcal T$ is \emph{transverse} to $y\in\mathcal T$ if for each $p\in\theta^{-1}(y)$, there exists $\epsilon>0$ such that $\theta((p-\epsilon,p))$ and $\theta((p,p+\epsilon))$ lie in distinct components of $\mathcal T-\{y\}$. Note that if $\theta$ is transverse to $y$, then $\theta^{-1}(y)$ is a discrete set. \end{defn} We denote by $\mathbf R^+$ a combinatorial sub-ray of the combinatorial line $\mathbf R$. \begin{lem}\label{lem:trichotomy_tt} Let $\mathcal T$ be an $\ensuremath{\field{R}}$-tree. Let $\mathcal T_0\subseteq\mathcal T$ have the property that $\mathcal T-\{y\}$ has two components for each $y\in\mathcal T_0$ and each open arc of $\mathcal T$ contains a point of $\mathcal T_0$. Let $\theta:\mathbf R\rightarrow\mathcal T$ or $\theta:\mathbf R^+\rightarrow\mathcal T$ be a continuous map. Suppose $\theta$ is transverse to every point in $\mathcal T_0$. Moreover, suppose that each edge $e$ of the domain of $\theta$ has connected intersection with the preimage of each point in $\mathcal T$. Then one of the following holds: \begin{enumerate} \item \label{item:odd}There exists a nontrivial arc $\alpha\subset\mathcal T$ such that $|\theta^{-1}(y)|$ is odd for all $y\in\alpha\cap\mathcal T_0$. \item \label{item:infinite_preimage}There exists a point $y\in\mathcal T$ with $|\theta^{-1}(y)|$ infinite. \item \label{item:bigpreimage_tt}For each $r\geq0$, there exists $y_r\in\mathcal T$ such that $\theta^{-1}(y_r)$ has diameter at least $r$. \end{enumerate} \end{lem} \begin{proof} For each $p\in\mathbf R$, we denote by $\bar p$ its image in $\mathcal T$ and by $|\theta^{-1}(x)|$ the number of components of the preimage of $x\in\mathcal T$ in $\mathbf R$. We now show that either $(3)$ holds or $\image(\theta)$ is locally compact. We first claim that either $(3)$ holds, or for each edge $e$ of $\mathbf R$, there are (uniformly) finitely many edges $f$ such that $\theta(f)\cap\theta(e)\neq\emptyset$. Indeed, if there are arbitrarily many such $f$, then for each $r\geq 0$, we can choose $f$ such that $\textup{\textsf{d}}_{\mathbf R}(e,f)>r$ but $\theta(e)\cap\theta(f)\neq\emptyset$, yielding~(3). Second, choose a point $p\in\mathcal T$. Our first claim shows that either~(3) holds or the set $\{e_i\}$ of edges with $p\in\theta(e_i)$ is finite. Assume the latter. Then for each $i$ we can choose $\epsilon_i>0$ such that the $\epsilon_i$-neighborhood of $p$ in $\theta(e_i)$ is disjoint from the image of each edge not in $\{e_i\}$. Let $\epsilon=\min_i\epsilon_i$. Then the $\epsilon$-neighborhood of $p$ in $\image(\theta)$ lies in $\cup_i\theta(e_i)$ and thus has compact closure. There exist sequences $\{a_i\}$ and $\{b_i\}$ in $\mathbf R=(-\infty,\infty)$ converging to $\infty$ and $-\infty$ respectively, whose images are sequences $\{\bar a_i\}$ and $\{\bar b_i\}$ that converge to points $\bar a_\infty$ and $\bar b_\infty$ in $\image(\theta)\cup\partial\image(\theta)$. Indeed, since $\image(\theta)$ is a locally compact $\ensuremath{\field{R}}$-tree, $\image(\theta)\cup\partial\image(\theta)$ is compact by~\cite[Exmp.~II.8.11.(5)]{BridsonHaefliger}. \iffalse If the set $\image(\theta)-\mathcal T_0$ has no limit point, then $\image(\theta)$ is homeomorphic to a finite valence simplicial tree, and hence there are convergent sequences $\{a_i\},\{b_i\}$ in $\mathbf R$ whose images in $\mathcal T$ converge, since the closure of the locally finite tree $\image(\theta)$ in $\mathcal T\cup\partial\mathcal T$ is compact. Suppose there is a convergent sequence of distinct points $\{\bar a_i\not\in\mathcal T_0\}$ in $\theta(\mathbf R^+)$. We claim that either conclusion~(3) holds, or $\{\bar a_i\}$ can be chosen so that $\{a_i\in\mathbf R\}$ converges to $+\infty$. Indeed, if $\{|\theta^{-1}(\bar a_i)|\}$ is not bounded then conclusion~$(3)$ holds, so assume $\{|\theta^{-1}(\bar a_i)|\}$ is bounded. If $\cup_i \theta^{-1}(\bar a_i)$ does not lie in any $[0,n]$ then we can choose $\{a_i\}$ with the desired property. Hence assume $\cup \theta^{-1}(\bar a_i)\subset [0,n]$ for some $n$. But then $\theta([0,n])$ is a finite tree containing infinitely many valence $>2$ points, which is impossible. Likewise, either conclusion~$(3)$ holds, or we can we choose $\{b_i\}$ converging to $-\infty$. \fi Suppose $\bar a_\infty \neq \bar b_\infty$. Let $\alpha$ be a nontrivial arc in the geodesic joining $\bar a_{\infty}$ and $\bar b_{\infty}$. Note that $\theta^{-1}(\bar c)$ has either odd or infinite cardinality for each $\bar c\in\alpha\cap\mathcal T_0$, since it must separate $a_i$ from $b_i$ for all but finitely many $i$. Hence either conclusion~$(1)$ or $(2)$ holds. Suppose $\bar a_\infty$ and $\bar b_\infty$ are equal to the same point $\bar p_\infty\not\in\image\theta$. Let $\bar o$ denote the image of the basepoint $o$ of $\mathbf R$. The intersections $\bar o\bar a_i \cap \bar o \bar p_\infty$ converge to the segment $\bar o \bar p_\infty$. The same holds for $\bar o \bar b_i$. We use this to choose a new pair of sequences $\{a_i'\}$ and $\{b_i'\}$ that still converge to $\pm\infty$, and with the additional property that $\bar a_i'=\bar b_i'$. We do this by choosing the image points far out in $\bar o \bar p_\infty$. We have thus found arbitrarily distant points in $\mathbf R$ with the same images, verifying conclusion~(3). In the remaining case, there exists $p\in\mathbf R$ such that $\bar p=\bar p_\infty$. Consider the restriction of $\theta$ to a ray $\mathbf R^+$ so that the initial vertex of $\mathbf R^+$ is $p$. Let $o$ be the vertex adjacent to $p$. Repeating the previous argument with $a_i=p$ and $b_i$ converging to $p_\infty$ verifies conclusion~(3). The case of the ray $\mathbf R^+$ is similar. \end{proof} By Lemma~\ref{lem:finite_intersection_leaf_edge} and Proposition~\ref{prop:leaf_properties}, for each regular leaf there is a pair $(\overleftarrow{\mathcal L},\overrightarrow{\mathcal L})$ of closed halfspaces in $\widetilde X$ such that $\overleftarrow{\mathcal L}\cup\overrightarrow{\mathcal L}=\widetilde X$ and $\overleftarrow{\mathcal L}\cap\overrightarrow{\mathcal L}=\mathcal L$. Points of $\rho(\widetilde X^0)$ are \emph{singular} points of $\mathcal Y$, and the other points are \emph{regular}. If $\mathcal L$ is a regular leaf, then $\mathcal Y-\rho(\mathcal L)$ has two components, namely the interiors of the images of $\overleftarrow{\mathcal L}$ and $\overrightarrow{\mathcal L}$. Since there are countably many singular points in $\mathcal Y$, each open arc in $\mathcal Y$ contains a regular point. \begin{lem}\label{lem:rho_transverse} For any geodesic $\gamma:\mathbf R\rightarrow\widetilde X^1$, the map $\theta=\rho\circ\gamma:\mathbf R\rightarrow\mathcal Y$ is transverse to regular points. \end{lem} \begin{proof} Let $y\in\mathcal Y$ be a regular point, so that each $x\in\rho^{-1}(y)$ lies in the interior of a vertical 1-cell, which in turn embeds in $\mathcal Y$ by Proposition~\ref{prop:leaf_properties}. The image of the vertical 1-cell is separated by $\rho(x)=y$. \end{proof} The goal of the rest of this subsection is to prove Corollary~\ref{cor:odd_intersection}, which depends on Corollary~\ref{cor:bh_separation}. We first give a proof of the latter in the case where $\phi$ is $\pi_1$-surjective incorporating the technology of~\cite{BestvinaFeighnHandelTits}, followed by a self-contained proof in the general case. \begin{cor}\label{cor:odd_intersection} Let $\gamma:\mathbf R\rightarrow\widetilde X^1$ be an $M$-deviating geodesic for some $M\geq 0$. Then there exists a regular leaf $\mathcal L$ such that $|\gamma\cap\mathcal L|$ is finite and odd. \end{cor} \begin{proof Consider the restriction of $\rho$ to $\gamma$. By Lemma~\ref{lem:rho_transverse}, $\rho|_{\gamma}$ is transverse to regular points. By Lemma~\ref{lem:trichotomy_tt}, one of the following holds: \begin{itemize} \item There exists a regular point $y\in\mathcal Y$ such that $\rho^{-1}(y)\cap\gamma$ has finite, odd cardinality. \item For all $r\geq 0$, there exists $y_r\in\mathcal Y$ such that $\diam(\rho^{-1}(y_r)\cap\gamma)>r$. (This includes the case in which some point in $\mathcal Y$ has infinite preimage.) \end{itemize} In the first case, note that $\rho^{-1}(y)$ is the union of regular leaves, one of which must therefore have odd intersection with $\gamma$. We will now show that the second case leads to a contradiction. In the second case, for each $r\geq 0$, there exists $m\in\ensuremath{\field{Z}}$ and forward rays $\sigma_1,\sigma_2$, originating at points of $\gamma$ and traveling through $\widetilde V_m$, such that $\rho(\sigma_1)=\rho(\sigma_2)$ and $\textup{\textsf{d}}_{\widetilde V_m}(\sigma_1\cap\widetilde V_m,\sigma_2\cap\widetilde V_m)>r$. Indeed, let $x_1,x_2\in\gamma$ be chosen so that $\rho(x_1)=\rho(x_2)=y_r$ and $q(x_1)\leq q(x_2)=m$ and $\textup{\textsf{d}}_{\widetilde X}(x_1,x_2)>r+M$. For some $k\geq 0$, we have $\tilde\phi^k(x_1)=x'_1\in\widetilde V_m$. We also have $\rho(x'_1)=y_r$. Since $\gamma$ is $M$-deviating, considering the $\delta$-thin triangle $x_1x_2x'_1$ shows that $\textup{\textsf{d}}_{\widetilde X}(x'_1,x_2)>r$. Hence $\textup{\textsf{d}}_{\widetilde V_m}(x'_1,x_2)>r$. We now apply Corollary~\ref{cor:bh_separation}. The rays $\sigma_1,\sigma_2$ cannot fellowtravel when $r$ is sufficiently large, since the conclusion of a thin quadrilateral argument would then contradict the hypothesis that $\gamma$ is $M$-deviating. Hence, by Corollary~\ref{cor:bh_separation}, we see that $\rho(\sigma_1)\neq\rho(\sigma_2)$, a contradiction. \end{proof} The \emph{tightening} of a path $P$ in a graph is the immersed path that is path-homotopic to $P$. A \emph{periodic Nielsen path} in $V$ is an essential path $P$ such that the tightening of $\phi^k(P)$ is path-homotopic to $P$ for some $k>0$. The following is a rephrasing of a special case of~\cite[Lem.~6.5]{Levitt:split}, which splits into ~\cite[Lem. 4.1.4, Lem. 4.2.6, Lem. 5.5.1]{BestvinaFeighnHandelTits}. \begin{lem}[Splitting lemma]\label{lem:splitting_lemma_tt} Let $\phi:V\rightarrow V$ be a $\pi_1$-surjective train track map. Let $P\rightarrow V$ be a path. Then there exists $n_0$ such that the tightening of $\phi^{n_0}(P)$ is a concatenation $Q_1\cdots Q_k$, where each $Q_s$ is of one of the following types: \begin{enumerate} \item a periodic Nielsen path; \item an edge of $V$; \item a subinterval of an edge of $V$, if $s\in\{1,k\}$; \end{enumerate} Moreover, for all $n\geq n_0$, the tightening of $\phi^n(P)$ is equal to a concatenation of the tightenings of the paths $\phi^{n-n_0}(Q_s)$. \end{lem} \begin{cor}\label{cor:bh_separation} Let $\sigma_1,\sigma_2$ be forward rays beginning on $\widetilde V_m$. Then either $N(\sigma_1),N(\sigma_2)$ lie at finite Hausdorff distance or $\rho(\sigma_1)\neq\rho(\sigma_2)$. \end{cor} Corollary~\ref{cor:bh_separation} means that for each $y\in\mathcal Y$, any two forward rays in $\rho^{-1}(y)$ fellowtravel, in the sense that they lie at finite Hausdorff distance. \begin{proof}[Proof of Corollary~\ref{cor:bh_separation} when $\phi$ is $\pi_1$-surjective] Let $P\rightarrow\widetilde V_m$ be a path from $\sigma_1$ to $\sigma_2$. Lemma~\ref{lem:splitting_lemma_tt} implies that for some $n\geq0$, the tightening of $\tilde\phi^n(P)$ splits as the concatenation of Nielsen paths and edges. If $\tilde\phi^n(P)$ is the concatenation of Nielsen paths, then $\sigma_1,\sigma_2$ fellowtravel. Otherwise the splitting contains an edge $e$ and for all $n'\geq n$, we have $\textup{\textsf{d}}_{n'}(\sigma_1\cap\widetilde V_{n'},\sigma_2\cap\widetilde V_{n'})\geq |e|$, whence $\rho(\sigma_1)\neq\rho(\sigma_2)$. \end{proof} \begin{proof}[Proof~of~Corollary~\ref{cor:bh_separation} in the general case] If $\sigma_1,\sigma_2$ do not fellowtravel, then by Lemma~\ref{lem:separated_or_fellow_travel} and Lemma~\ref{lem:separated_implies_band}, the geodesic of $\widetilde V_m$ joining the initial points of $\sigma_1,\sigma_2$ contains an open arc $\alpha\subset e$, for some edge $e$, such that each regular leaf intersecting $\alpha$ separates $\sigma_1,\sigma_2$. For each $n\geq m$, let $a_n=\sigma_1\cap\widetilde V_n$ and $b_n=\sigma_2\cap\widetilde V_n$. Then for each $n$, the geodesic of $\widetilde V_n$ joining $a_n,b_n$ contains $\phi^n(\alpha)$. Regarding $e$ as a copy of $[0,1]$ with weight $|e|$, and $\alpha=(t_1,t_2)\subset[0,1]$, we see that $\textup{\textsf{d}}_n(a_n,b_n)\geq|e|(t_2-t_1).$ Hence $\textup{\textsf{d}}_{\infty}(\rho(\sigma_1),\rho(\sigma_2))>0$. \end{proof} \begin{lem}\label{lem:separated_or_fellow_travel} Let $\sigma_1,\sigma_2$ be forward rays beginning on $\widetilde V_m$. Then either $N(\sigma_1),N(\sigma_2)$ are at finite Hausdorff distance or there exists a regular leaf separating $\sigma_1$ from $\sigma_2$. \end{lem} \begin{proof} We claim that if $\sigma_1,\sigma_2$ are not separated by a regular leaf, then $\rho(\sigma_1)=\rho(\sigma_2)$. If $\rho(\sigma_1)\neq\rho(\sigma_2)$, then these points are separated by a point $y\in\mathcal Y$ whose preimage is the union of regular leaves. Any path joining $\sigma_1,\sigma_2$ must intersect the union of these leaves in an odd-cardinality set, so one of them separates $\sigma_1,\sigma_2$. Hence suppose $\rho(\sigma_1)=\rho(\sigma_2)$ and let $z=\rho(\sigma_i)\in \mathcal Y$. Let $p\geq m$ be such that there is a vertical geodesic $I_{p}\rightarrow\widetilde V_{p}$ joining $\sigma_1\cap\widetilde V_p,\sigma_2\cap\widetilde V_p$ and with the property that $\rho^{-1}(z)\cap I_p$ has minimal cardinality. For simplicity, having chosen $p$, we will translate so that $p=0$. Having chosen $I_0$, we now inductively define paths $I_n\rightarrow\widetilde V_n$ joining $\sigma_1$ to $\sigma_2$ as follows. For $n\geq 0$, express $I_n=e_1e_2\cdots e_k$ as a concatenation of partial edges: $e_1,e_k$ are closed subintervals of edges and the other $e_i$ are entire edges. Let $I_{n+1}\rightarrow\widetilde V_{n+1}$ be the path $\tilde\phi(e_1)\cdots\tilde\phi(e_k)$. Let $\bar I_n$ be the image of $I_n$ in $\widetilde X$ and note that $\bar I_n$ is a finite subtree of $\widetilde V_n$. Observe that $T=\rho(\bar I_0)\subset\mathcal Y$ is a finite tree, since it is the union of finitely many closed embedded arcs. Let $\rho_n:I_n\rightarrow\mathcal Y$ be the composition $I_n\rightarrow\widetilde X\stackrel{\rho}{\rightarrow}\mathcal Y$. Since each $\bar I_n\rightarrow\bar I_{n+1}$ is surjective, $\rho(\bar I_n)=T$ for all $n\geq 0$. The maps $e_i\rightarrow\tilde\phi(e_i)$ induce a map $I_n\rightarrow I_{n+1}$ so that the following diagram commutes. \begin{center} $ \begin{diagram} \node{I_n}\arrow{e}{}\arrow{s}{}\node{I_{n+1}}\arrow{s}{}\arrow{se,l}{\rho_{n+1}}\\ \node{\bar I_n}\arrow{e,t}{\tilde\phi}\node{\bar I_{n+1}}\arrow{e,t}{\rho}\node{T\subset\mathcal Y} \end{diagram} $ \end{center} Since $\rho(\sigma_1)=\rho(\sigma_2)$, each $\rho_n:I_n\rightarrow T$ is a closed path in $T$ beginning and ending at $z\in T$. If $d_{\widetilde V_n}(\sigma_1\cap\bar I_n,\sigma_2\cap\bar I_n)$ is uniformly bounded as $n\rightarrow\infty$, then $\sigma_1,\sigma_2$ lie at uniformly bounded vertical distance, and so $N(\sigma_1)$ and $N(\sigma_2)$ lie at finite Hausdorff distance. Since $I_n$ is vertical, $\rho_n^{-1}(z)$ is finite, and $I_n=Q_1Q_2\cdots Q_r$, where the interiors of the $Q_i$ are the components of $I_n-\rho_n^{-1}(z)$. Let $\bar Q_i$ denote the image of $Q_i$ in $\widetilde V_n$. Note that $r$ is independent of $n$; indeed, this is ensured by the minimality achieved through our choice of $p$. It follows that no regular leaf intersects $\bar Q_i$ and $\bar Q_j$ for $i\neq j$, for otherwise we could apply $\tilde\phi$ finitely many times and reduce $r$. Let $a_i$ and $b_i$ be the endpoints of $Q_i$, and let $\bar a_i,\bar b_i$ be their images in $\bar I_n$. We will show that there exists $M$, independent of $n$, such that $d_{\widetilde V_n}(\bar a_i,\bar b_i)\leq M$. We conclude that $d_{\widetilde V_n}(N(\sigma_1),N(\sigma_2))\leq rM$ for all sufficiently large $n$. To verify the existence of $M$, we shall show that there exists a leaf $\mathcal L_i$ that intersects the initial and terminal (possibly partial) edges of $\bar Q_i$, intersecting these edges in points $c_i,d_i$ respectively. This leaf $\mathcal L_i$ must intersect $\bar I_0$ in points $\hat c_i,\hat d_i$ with $\tilde\phi^n(\hat c_i)=c_i$ and $\tilde\phi^n(\hat d_i)=d_i$. Hence there are forward paths $\hat c_ic_i$ and $\hat d_id_i$ of $N(\mathcal L_i)$ whose intersections with $\widetilde X^1$ are $\lambda$-quasigeodesics lying in forward rays of $N(\mathcal L_i)$. The quasigeodesic quadrilateral $\hat c_ic_id_i\hat d_i$ shows that $\hat c_ic_i$ and $\hat d_id_i$ fellowtravel at distance $M'=M'(\delta,\lambda,|I_0|)$, and hence $d_{\widetilde V_n}(\bar a_i,\bar b_i)\leq M$, where $M=M'+2$. It remains to find the leaf $\mathcal L_i$. Note that if $\bar a_i,\bar b_i$ lie on a common leaf, we are done. We can assume that no regular leaf separates $\bar a_i$ from $\bar b_i$. Indeed, any such leaf could not separate $\sigma_1,\sigma_2$ by our assumption, and thus must end on $\bar Q_j$ for some $i\neq j$, which was ruled out above. Hence each leaf emanating from the image of the initial edge of $\bar Q_i$ intersects the images of an even number of edges of $\bar Q_i$. Let $\mathcal L$ be such a regular leaf, and let $b'_i\in\mathcal L\cap\bar Q_i$ be a point outside of the image of the initial partial edge of $\bar Q_i$. We claim that by choosing $\mathcal L$ to intersect the $\bar Q_i$ at a point $a'_i$ sufficiently close to $\bar a_i$, we can ensure that $b'_i$ lies in the image of the terminal partial edge of $\bar Q_i$. Indeed, choose a sequence $\{a'_{ik}\}_k$ of regular points in the initial edge of $\overline Q_i$, with $a'_{ik}\rightarrow\bar a_i$. For each $k$. let $\mathcal L_{ik}$ be the regular leaf containing $a'_{ik}$. If $\mathcal L_{ik}$ intersects the terminal edge of $\overline Q_i$, we are done, so we let $b'_{ik}$ be a point of $\mathcal L_{ik}\cap\overline Q_{ik}$ that lies in a non-terminal, non-initial edge. By possibly passing to a subsequence, compactness allows us to assume that $\{b'_{ik}\}$ converges to some $b_i'\in\overline Q_i$ different from $\bar b_i$. Since $\rho$ is continuous and $\rho(b'_{ik})=\rho(a'_{ik})\rightarrow z$, we have $\rho(b_i')=z$, contradicting the fact that the interior of $\overline Q_i$ contains no point in $\rho^{-1}(z)$. \end{proof} \begin{lem}\label{lem:separated_implies_band} Let $\sigma_1,\sigma_2$ be forward rays beginning on $\widetilde V_m$ that do not fellowtravel. Suppose there exists a regular leaf $\mathcal L$ separating $\sigma_1,\sigma_2$. Then the geodesic of $\widetilde V_m$ joining the initial points of $\sigma_1,\sigma_2$ contains an open arc $\alpha\subset e$, for some edge $e$, such that each regular leaf intersecting $\alpha$ separates $\sigma_1,\sigma_2$. Hence for all $n\geq m$, the geodesic of $\widetilde V_n$ joining $\sigma_1\cap\widetilde V_n$ and $\sigma_2\cap\widetilde V_n$ contains $\tilde\phi^n(\alpha)$. \end{lem} \begin{proof} Let $P\rightarrow\widetilde V_m$ be a vertical geodesic joining $\sigma_1,\sigma_2$. For any $n\geq m$, given a path $U\rightarrow\widetilde V_n$ joining $\sigma_1,\sigma_2$, a \emph{syllable} of $U$ is a maximal subpath $Q$ that is \emph{legal} in the sense of~\cite{BestvinaHandel}; since $\phi$ is a train track map, this means that $\tilde\phi^k(Q)$ is embedded for all $k\geq 0$. Consider the vertical geodesic $T_k$ joining the endpoints of $\tilde\phi^k(P)$ for $k\geq 0$. Each $T_k$ can be expressed as a concatenation of syllables, since $\phi$ is a train track map, and this decomposition is unique. Choose $k\geq 0$ such that the number of syllables in the decomposition of $T_k$ is equal to the number of syllables in $T_{k'}$ for all $k'\geq k$. Let $T_k=Q_1\cdots Q_n$ be a decomposition into syllables. Observe that nonconsecutive syllables of $T_k$ cannot intersect a common leaf, for otherwise applying some iterate of $\tilde\phi$ would result in a path with fewer syllables. Since $\mathcal L$ intersects each syllable in at most one point, the minimality of $T_k$ guarantees that $|\mathcal L\cap T_k|<3$ and hence, since this cardinality is odd, $|\mathcal L\cap T_k|=1$. Hence there exists a unique $Q_i$ such that $\mathcal L\cap T_k$ is contained in $\interior{Q_i}$. For each $p\in\ensuremath{\field{N}}$, let $B_i^{\pm}(\frac{1}{p})$ be the two half-open $\frac{1}{p}$-neighborhoods in $Q_i$ bounded at $\mathcal L\cap Q_i$. If the lemma does not hold, then for each $p$ there exists a regular leaf $\mathcal L^{\pm}(\frac{1}{p})$ intersecting $B_i^{\pm}(\frac{1}{p})$ but failing to separate $\sigma_1$ and $\sigma_2$. Each $\mathcal L^{\pm}(\frac{1}{p})$ has even intersection with $T_k$ and thus also intersects $Q_{i\pm 1}$ in a single point. The sequence $\{\mathcal L^{\pm}(\frac{1}{p})\cap Q_{i\pm1}\}_p$ has a subsequence that converges to a point $z_{\pm}\in Q_{i\pm1}$ such that $\rho(z_{\pm})=\rho(\mathcal L)$. Observe that no regular leaf separates $z_{\pm}$ from $\mathcal L\cap Q_i$, since such a separating regular leaf would have to intersect some $\mathcal L^{\pm}(\frac{1}{p})$, which is impossible since leaves are disjoint. Hence, by Lemma~\ref{lem:separated_or_fellow_travel}, the forward rays $\sigma_{z\pm}$ and $\sigma$, respectively emanating from $z_{\pm}$ and $\mathcal L\cap Q_i$, must fellowtravel. \begin{figure}[ht] \includegraphics[width=0.5\textwidth]{sigma_ft} \caption{The forward rays and leaves in the proof of Lemma~\ref{lem:separated_implies_band}}\label{fig:sigma_ft}. \end{figure} If $\sigma$ fellowtravels with $\sigma_1$ [resp. $\sigma_2$] and $\sigma_{z^{\pm}}$ fellowtravels with $\sigma_2$ [resp. $\sigma_1]$, then since $\sigma,\sigma_{z^{\pm}}$ fellowtravel, we conclude that $\sigma_1,\sigma_2$ fellowtravel, contradicting our hypotheses. See Figure~\ref{fig:sigma_ft}. If $\sigma,\sigma_1$ (for example) do not fellowtravel, then $\sigma_1,\sigma_{z^+}$ (say) also do not fellowtravel. Lemma~\ref{lem:separated_or_fellow_travel} implies that a regular leaf $\mathcal L_1$ separates $\sigma_{z^+},\sigma_1$. The part of $T_k$ subtended by $\sigma_{z^+},\sigma_1$ has strictly fewer syllables than $T_k$, so by induction, there is an open interval $\alpha'\subset T_k$ with the following properties: \begin{enumerate} \item $\alpha'$ is contained in the interior of some edge. \item $\alpha'$ intersects a regular leaf $\mathcal L'_1$ that separates $\sigma_{z^+}$ and $\sigma_1$. \item $\alpha'$ lies on the part of $T_k$ between $\sigma_{z^+}$ and $\sigma_1$. \item All regular leaves intersecting $\alpha'$ separate $\sigma_{z^+}$ from $\sigma_1$. \end{enumerate} Let $\mathcal L_2$ be a regular leaf intersecting $\alpha'$ between $\mathcal L_1'$ and $\sigma_1$. Then $\mathcal L_2$ separates $\sigma_1$ from $\sigma_{z^+}$ by the induction hypothesis, and therefore $\mathcal L_2$ separates $\sigma_1$ from $\sigma_2$. The subinterval of $\alpha'$ above $\mathcal L'_1$ (and so containing all such $\mathcal L_2$) is the desired interval $\alpha$. See Figure~\ref{fig:sigma_ft}. In the base case, $T_k$ has a single syllable, and any open subinterval of an edge suffices. Finally, let $n\geq m$ and let $P_n\rightarrow\widetilde V_n$ be the geodesic joining $\sigma_1,\sigma_2$. For any $x\in\tilde\phi^n(\alpha)$, and any $\epsilon>0$, there exists a regular point $y\in\tilde\phi^n(\alpha)$ at distance less than $\epsilon$ from $x$, since edges are expanding. The regular leaf $\mathcal L_y$ separates $\sigma_1,\sigma_2$, so that $y\in P_n$. Since this holds for arbitrarily small $\epsilon$ and $P_n$ is closed, $x\in P_n$. \end{proof} We have now arrived at the main goal of this subsection: \begin{lem}\label{lem:tt_level_separated} Suppose that $\phi$ is a train track map, that every edge of $V$ is expanding, and that $\mathfrak M$ is irreducible. Then $\widetilde X$ is level-separated. \end{lem} \begin{proof} Let $\gamma:\mathbf R\rightarrow\widetilde X^1$ be an $M$-deviating geodesic and let $K\geq0$. By Corollary~\ref{cor:odd_intersection}, there is a regular leaf $\mathcal L$ such that $|\mathcal L\cap\gamma|$ is finite and odd. Let $C_0=\mathcal L\cap\gamma$ and choose $y\in\mathcal L$ such that $q(c)-q(y)>M+K$ for all $c\in C_0$. Then for all sufficiently large $n$, there is a level $T^n_o(\tilde\phi^n(y))\subset\mathcal L$ that contains $y$ as one of its leaves and satisfies $T^n_o(\tilde\phi^n(y))\cap\gamma=C_0$. Hence $\widetilde X$ is level-separated. \end{proof} \subsection{Proof of Theorem~\ref{thmi:general}}\label{sec:irreducible_case} \begin{thm}\label{thm:irreducible} Let $\phi:V\rightarrow V$ be a train track map of a finite graph $V$. Suppose that $\phi$ is $\pi_1$-injective and that each edge of $V$ is expanding. Moreover, suppose that the transition matrix $\mathfrak M$ of $\phi$ is irreducible and that the mapping torus $X$ of $\phi$ has word-hyperbolic fundamental group $G$. Then $G$ acts freely and cocompactly on a CAT(0) cube complex. \end{thm} \begin{proof} Let $\mathcal Y$ be the forward space arising from the map $\tilde\phi:\widetilde X\rightarrow\widetilde X$. Since $\phi$ is a train track map, $\widetilde X$ has bounded level intersection by Remark~\ref{rem:bounded_level_intersection} and is level-separated by Lemma~\ref{lem:tt_level_separated}. By Lemma~\ref{lem:periodic_paths_everywhere}, each finite forward path uniformly fellow-travels with a periodic forward path. Hence by Proposition~\ref{prop:everything_cut}, it suffices to show that $\widetilde X$ has many effective walls by verifying Conditions~\eqref{item:bust_point} and~\eqref{item:w_a_periodic} of Definition~\ref{defn:many_effective_walls}. \textbf{Condition~\eqref{item:bust_point}:} Let $y\in V$ be regular and let $\epsilon>0$. Let $\mathbf S$ be a finite subtree of $\widetilde V_0$ such that each contractible subspace of $V$ has one or more lifts to $\mathbf S$. Let $x_0\in V$ be a periodic point in the interior of the edge $e_0$ containing $y$, chosen so that $\textup{\textsf{d}}_{e_0}(x_0,y)<\epsilon$. This choice is possible since periodic points are dense in each edge of $V$. Indeed, by irreducibility of $\mathfrak M$, for each edge $e$ of $V$, and each subinterval $d\subset e$, there exists $k>0$ such that the path $\phi^k(d)$ properly contains $e$. Brouwer's fixed point theorem implies that $d$ contains a $\phi^k$-fixed point. Let $e_1,\ldots,e_r$ be the edges of $V$, except $e_0$. Suppose that we have chosen periodic points $\{x_i\in\interior{e_i}\}$ for $0\leq i< s$, for some $s\leq r$, with the property that $\rho(\tilde x_i)\neq\rho(\tilde x_j)$ for $i\neq j$ and any lifts $\tilde x_i,\tilde x_j$ of $x_i,x_j$ to $\mathbf S$. Let $\tilde e_{i1},\ldots,\tilde e_{ip_i}$ be the lifts of $e_i$ to $\mathbf S$. Likewise, let $\tilde x_{ij}$ be the lift of $x_i$ to $\tilde e_{ij}$. Choose $x_{s}\in\interior{e_{s}}$ to be a periodic point with the property that no lift of $x_s$ to $\mathbf S$ lies in $\cup_{i<s,j}\rho^{-1}(\{\rho(\tilde x_{ij})\})$. Let $\tilde x_{s1},\ldots,\tilde x_{sp_s}$ be the lifts of $x_s$ to $\mathbf S$. Iterating this procedure, we obtain a set $\{x_0,\ldots,x_r\}$ of periodic points in $V$ such that: \begin{enumerate} \item Each edge $e_i$ of $V$ contains exactly one point $x_i$ in its interior. \item The point $x_0$ lies in the interior of the edge $e_0$ containing $y$ and $\textup{\textsf{d}}_{e_0}(x_0,y)<\epsilon$. \item Let $\tilde x_{ip}\neq\tilde x_{jq}$ be lifts of $x_i,x_j$ in the closure of a lift of a component of $V-\cup_i\{x_i\}$. Then $\rho(\tilde x_{ip})\neq\rho(\tilde x_{jq})$. This holds by construction when $i\neq j$, and holds by Proposition~\ref{prop:R_tree_base case}.(5) when $i=j$ and $p\neq q$. \end{enumerate} For each $\tilde x_{ip}$, let $\periodicline_{ip}$ be the bi-infinite periodic forward path containing $\tilde x_{ip}$. Let $N(\periodicline_{ip})$ be the 1-skeleton of the smallest subcomplex of $\widetilde X$ containing $\periodicline_{ip}$, so that $N(\periodicline_{ip})$ is $\lambda$-quasiconvex in $\widetilde X^1$ by Proposition~\ref{prop:forward_ladder_quasiconvex}. We now show that for each $R\geq 0$ there exists $B_R$ such that $$\diam(\mathcal N_R(N(\periodicline_{ip}))\cap\mathcal N_R(N(\periodicline_{jq})))\leq B_R$$ whenever $\periodicline_{ip}\neq\periodicline_{jq}$. Since $\periodicline_{ip},\periodicline_{jq}$ are periodic, they either fellow-travel or have bounded coarse intersection; the following argument precludes the former possibility, whence the claimed $B_R$ exists since there are finitely many pairs $\periodicline_{ip},\periodicline_{jq}$. Let $\textup{\textsf{d}}_{(ip,jq)}=\textup{\textsf{d}}_{\infty}(\rho(\tilde x_{ip}),\rho(\tilde x_{jq}))$. By definition of $\textup{\textsf{d}}_{\infty}$, when $\periodicline_{ip}\neq\periodicline_{jq}$, there exists $n^o_{(ip,jq)}>0$ such that for all $n\geq n^o_{(ip,jq)}$ we have $$\textup{\textsf{d}}_{n}(\tilde\phi^{n}(\tilde x_{ip}),\tilde\phi^{n}(\tilde x_{jq}))\geq\frac{\varpi^{n}\textup{\textsf{d}}_{(ip,jq)}}{2}.$$ Let $n_{(ip,jq)}\geq n^o_{(ip,jq)}$ have the property that $\varpi^{n_{(ip,jq)}}\textup{\textsf{d}}_{(ip,jq)}\geq 2R$. Let $m=\max\{n_{(ip,jq)}\}$. Then for all $\periodicline_{ip}\neq\periodicline_{jq}$, and all $n\geq m$ we have $$\textup{\textsf{d}}_{n}(\tilde\phi^{n}(\tilde x_{ip}),\tilde\phi^{n}(\tilde x_{jq}))\geq R.$$ We now construct the uniformly sub-quasiconvex spreading set $\mathbb W$. For $L\geq 1$, let $\epsilon'=\frac{\epsilon}{\varpi^L}$. By Lemma~\ref{lem:choosing_busts_1}, there exist primary busts $d_i\subset e_i$, each disjoint from its $\phi^L$-preimage, with $d_i\subset\mathcal N_{\epsilon'}(x_i)$. Let $W\rightarrow X$ be the immersed wall with tunnel-length $L$ and primary busts $d_i$. Choose $J$ such that $\phi^J(x_s)=x_s$ for all $0\leq s\leq r$. We choose $\mathbb W$ to be the set of all walls constructed in this way, where $L$ is divisible by $J$. $\mathbb W$ is uniformly sub-quasiconvex since each component of $V-\cup_i{\interior{e_i}}$ is a finite tree. Let $T_i,T_j$ be distinct tunnels of $\overline W$ and suppose that $\comapp(T_i),\comapp(T_j)$ intersect a common nucleus approximation $\mathbf N$. The forward parts of $\comapp(T_i),\comapp(T_j)$ begin at endpoints of primary busts $\tilde d_{ip},\tilde d_{jq}$ which are lifts of primary busts $d_{i},d_{j}$ near the periodic points $x_i,x_j$ respectively. Let $\tilde x_{ip},\tilde x_{jq}$ be the lifts of $x_i,x_j$ at distance $\epsilon'$ from $\tilde d_{ip},\tilde d_{jq}$. There are three cases according to whether each of $\comapp(T_i),\comapp(T_j)$ is incoming or outgoing at $\mathbf N$. In the case where one is incoming and the other outgoing, consideration of the map $q$ shows that the diameter of the intersection of the $R$-neighborhoods of $\comappn{T_i}$ and $\comappn{T_j}$ is bounded by a function of $R$. Suppose that $\comapp(T_i)$ and $\comapp(T_j)$ are both outgoing from $\mathbf N$. Our choice of $\epsilon'$ ensures that $\comapp(T_i)$ fellow-travels at distance $\epsilon$ with the forward path of length $L$ emanating from $\tilde x_{ip}$ and similarly for $\comapp(T_j)$ and $\tilde x_{jq}$. (More precisely, each point of $\comapp(T_i)\cap\widetilde X^1$ is at distance at most $\epsilon$ from the corresponding point of the forward path emanating from $\tilde x_{ip}$.) Hence the coarse intersection of $\comapp(T_i)$ and $\comapp(T_j)$ is controlled by the function $R\mapsto B_R$ and the uniform constant $\epsilon$. Suppose that $\comapp(T_i)$ and $\comapp(T_j)$ are both incoming to $\mathbf N$. By translating, we can assume that $\mathbf N\subset\mathbf S$. Because $J\mid L$, we have that $\tilde\phi^L(\tilde x_{ip})$ and $\tilde\phi^L(\tilde x_{jq})$ are again lifts of $x_i,x_j$ to $\mathbf N\subset\mathbf S$ and thus lie on the bi-infinite periodic forward paths $\periodicline_{ip},\periodicline_{jq}$ that diverge according to the map $R\mapsto B_R$. As before, $\comapp(T_i)$ and $\comapp(T_j)$ are (uniformly) coarsely contained in the $\epsilon$-neighborhoods of $\periodicline_{ip}$ and $\periodicline_{jq}$. \textbf{Condition~\eqref{item:w_a_periodic}:} Let $a\in\widetilde V_0$ and let its image $\bar a\in V$ be periodic with period $J_a$. As before, let $\mathbf S$ be a finite subtree of $\widetilde V_0$ containing $a$ and having the property that every contractible subspace of $V$ lifts to $\mathbf S$ and let $e_0,\ldots,e_r$ be the edges of $V$, with $\bar a\in e_0$. Let $x_{-1}=\bar a$. We temporarily subdivide $e_0$, writing $e_0=e_{-1}'e_0'$ with $x_{-1}\in e'_{-1}$. We now apply Lemma~\ref{lem:r_tree_quotient} to $V$, and then remove the subdivision vertex, yielding periodic points $x_i\in\interior{e_i}, 0\leq i\leq r$ so that: for all $i,j\geq -1$ and all $n\geq 0$, any lifts $\tilde x_{ip},\tilde x_{jq}$ of $\phi^n(x_i),\phi^n(x_j)$ to $\mathbf S$ satisfy $\rho(\tilde x_{ip})\neq\rho(\tilde x_{jq})$. As before, let $J$ be the least common multiple of the periods of the $x_i$. Let $L\geq 0$. Applying Lemma~\ref{lem:choosing_busts_1}, for each $i\geq 0$ let $d_i\subset\interior{e_i}$ be a primary bust such that $d_i\subset\mathcal N_{\frac{C}{\varpi^L}}(x_i)$ and such that there is an immersed wall $W\rightarrow X$ with tunnel length $L$ and primary busts $d_i$. The collection $\mathbb W_a$ of such walls with $J\mid L$ is uniformly bust-quasiconvex since each component of the complement of the primary busts is contractible. Arguing as in the verification of Condition~\eqref{item:bust_point}, the characteristic property of $\{x_i\}$ ensures that $\mathbb W_a$ has uniformly bounded ladder overlap (the bound is independent of $L$). Likewise, there is a uniform bound $k(a)$ on $3\delta+2\lambda$ fellow-traveling between two forward ladders, one emanating from an endpoint of $\tilde d_{ip}$ and one from $a=\tilde x_{0q}$, whenever $\tilde d_{ip}$ is a lift of some $d_i$ that is joined to $a$ by a path in a knockout of $\overline W$. Indeed, in this situation, $\tilde\phi^L(a)$ is a lift of $\bar a$ to the finite nucleus approximation containing the lift $\tilde\phi^L(\tilde x_{ip})$ of $x_i$, whence the forward paths emanating from $\tilde d_{ip}$ and $a$ have uniformly bounded coarse intersection. The other case, where $a$ and $\tilde d_{ip}$ lie on the same nucleus approximation, is handled as in the analogous case in the verification of Condition~\eqref{item:bust_point}. \end{proof} \begin{lem}\label{lem:r_tree_quotient} Let $x_{-1}\in V$ be a periodic point in an edge $e_{-1}$ and let $e_0,\ldots, e_r$ be a collection of edges in $V$. Then for $0\leq i\leq r$, there exist periodic points $x_i\in\interior{e_i}$ such that for all $i,j\geq -1,n\geq 0$ and for all distinct lifts $\tilde x_{ip},\tilde x_{jq}$ of $\phi^n(x_i),\phi^n(x_j)$ to $\mathbf S$, we have $\rho(\tilde x_{ip})\neq\rho(\tilde x_{jq})$. \end{lem} \begin{proof} For all $i$, any two distinct lifts of $\phi^n(x_i)$ to $\mathbf S$ have distinct images in $\mathcal Y$ by Proposition~\ref{prop:R_tree_base case}.(5). It therefore suffices to verify the claim of the lemma for points $\tilde x_{ip},\tilde x_{jq}$ with $i\neq j$. We argue by induction on $r$. In the base case where $r=-1$, there is nothing to prove. Supposing that $x_{-1},\ldots,x_{r-1}$ satisfy the conclusion of the lemma, we will choose $x_r$. Since $\rho$ is an embedding on each edge and $\mathbf S$ is the union of finitely many edges, there exists $K\in\ensuremath{\field{N}}$ such that for all $y\in\rho(\mathbf S)$, we have $|\rho^{-1}(y)\cap\mathbf S|\leq K$. Let $$Q=|\{\rho(\tilde x_{ip}):-1\leq i\leq r-1,\,1\leq p\leq p_i\}|,$$ where $p_i$ is the number of lifts of $x_i$ to $\mathbf S$. Choose $m\in\ensuremath{\field{N}}$ such that $e_r$ intersects at least $KQ+1$ $\phi$-orbits of $m$-periodic points. This choice is possible because, for arbitrarily large $m$, the number of $m$-periodic points in $e_r$ is approximately $C\varpi^m$ for some $C>0$, while the claimed $\phi$-orbits exist as long as there are at least $(KQ+1)m$ periodic points in $e_r$ with period $m$. For each such $m$-periodic $u$, a \emph{lifted orbit} of $u$ is the set of all lifts to $\mathbf S$ of all points $\phi^k(u)$ with $0\leq k<m$. Note that if $u,u'$ are $m$-periodic points with distinct $\phi$-orbits, then their lifted orbits are disjoint since their projections to $V$ are distinct $\phi$-orbits of the same cardinality and are hence disjoint. By the pigeonhole principle, there exists an $m$-periodic point $x_r\in e_r$ with the desired property. Indeed, the points $\rho(\tilde x_{ip})$ with $i<r$ ruled out at most $KQ$ of the $KQ+1$ lifted orbits. \end{proof} \begin{lem}\label{lem:periodic_paths_everywhere} Let $\phi$ be as in Theorem~\ref{thm:irreducible}. Then for any finite forward path $\sigma\rightarrow\widetilde X$, there exists a periodic forward path $\chi$ with $\sigma\subset N(\chi)$. If $\sigma$ is regular, then $N(\sigma)=N(\chi)$. \end{lem} \begin{rem} The period of $\chi$ is unbounded as the length of $\sigma$ increases. \end{rem} \begin{proof} This follows from the fact that periodic points are dense in $V$ and the fact that distinct forward rays diverge at a rate governed by $\varpi$. Let $x$ be the initial point of $\sigma$ and let $n=|\sigma|$. We choose $y$ to be a point at distance at most $\frac{\varpi^n}{2K}$ from $x$, with the image of $y$ in $V$ periodic regular, where $K$ is the distance from $x$ to a nearest vertex. Then the length-$n$ forward path $\sigma_y$ fellow-travels with $\sigma$ at distance $\frac{1}{2K}$, and hence the first and last vertical edges in the carriers of $\sigma,\chi$ are equal. \end{proof} We conclude with the following: \begin{cor}\label{cor:irreducible_case} Let $\Phi:F\rightarrow F$ be a monomorphism of the finitely generated free group $F$. Suppose that $\Phi$ is irreducible and that the ascending HNN extension $G=F*_{\Phi}$ is word-hyperbolic. Then $G$ acts freely and cocompactly on a CAT(0) cube complex. \end{cor} \begin{proof} This follows from the fact that such $\Phi$ is represented by a map $\phi:V\rightarrow V$ satisfying the hypotheses of Theorem~\ref{thm:irreducible}. Indeed, any irreducible endomorphism has an irreducible train track representative~\cite{BestvinaHandel,Reynolds,DicksVentura}. \end{proof} \bibliographystyle{alpha}
\section{Numerical procedure} \label{Sec:Numerical} For a given realization of CN impurities, the low energy state of a KBr$_{1-x}$(CN)$_x$ sample corresponds to a specific orientation of each of the CN impurities in the sample, which is dictated by the impurity-impurity interactions. At the same time, the exact positions of all ions in the sample giving energy minimization have to be calculated. We therefore use a technique that hybridizes local energy minimization (conjugate gradients) and MC simulation. We start by creating a 3D grid of volume $N \times N \times N$ ($N \times N$ in 2D, $N$ is even) of $K$\textsuperscript{+} and $Br$\textsuperscript{-} ions, having distance $3.1974\AA$ ($3.2735\AA$ in 2D) between the ions. These distance values were calculated by the energy minimization procedure of pure KBr grid. We replace randomly some of the $Br$\textsuperscript{-} ions by $CN$\textsuperscript{-} ions, according to our chosen concentration $x=0.25$. The charges of $K$\textsuperscript{+} and $Br$\textsuperscript{-} ions are taken as +1 and -1 respectively, and the charge of the $CN$\textsuperscript{-} ion is represented by fractional charges q$_{C}=-1.28$, q$_{N}=-1.37$ and q$_{center}=+1.65$ placed respectively on the carbon atom, nitrogen atom and at the center of mass \cite{KM83}. The distance of the carbon and nitrogen atoms from the center of mass are taken as $0.63\AA$ and $0.54\AA$ respectively \cite{KM83}. Interatomic potential is calculated by the formula: \begin{equation} V_{\alpha \beta}(R) = A_{\alpha \beta}\exp(-a_{\alpha \beta}R) + {B_{\alpha \beta} \over R^6} + K{q_{\alpha}q_{\beta} \over R}. \label{potential} \end{equation} The interatomic potential parameters A$_{\alpha \beta}$, $a_{\alpha \beta}$ and B$_{\alpha \beta}$, see Table~\ref{tab:params}, are taken from Ref. \cite{BK82}, and $K=1389.35 \AA$kJ/mol. \begin{table} \def\arraystretch{1.5} \begin{tabular}{|c|c|c|c|} \hline $_{\alpha \alpha}$ & $A_{\alpha \alpha}$ (kJ/mol) & $a_{\alpha \alpha}$ (1/$\AA$) & $B_{\alpha \alpha}$ ($\AA^6$kJ/mol) \\ \hline KK & 158100 & 2.985 & -1464 \\ CC & 259000 & 3.600 & -2110 \\ NN & 205020 & 3.600 & -1803 \\ BrBr & 429600 & 2.985 & -12410 \\ \hline \end{tabular} \caption {Interatomic potential parameters (taken from Ref. \cite{BK82}). Cross-interaction parameters were calculated by $A_{\alpha \beta}=(A_{\alpha \alpha}A_{\beta \beta})^{1/2}$, $a_{\alpha \beta}=(a_{\alpha \alpha}a_{\beta \beta})/2$, $B_{\alpha \beta}=-(B_{\alpha \alpha}B_{\beta \beta})^{1/2}$. } \label{tab:params} \end{table} For a single CN impurity in an otherwise pure KBr lattice we find that its preferred orientation is along the eight-fold degenerate in-space diagonals, in agreement with Ref.~[\onlinecite{Bey75}]. In two-dimensions the degeneracy is four-fold, along the plane diagonals. These degeneracies are lifted by tunneling. However, in strongly disordered systems, of interest to us here, the bias disorder is much larger than the tunneling amplitudes\cite{SS09}. Calculation of the bias energy can therefore be carried out without consideration of the tunneling. The tunneling amplitude, which for almost all TTLSs is much smaller than the bias energy, can then be considered after the distribution of biases, which is of interest for us here, has been established. We thus start by orienting each $CN$\textsuperscript{-} ion in one of the inspace diagonals (plane diagonals in 2D) randomly. We then relax the system for the given initial orientation of the $CN$\textsuperscript{-} impurities to the local energy minimum, optimizing all ion locations as well as $CN$\textsuperscript{-} orientations using the non-linear Fletcher-Reeves conjugate gradients method. Periodic boundary conditions are used to simulate the infinite crystal. The conjugate gradient method, however, does not allow the crossings of the high barriers separating different single CN states. Such crossings are required, though, to find a low energy state of the full system, as is dictated by the impurity-impurity interactions. We thus combine the conjugate gradients method with intervening MC steps. After the initial energy minimization, the orientation of one $CN$\textsuperscript{-} ion is changed randomly to another in space diagonal (in plane diagonal in 2D), and the system's energy is minimized again using conjugate gradient method. The new configuration is accepted or rejected according to the standard Metropolis algorithm and the assigned temperature. These steps are carried at 40 different temperatures from $300$K to $0.02$K, with 256 MC steps for each temperature, where a single MC step involves a single orientation change of each $CN$\textsuperscript{-} ion in the sample. We then do 256 MC steps at $10^{-5}$ K, and reach a low energy metastable state. The final state of the system is not necessarily the ground state. However, such metastable states are expected to represent well the real low temperature glassy state of KBr:CN and its low energy DOS. The latter is dictated by the Efros Shklovskii gap\cite{ES75}, which relies only on stability to single and double $CN$\textsuperscript{-} flips or rotations\cite{BEGS79}. Once we reach the final state of the simulation we measure all single particle excitation energies. We also keep track of the symmetry of the excitations. The symmetric excitations, i.e. those where the final orientation of the $CN$\textsuperscript{-} ions is nearly inverse to its initial orientation (within $14^\circ$) are denoted $\tau$-TTLSs, and the asymmetric excitations are denoted $S$-TTLSs\cite{SS09}. This procedure is then repeated for 3000 samples in 3D and 5000 samples in 2D. \begin{figure} \subfigure[]{ \includegraphics[width=8.5cm]{Fig1a.eps} \label{fig:Fig1subfig1} } \subfigure[]{ \includegraphics[width=8.5cm]{Fig1b.eps} \label{fig:Fig1subfig2} } \subfigure[]{ \includegraphics[width=8.5cm]{Fig1c.eps} \label{fig:Fig1subfig3} } \caption{ The DOS of single impurity tunneling in 3D in 4x4x4 unit cell with CN concentration x=0.25. (a) Full DOS up to $1000$K. (b) DOS of CN flips ($\tau$-excitations). Solid line denotes a one parameter Gaussian fit, with standard deviation $E_o=8.99$. (c) DOS of CN rotations ($S$-excitations). Solid line in the inset denotes a fit to the function $A/ln(B/E)$ ($A=1.56 \times 10^{-3}, B=500$). } \label{fig:DOS3D} \end{figure} \begin{figure} \subfigure[]{ \includegraphics[width=8.5cm]{Fig2a_2.eps} \label{fig:Fig2subfig1} } \subfigure[]{ \includegraphics[width=8.5cm]{Fig2b_2.eps} \label{fig:Fig2subfig2} } \subfigure[]{ \includegraphics[width=8.5cm]{Fig2c_2.eps} \label{fig:Fig2subfig3} } \caption{ The DOS of single impurity tunneling in 2D in 8x8 unit cell with CN concentration x=0.25. (a) Full DOS up to $1000$K. (b) DOS of CN flips ($\tau$-excitations). Solid line denotes a one parameter Gaussian fit, with standard deviation $E_o=8.65$. (c) DOS of CN rotations ($S$-excitations). Solid line in the inset denotes a fit to the function $A/ln(B/E)$ ($A=7.7 \times 10^{-4}, B=1500$). } \label{fig:DOS2D} \end{figure} \section{Results} \label{Sec:Results} In Fig.\ref{fig:DOS3D}(a) and Fig.\ref{fig:DOS2D}(a) we plot the full DOS of single impurity tunneling in 3D ($4\times4\times4$ unit cells) and 2D ($8\times8$ unit cells) respectively. The DOS has a maximum at low energies, but reduces considerably at an energy $\approx 3$K. We now plot separately the single particle excitations of CN flips ($\tau$-TTLSs) and rotations ($S$-TTLSs) in Fig.\ref{fig:DOS3D}(b),(c) for the 3D samples and in Fig.\ref{fig:DOS2D}(b),(c) for the 2D samples. The difference between the DOS of the two types of excitations is striking. The inversion symmetric flip excitations are peaked at low energies, with typical energies of $\approx 7$K for both 3D and 2D. The inversion non symmetric rotation excitations, however, are broadly distributed, with typical energies of $\approx 300$K for 3D and $\approx 500$K for 2D. The peak value of the $\tau$-TTLS DOS is $\sim 14$ and $\sim 40$ times larger than the peak value of the $S$-TTLS DOS in 3D and 2D respectively. The difference between the peak values of the $S$-TTLSs in 3D and in 2D is a consequence of each CN impurity having two $S$ excitations in 2D, but six $S$ excitations in 3D. Focusing on the CN flips ($\tau$-TTLSs), their DOS is quite close to a Gaussian, except for a small dip in the DOS at very low energies [see inset of Figs.~\ref{fig:DOS3D}(b),~\ref{fig:DOS2D}(b)], and some structure at $\sim 10$ K. The former is generic to the two-TTLS model (see below) whereas the latter is a result of details of the lattice structure and form of the interaction. With regard to the $S$-TTLSs, their DOS is well fit to a log function at low temperatures, except below $10$ K, the energy at which the $\tau$-TTLSs appear, see further discussion below. \subsection{Two-TTLS model} \label{Sec:TwoTLS} Our results for the DOS are intimately connected with the fact that the coupling strength to the phonon field of inversion symmetric excitations $\gamma_{\rm w}$ is much smaller than that of the asymmetric excitations $\gamma_{\rm s}$. For KBr:CN $\gamma_{\rm w} \approx 0.1eV$, whereas $\gamma_{\rm s} \approx 3eV$\cite{GS11}. This bi-modality in the coupling strengths affects the phonon mediated acoustic interactions, which then result in an effective TTLS-TTLS interaction Hamiltonian of the form\cite{CBS13,SS09} \begin{equation} H_{\rm S\tau}^{\rm eff} = \sum_{ij} U_{ij}^{SS} S_i^z S_j^z + \sum_{ij} U_{ij}^{S \tau} S_i^z \tau_j^z + \sum_{ij} U_{ij}^{\tau \tau} \tau_i^z \tau_j^z \label{effectiveStau} \end{equation} where in three dimensions (two dimensions) all interactions decay as $1/r^3$ ($1/r^2$) at distances $r \gg a_0$, with a short distance cutoff $\tilde{a} \sim a_0$, and their typical values at near neighbor distance are related by\cite{SS08b,SS09,CBS13} \begin{equation} U_0^{\tau \tau} \approx g U_0^{S \tau} \approx g^2 U_0^{S S} . \label{gratios} \end{equation} Here $g \equiv \gamma_{\rm w}/\gamma_{\rm s} \approx 1/30$\cite{SS09,GS11}. This effective Hamiltonian was shown to lead to quantitative universality and to account for the energy scale of $\approx 0.2 U_0^{S \tau} \approx 0.2gU_0^{SS} \approx 3$K related to it\cite{SS09}. The DOS for the two types of TTLSs within the Hamiltonian~\ref{effectiveStau} was analyzed in detail in Ref.\cite{CGBS13}. Indeed, it was found that the weakly interacting ($\tau$) TTLSs DOS are well fit by a Gaussian, except for a small dip, at very low energies, of relative magnitude $\sim g$. The Gaussian width is dictated by the strength of the $S-\tau$ interaction $U_0^{S \tau}$. The strongly interacting $S$-TTLSs were found to have a typical energy dictated by $U_0^{SS}$, a logarithmic gap at intermediate energies\cite{BSE80,Bur95}, and a power-law gap below $\approx U_0^{S \tau}$, the energy where the DOS of the $\tau$-TTLSs becomes appreciable. As a consequence of this gapping of the $S$-TTLS DOS $n_S(E)$, one finds below $\approx 3$ K that $n_S(E) \gamma_S^2 < n_{\tau}(E) \gamma_w^2$ (i.e. $n_S(E)/n_{\tau}(E) < g^2$). This defines the condition for the $\tau$-TTLS to dominate phonon attenuation, resulting in universality below this energy scale. Our results here validate the applicability of the above two-TTLS model to real systems, as we verify, in a calculation relying only on the bare interatomic interactions of a real glass, the central features predicted by the model: (i) The width of the distributions (typical energies) of the $S$-TTLSs and $\tau$-TTLSs - both their relative magnitudes and the absolute magnitude of each one. We further note that the value of $g$ inferred from the ratio between the typical energies is in good agreement with that obtained from the ratio of the interaction constants in Ref.\cite{GS11}; (ii) The gapping of the $S$-TTLSs but not the $\tau$-TTLSs at low energies, and the small dip, of relative magnitude $\sim g$, of the $\tau$-TTLSs at the lowest energies: (iii) The change in the functional form of the gap of the $S$-TTLSs at the energy scale $\sim 10$ K where the $\tau$-TTLSs appear: (iv) The relation $n_S(E)/n_{\tau}(E) < g^2$ is fulfilled at energies smaller than $\sim 3$ K. We note that the energy below which the above condition is fulfilled in KBr:CN is slightly larger than what is inferred by our results, because as a consequence of finite size effect our results over estimate the $S$-TTLS DOS at the lowest energies. \section{Discussion} \label{Sec:Discussion} In contrast to the assumption of the STM, we find that in KBr:CN TTLSs are divided to two distinct classes by their local inversion symmetry, with a DOS distribution which is particular to each class of TTLSs, and has a strong energy dependence at low energies, of order $10$ K. Many disordered lattices, of which KBr:CN is a primary example, share the low temperature universal characteristics with amorphous solids. Furthermore, experiments strongly support the notion that it is the same mechanism, pertaining to the disordered state of matter itself, that dictates universality in amorphous systems and disordered lattices alike. The structure we find for the DOS of the symmetric and asymmetric TTLSs is not only very different from the flat DOS suggested by the STM, but it is this very structure of DOS that provides an explanation of the quantitative universality of phonon attenuation and the energy scale of $\approx 3$K below which universality is observed\cite{SS09}. As it is expected that the same model would be relevant to both amorphous solids and disordered lattices showing universality, our results suggest that also in amorphous solids two types of TTLSs exist, with a similar structure of DOS to the one found here. Verification of this idea could lead not only to a resolution of the long standing problem of the low temperature universality in disordered systems, but also to an enhanced understanding of the microscopic structure of amorphous solids, and its relation to the physical properties of amorphous solids in general. The standard tunneling model uses only one feature of the complicated structure of the density of states found here - the rather homogeneous DOS of the inversion symmetric TTLSs below $3$K. It completely neglects not only the sharp reduction of the DOS at higher energies, but also the existence of the asymmetric TTLSs with their much larger interaction with the phonon field. These characteristics of the DOS, however, may lead to the explanation of further intriguing properties of disordered lattices and amorphous solids at low and intermediate temperatures. Some examples at intermediate temperatures are the full temperature dependence of the thermal conductivity including the rather universal plateau between $3-10$K\cite{YKMP86,FA86}, and the boson peak. At low temperatures, the existence of the strongly interacting $S$-TTLSs, although small in number, may be useful in explaining equilibrium and non-equilibrium properties of phonon attenuation, as well as the relaxation and decoherence of the prevalent $\tau$-TTLSs. In addition to the standard TTLSs responsible to the universal phonon attenuation at low temperatures, a second type of local excitations was introduced previously, phenomenologically, by Black and Halperin\cite{BH77} and by Yu and Freeman\cite{YF87}. Black and Halperin suggested the existence of TTLSs with a much smaller interaction with the strain, thus affecting the specific heat but not phonon attenuation. Yu and Freeman suggested the existence of local phononic modes with a similar interaction with the phonon field to that of the TTLSs, but with a DOS which is gapped below $43$ K, to fit experimental data for thermal conductivity at intermediate and high temperatures. Our results differ from the above two approaches in that TTLSs of the second type have a much larger interaction with the strain, and are softly gapped below $\sim 3$ K. Experimentally, TTLSs are observed via a plethora of techniques, including echo\cite{GG76,GSHD79,NFHE04}, coupling to superconducting qubits\cite{LSC+10,GPL+12}, specific heat\cite{DFA86} and dielectric response\cite{EWNS88}. In general, our results suggest that care should be taken in experiments to characterize the TTLSs according to their interaction strength. Whereas most experiments are susceptible to the weakly interacting TTLSs abundant at low energies, strongly interacting TTLSs can dominate experimental observations susceptible to their high bias energy or strong interaction with the strain. Specifically, our results disagree with previous estimates for typical energy biases of CN flips in KBr:CN\cite{SC85,SK94}, and suggest that experimental values of $300-400$ K obtained for TTLS-TTLS interaction at short distances, and for TTLS bias energies, in KBr:CN\cite{DFA86,EWNS88} concern CN rotations rather than CN flips. The use of the empirical potential of the (Buckingham-Coulomb) form given in Eq.~(\ref{potential}) rather than the more accurate DFT or ab-initio formalisms is dictated by the required complexity of the problem: the need to consider a relatively large system, to find its low energy state using Monte Carlo, and to repeat the calculation for many samples for good statistics. This use of the empirical potential has been shown to be useful in obtaining qualitative, and to some degree also quantitative, predictions for ionic crystals (for details see Refs.~\cite{BK82,KM83,LC85,HUH+05}). For the properties which are of interest to us here, i.e. the DOS of the $CN$\textsuperscript{-} flips ($\tau$-TLSs) and $CN$\textsuperscript{-} rotations ($S$-TLSs), all qualitative features are dictated by symmetry and are therefore robust. Comparison to results for the model Hamiltonian in Eq.(\ref{effectiveStau})\cite{GS11,CGBS13} suggest that all qualitative features of the DOS for both $CN$\textsuperscript{-} flips and $CN$\textsuperscript{-} rotations are indeed retrieved by our calculation. Quantitatively, the value of $g$ inferred from our results for the typical energies of $CN$\textsuperscript{-} flips and $CN$\textsuperscript{-} rotations is in very good agreement with that found by the ratio of their corresponding couplings to the phonon field using DFT and ab-initio calculations\cite{GS11}. We also find good quantitative correspondence with the calculations for the model Hamiltonian in Eq.(\ref{effectiveStau}) as obtained in Ref.~\cite{CGBS13}. This suggests that our results provide also reasonable quantitative accuracy, except for the over estimate of the DOS of the $S$-TLSs at the very low energies, as mentioned above. We are concerned here with the calculation of the bias energies of the $S$-TLSs and the $\tau$-TLSs, but not their tunneling amplitudes. The latter are crucial for the discussion of dynamic and non-equilibrium properties. A plausible assumption for the distribution of the tunneling amplitudes $\Delta_0$ is that suggested in Refs.\cite{AHV72,Phi72}, $P(\Delta_0) \propto 1/\Delta_0$. For the $\tau$-TLSs which dominate the low energy universal phenomena, this assumption has been shown to be consistent with the various experimental findings. However, its applicability for the $S$-TLSs needs to be checked. Such a verification may become possible by using our results for the bias energies here, and for the coupling constant of the $S$-TLSs in Ref.\cite{GS11} to calculate measurable quantities dominated by $S$-TLSs and comparing to experimental results. {\it Acknowledgments} --- We thank A. Burin, A. Gaita-Ari\~no, Niels Gr{\o}nbech-Jensen, and Joerg Rottler for very useful discussions. This research was supported by the Israel Science Foundation (Grant No. 982/10).
\section{Introduction} Let $T>0$, $(\Omega,\mathcal{F},\mathbb{P})$ be a probability space and $V$ is a Banach space. Suppose the space-time continuous stochastic process $X:[0,T]\times \Omega \rightarrow V$ is the unique solution of the the following stochastic partial differential equation (SPDE) \begin{equation}\label{intr1} dX_{t}=\left[AX_{t}+F(X_{t})\right]dt+dW_{t}, \qquad X_{t}(0)=X_{t}(1)=0, \quad X_0=0, \end{equation} for $t\in [0,T]$ and $x\in (0,1)$, where the operator $A$ denotes an unbounded operator, for example the Laplacian. The noise is given by a Wiener process $W_{t}$, $t\in [0,T]$ defined later. The main purpose of this article is to consider a spectral Galerkin approximation of (\ref{intr1}) in $L^\infty$, where the noise is colored. The main results are formulated in an abstract way so that in principle they should apply to other approximation methods like finite elements, but here we only verify the applicability for spectral Galerkin methods for simplicity of presentation. A key point is the uniform bound on the numerical data. Alternatively, we can uniformly bound the Galerkin approximation, which is for spectral methods frequently straightforward to verify, by using energy-type a-priori estimates. Of course the result should apply for higher dimensional domains, differential operators of higher order, or other boundary conditions like Dirichlet, but as an example we stick with this relatively simple situation here. In \cite{Ref1362} the Galerkin approximation was already considered for a stochastic Burgers equation with colored noise, but here we present this method in a more general setting, and not only for the Burgers equation. The main novelty, as in \cite{Ref1362} or \cite{Ref1}, is to bound the spatial and temporal discretization error in the uniform topology. The space of continuous or H\"older-continuous functions is a natural space for stochastic convolutions. For instance, if for space-time white noise the stochastic convolution is in $L^2$ in space, it is already continuous. In a recent publication \cite{Ref2014} Cox \& van Neerven established a time-discretization error in H\"older spaces, but the spatial error in UMD-spaces. We strongly believe, that working in fractional Sobolev-spaces $W^{\alpha,p}$ with $\alpha>0$ small and $p\gg1$ large should yield similar results than ours, but we present here a simple proof yielding uniform bounds in time only. In \cite{Ref1, Ref2} the Galerkin approximation was considered for a simple case of SPDEs of the type of (\ref{intr1}), either without time-discretization or in different spaces. Moreover, the Brownian motions in the Fourier expansion of the noise are independent. But in general the spatial covariance operator of the forcing does not necessarily commute with the linear operator $A$, thus we consider here the case where the Brownian motions are not independent. Many authors have investigated the spectral Galerkin method for this kind of equation with space-time white noise. See for example \cite{Ref13, Ref1303, Ref2009, Ref2, Ref2310, Ref1312, Ref2304}. There are also many articles about finite difference methods \cite{Ref5, Ref4, Ref13, Ref1308, Ref16}. The existence and uniqueness of solutions of the stochastic equation was studied in \cite{Ref1995, Ref6} for space-time white noise. In our proofs, as the nonlinearity allows for polynomial growth, we do not rely on the global existence of solutions, but assume that the numerical approximation remains uniformly bounded. In the limit of fine discretization, this will ensure global existence of the solutions and a global error bound for the numerical approximation. Our aim here is to extend the results of \cite{Ref1362} to the case of more general nonlinearities, with local Lipschitz conditions and polynomial growth. For spatial discretization of equation (\ref{intr1}) we apply a spectral Galerkin approximation as already discussed in \cite{Ref1} and for the time discretization we follow the method proposed in \cite{Ref2}. It should be mentioned that the spatial discretization error is obtained by the results of \cite{Ref1, Ref1362}. We will recall their main results in Section 1. In this article we focus on the time discretization. Not treated in \cite{Ref1} but already in \cite{Ref1362}, we consider here also the case of colored noise being not diagonal with respect to the eigenfunctions of the Laplacian. As the final result we obtain an error estimate for the full space-time discretization. The main result of \cite{Ref1362} in combination with the results presented in this paper yields the convergence results in the uniform topology of continuous functions for the numerical approximations of a wider class of SPDEs with colored noise. The key assumption is a uniform bound on the numerical approximations, that allows for local Lipschitz-conditions only. The paper is organized as follows. Section 2 gives the setting and the assumptions. In Section 3 we recall the results on the spatial discretization error, and in Section 4 estimates for the temporal error are derived. Finally, in the last section a simple numerical example is presented, in order to illustrate the results. \section{Setting and assumptions} Let $V,W$ be two $\mathbb{R}$-Banach spaces such that $V\subseteq W$. Suppose that the unbounded and invertible linear operator $A$ generates an analytic semigroup $S_{t}$ on $V$ that extends to the larger space $W$ , i.e., $S_{t}:W \rightarrow V$. Especially, $S_{t+s}=S_tS_s$ and $S_0=Id$. Consider the following equation \begin{equation}\label{intr} dX_{t}=\left[AX_{t}+F(X_{t})\right]dt+dW_{t}, \qquad X_{t}(0)=X_{t}(1)=0, \quad X_0=0, \end{equation} for $t\in [0,T]$ and $x\in (0,1)$. It should be mentioned that just for simplicity we assumed these initial and boundary conditions. \\ Suppose there are bounded linear operators $P_{N} :V\rightarrow V$. The example we have in mind is the spectral Galerkin method given by the orthogonal projection $P_{N}v=\sum_{i=1}^{N}\int_{0}^{1}e_{i}(s) v(s) ds \cdot e_{i}$, where $\{e_i\}_{i\in\mathbb{N}}$ are an orthonormal basis of eigenfunctions of $A$. But any other approximation method like finite elements should work in a similar way, if we can satisfy our assumptions for the projections. Consider the following assumptions already made in \cite{Ref1}. \begin{assumption}[Semigroup] Suppose for the semigroup, that $S:(0,T]\rightarrow L(W,V)$ is a continuous mapping that commutes with $P_{N}$ and satisfies for given constants $\alpha, \theta \in [0,1)$ and $\gamma \in (0,\infty)$ \begin{equation}\label{S1} \sup_{0<t\leq T}\left(t^{\alpha}\|S_{t}\|_{L(W,V)}\right)< \infty, ~~\sup_{N\in \mathbb{N}} \sup_{0\leq t\leq T}(t^{\alpha}N^{\gamma}\|S_{t}-P_{N}S_{t}\|_{L(W,V)})< \infty, \end{equation} and \begin{equation} \label{e:SGtime} \|A^{\theta} S_{t}\|_{L(V,V)} \leq C t^{-\theta} \quad\text{together with} \quad \|A^{-\theta}(S_{t}-I)\|_{L(V,V)} \leq t^{\theta}. \end{equation} \end{assumption} The first assumption is crucial for the spatial discretization, while the second assumption (\ref{e:SGtime}) is mainly needed for the result on time-discretization, in order to bound differences of the semigroup. For example, for analytic semigroups generated by the Laplacian, this is usually straightforward to verify. See for example \cite{Ref2013}. \begin{assumption}[Nonlinearity] \label{ass:nonlin} Let $F:V\rightarrow W$ be a continuous mapping, which satisfies the following local Lipschitz condition. There is a nonnegative integer $p$ and a constant $L>0$ such that for all $u,v \in V$ \begin{equation}\label{drift condi} \|F(u)-F(v)\|_{W}\leq L\|u-v\|_{V}(1+\|u\|_{V}^{p}+\|v\|_{V}^{p})\;. \end{equation} \end{assumption} Let us remark that it is not a major restriction that we assumed the operator $A$ to be invertible, as we can always consider for some constant $c$ the operator $\tilde A=A+cI$ and the nonlinearity $\tilde F=F-cI$. \subsection{The Ornstein-Uhlenbeck process} \label{suse:OU} \begin{assumption}[Ornstein-Uhlenbeck] Let $O:[0,T]\times \Omega\rightarrow V$ be a stochastic process with continuous sample paths and there exists some $\gamma\in (0,\infty)$ such that \begin{equation}\label{S3} \sup_{N\in \mathbb{N}} \sup_{0\leq t\leq T} N^{\gamma}\|O_{t}(\omega)-P_{N}(O_{t}(\omega))\|_{V}< \infty, \end{equation} for every $\omega \in \Omega$.\\ Moreover, \begin{equation}\label{eq45} \sup_{0\leq t_1\leq t_2\leq T}\frac{\big\|O_{t_2}(\omega)-O_{t_1}(\omega)\big\|_V}{(t_2-t_1)^{\theta}}<\infty,\\ \end{equation} for some $\theta \in (0,\frac{1}{2})$. \end{assumption} In order to give an example for this assumption we focus for the remainder of this subsection now on $L^2[0,1]$ with basis functions $e_k$ are given by the standard Dirichlet basis, where for every $k\in \mathbb{N}$ \begin{equation*}\label{eq2} e_{k}:[0,1]\rightarrow \mathbb{R},~~~ e_{k}(x)=\sqrt{2}\sin(k\pi x),~~x\in [0,1], \end{equation*} are smooth functions. For every $k \in N$ define the real numbers $\lambda_{k}=(\pi k)^{2}\in \mathbb{R}$. Furthermore, let $Q$ be a symmetric non-negative operator, given by the convolution with a translation invariant positive definite kernel $q$. This means \begin{equation}\label{eq:defQ} <Q e_{k},e_{l}>=\int_0^{1}\int_0^{1}e_{k}(x)e_{l}(y)q(x-y)dy dx, \end{equation} for $k,l\in \mathbb{N}$. Note that $Q$ is diagonal with respect to the standard Fourier basis, but in general not with the Dirichlet basis. We think of $Q$ being the covariance operator of a Wiener process $W$ in $L^2(0,1)$ and $q$ being the spatial correlation function of the noise process $\partial_t W(t)$. See for example \cite{DB:noise} for a detailed discussion. Let $\beta^{i}:[0,T]\times \Omega\rightarrow \mathbb{R},i\in \mathbb{N},$ be a family of Brownian motions that are not necessarily independent. We usually think of $\beta^{i}(t)=\langle W(t),e_i\rangle$. See the discussion at end of this subsection. Note that the variance of the Brownian motion is $\sigma_k^2 = <Q e_{k},e_{k}>$, which means that for $\sigma_k\not=0$ the process $\sigma_i^{-1} \beta^{i}(t)$ is a standard Brownian motion. Moreover, the $\beta^{i}$'s are correlated as given by \begin{equation*}\label{eqn2} \mathbb{E}\left[\beta^{k}(t)\beta^{l}(t)\right]= \langle Q e_{k},e_{l} \rangle \cdot t,~~ \text{for } k,l\in \mathbb{N}. \end{equation*} For the regularity assume that for some $\rho>0$ we have \begin{equation}\label{MatQ} \sum_{i\in \mathbb{N}}\sum_{j\in \mathbb{N}}\|i\|_2^{\rho-1}\|j\|_2^{\rho-1}|\langle Qe_i,e_j\rangle |< \infty. \end{equation} This is for a diagonal operator $Q$ a condition on the trace of $\Delta^{\rho-1}Q$ being finite. Using (\ref{MatQ}) together with Lemma 4 in \cite{Ref1362}, there exists a stochastic process $O:[0,T]\times \Omega\rightarrow V$, which is the Ornstein-Uhlenbeck process (or stochastic convolution) given by the semigrup generated by the Dirichlet Laplacian and the Wiener process $W(t)=\sum_{k\in\mathbb{N}} \beta^{k}(t)e_k$. Furthermore, Lemma 4 in \cite{Ref1362} assures that $O$ satisfies Assumption 3, for all $\theta \in (0,\min\{\frac{1}{2},\frac{\rho}{2}\})$ and $\gamma \in (0,\rho)$ with \begin{equation}\label{eq46} \mathbb{P}\Big[\lim_{N\rightarrow \infty}\sup_{0\leq t\leq T} \Big\|O_{t}- \sum_{i=1}^N \Big(-\lambda_i\int_0^te^{-\lambda_i(t-s)}\beta^{i}(s)ds+\beta^{i}(t)\Big)e_i\Big\|_{C^0([0,1)]}=0\Big]=1. \end{equation} Let us comment a little bit more on the $Q$-Wiener process. As $Q$ is a symmetric Hilbert-Schmidt operator, there exists an orthonormal basis $f_k$ given by eigenfunctions of $Q$ with $\alpha_k^2f_k=Qf_k$. Using standard theory of \cite{Ref1995}, there is a family of i.i.d.\ Brownian motions $\{B_k\}_{k\in\mathbb{N}}$ such that $W(t)=\sum_{ k\in\mathbb{N}}\alpha_k B_k(t) f_k \in L^2([0,1])$. We can then define \[\beta_k(t)=\langle W(t),e_k\rangle_{L^2} = \sum_{ \ell\in\mathbb{N}}\alpha_\ell B_\ell(t) \langle f_\ell,e_k\rangle_{L^2}. \] \subsection{Bounds and solutions} Let us first assume boundedness of the spectral Galerkin approximation. This will assure the existence of mild solutions later on. We will discuss later how to relax this condition to boundedness of the numerical data alone. \begin{assumption} Let $X^{N}:[0,T]\times \Omega\rightarrow V,~~N\in \mathbb{N}$, be a sequence of stochastic processes with continuous sample paths such that \begin{equation} \label{e:Galbou} \sup_{M\in \mathbb{N}}\sup_{0\leq s\leq T}\|X_s^{M}(\omega)\|_{V}<\infty \end{equation} and \begin{equation}\label{S4} X_{t}^{N}(\omega)=\int_0^{t}P_{N}S_{t-s}F(X_s^{N}(\omega))ds+P_{N}(O_{t}(\omega)), \end{equation} for every $t\in [0,T], \omega\in \Omega$ and every $N\in \mathbb{N}$. \end{assumption} From \cite{Ref1} we have the following theorem about existence of solutions. \begin{theorem}\label{theoremBJ} Let Assumptions 1-4 be fulfilled. Then, there exists a unique stochastic process $X : [0, T ]\times \Omega\rightarrow V$ with continuous sample paths, which satisfies \begin{equation}\label{eqBJ1} X_{t}(\omega)=\int_0^{t}S_{t-s}F(X_s(\omega))ds+O_{t}(\omega), \end{equation} for every $t \in [0, T ]$ and every $\omega \in \Omega$. Moreover, there exists a $\mathcal{F}/\mathcal{B}([0,\infty))$-measurable mapping $C:[0,\infty)\rightarrow \Omega$ such that \begin{equation}\label{eqBJ2} \sup_{0\leq t \leq T} \parallel X_{t}(\omega)-X_{t}^{N}(\omega)\parallel_{V} \leq C(\omega)\cdot N^{-\gamma}, \end{equation} holds for every $N \in \mathbb{N}$ and every $\omega \in \Omega$, where $\gamma \in (0,\infty)$ is given in Assumption 1 and Assumption 3. \end{theorem} \section{Time discretization} For the time discretization of the finite dimensional SDE (\ref{S4}) we follow the method proposed in \cite{Ref2}, which was also used in \cite{Ref1362}. Fix a small time-step $\Delta t>0$ and define the discrete points via the mapping $Y_{m}^{N,M}:\Omega\rightarrow V$ for $m\in \{1,...,M\}$ by \begin{equation}\label{Teq1} Y_{m+1}^{N,M}(\omega)=S_{\Delta t}\Big(Y_{m}^{N,M}(\omega) +\Delta t (P_{N}F)(Y_{m}^{N,M}(\omega))\Big) +P_{N}\Big(O_{(m+1)\Delta t}(\omega)-S_{\Delta t}O_{m\Delta t}(\omega)\Big). \end{equation} Thus $Y_{m}^{N,M}$, $m\in \{1,...,M\}$ should be the approximation of the spectral Galerkin approximation $X^N$ (see (\ref{Teqq2}) below) at times $m \cdot (\Delta t)$. For simplicity of presentation, we first assume in addition to (\ref{e:Galbou}) that our numerical data is uniformly bounded: \begin{assumption} For the numerical scheme (\ref{Teq1}) we assume \begin{equation}\label{Numeric data} \sup_{0\leq m\leq M}\sup_{N,M\in \mathbb{N}}\|Y_{m}^{N,M}\|_{V}< \infty. \end{equation} \end{assumption} Therefore, in all the examples that one wants to study, we need to verify that both bounds (\ref{Numeric data}) and (\ref{Teq1}) are true, which might be quite involved. We will comment later on the extension of the approximation result, in case either (\ref{Numeric data}) or (\ref{Teq1}) is not verified. Our aim is now to obtain the discretization error in time \begin{equation}\label{Teq2} \|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V}, \end{equation} where \begin{equation}\label{Teqq2} X_{m\Delta t}^{N}(\omega)=\int_0^{m\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds+O_{m\Delta t}^{N}(\omega) \end{equation} is the solution of the spatial discretization, which is evaluated at the grid points. It should be mentioned that for simplicity of notation, during this section $C(\omega,\alpha,\theta)>0$ is a random constant which changes from line to line. \begin{lemma}\label{lemma8} Suppose Assumptions 1-4 are true. Let $X^{N}:[0,T]\times \Omega\rightarrow V$ be the unique adapted stochastic process with continuous sample paths in (\ref{S4}) and $O^{N}:[0, T]\times \Omega \rightarrow V$ is the stochastic process defined in Assumption 4 in (\ref{eq46}). Then for all $\vartheta \in (0,1-\alpha)$, then there is a random variable $C:\Omega\rightarrow [0,\infty)$ such that \begin{equation*}\label{Teq3} \begin{split} &\Big\|(X_{t_2}^{N}(\omega)-O_{t_2}^{N}(\omega))-(X_{t_1}^{N}(\omega)-O_{t_1}^{N}(\omega))\Big\|_{V} \leq C(\omega)(t_2-t_1)^{\vartheta}, \end{split} \end{equation*} for every $N\in\mathbb{N}$, $\omega\in \Omega$ and all $t_1,t_2\in [0,T]$, with $t_1<t_2$. \end{lemma} Note that $\alpha$ was introduced in Assumption 1. \begin{proof} \begin{equation} \begin{split} &\Big\|(X_{t_2}^{N}(\omega)-O_{t_2}^{N}(\omega))-(X_{t_1}^{N}(\omega)-O_{t_1}^{N}(\omega))\Big\|_{V} \\& = \Big\|\int_{t_{1}}^{t_{2}}P_{N}S_{t_{2}-s}F(X_{s}^{N}(\omega)) ds+\int_{0}^{t_{1}}P_{N}(S_{t_{2}-s}-S_{t_{1}-s})F(X_{s}^{N}(\omega))ds\Big\|_{V} \\& \leq \int_{t_{1}}^{t_{2}}\|P_{N}S_{t_{2}-s}\|_{L(W,V)}.\|F(X_{s}^{N}(\omega)\|_{W}ds+\int_{0}^{t_{1}}\|P_{N}(S_{t_{2}-s}-S_{t_{1}-s})\|_{L(W,V)}\|F(X_{s}^{N}(\omega))\|_{W} ds \\& \leq \sup_{0\leq s\leq T} \|F(X_{s}^{N}(\omega))\|_{W} \int_{t_{1}}^{t_{2}}(t_{2}-s)^{-\alpha}ds +\int_{0}^{t_{1}}\|P_{N}S_{t_{1}-s}(S_{t_{2}-t_{1}}-I)\|_{L(W,V)}\|F(X_{s}^{N}(\omega))\|_{W}ds \\& \leq \Big(\int_{t_{1}}^{t_{2}}(t_{2}-s)^{-\alpha}ds+\int_{0}^{t_{1}}\|P_{N}S_{t_{1}-s}A^{\vartheta}\|_{L(W,V)}\|A^{-\vartheta}(S_{t_{2}-t_{1}}-I)\|_{L(V,V)}ds\Big)\sup_{0\leq s\leq T} \|F(X_{s}^{N}(\omega))\|_{W} \\& \leq C(\omega) \Big(\int_{t_{1}}^{t_{2}}(t_{2}-s)^{-\alpha}ds+\int_{0}^{t_{1}}\|P_{N}S_{\frac{t_{1}-s}{2}}\|_{L(W,V)}\|S_{\frac{t_{1}-s}{2}}A^{\vartheta}\|_{L(V,V)}\|A^{-\vartheta}(S_{t_{2}-t_{1}}-I)\|_{L(V,V)}ds\Big) \\& \leq C(\omega)\Big( (t_{2}-t_{1})^{1-\alpha}+\int_{0}^{t_{1}}(t_{1}-s)^{-\alpha-\vartheta}ds\cdot(t_{2}-t_{1})^{\vartheta}\Big) \\& \leq C(\omega)\Big((t_{2}-t_{1})^{1-\alpha}+(t_{2}-t_{1})^{\vartheta}\Big) \leq C(\omega) (t_{2}-t_{1})^{\vartheta}, \end{split} \end{equation} where we have used (\ref{e:SGtime}). \end{proof} Before we start to bound the first part of the error, we define \begin{equation*}\label{Teq8} \begin{split} R(\omega):= &\sup_{N\in \mathbb{N}}\sup_{0\leq s\leq T}\|F(X_s^{N}(\omega))\|_{W} +\sup_{N\in \mathbb{N}}\sup_{0\leq s\leq T}\|X_s^{N}(\omega)\|_{V} \\& +\sup_{0\leq t_1,t_2\leq T}\|O_{t_2}(\omega)-O_{t_1}(\omega)\|_{V}|t_2-t_1|^{-\vartheta} \\& +\sup_{N\in \mathbb{N}}\sup_{0\leq t_1,t_2\leq T} \|X_{t_2}^{N}(\omega)-O_{t_2}^{N}(\omega)-(X_{t_1}^{N}(\omega)-O_{t_1}^{N}(\omega))\|_{V}|t_2-t_1|^{-\vartheta}. \end{split} \end{equation*} provided $\vartheta \in (0,\min\{\theta,1-\alpha\})$. From Assumption 2, 4, (\ref{eq45}) and Lemma \ref{lemma8}, $R:\Omega\rightarrow \mathbb{R}$ is a finite random variable. \subsection{Theorem} The first main result of this section is stated below. \begin{theorem}\label{maintheorem} Let Assumptions 1-5 be fulfilled and suppose $\vartheta \in (0,\min\{\theta,1-\alpha\})$. Then there exists a finite random variable $C:\Omega\rightarrow [0,\infty)$ such that for all $m\in\{0,1,...,M\}$ and every $M,N\in \mathbb{N}$ \begin{equation*} \|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V} \leq C(\omega)(\Delta t)^{\vartheta}, \end{equation*} for all $\omega \in \Omega$, where $X^{N}:[0,T]\times \Omega\rightarrow V$ is the unique adapted stochastic process with continuous sample paths, defined in Assumption 4, and $Y_{m}^{N,M}:\Omega\rightarrow V$, for $m\in \{0,1,...,M\},$ and $N,M\in \mathbb{N}$, is given in (\ref{Teq1}). \end{theorem} \begin{proof} For the proof it is sufficient to prove the result for sufficiently small $|t_2-t_1|$. From (\ref{Teqq2}) we have \begin{equation}\label{Teq6} \begin{split} X_{m\Delta t}^{N}(\omega)&=\int_0^{m\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds+O_{m\Delta t}^{N}(\omega)\\ &=\sum_{k=0}^{m-1}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds+O_{m\Delta t}^{N}(\omega), \end{split} \end{equation} for every $m\in \{0,1,...,M \}$, and every $M \in \mathbb{N}$. as an intermediate discretization, we consider the mapping $Y_{m}^{N}:\Omega\rightarrow V,~m=1,2,...,M$ by \begin{equation}\label{Teq7} Y_{m}^{N}(\omega)=\sum_{k=0}^{m-1}\int_{k\Delta t}^{(k+1)\Delta t}P_{N} S_{m\Delta t-k\Delta t}F(X_{k\Delta t}^{N}(\omega))ds+O_{m\Delta t}^{N}(\omega). \end{equation} Our aim is to bound $\|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V}$. We split this into two Lemmas. First for an error between $Y_{m}^{N}$ and the spectral Galerkin method. \begin{lemma} \label{lem:bou1} Under the assumptions of Theorem \ref{maintheorem} there exists a finite random variable $C:\Omega\rightarrow [0,\infty)$ such that for all $m\in\{0,1,...,M\}$ and every $M,N\in \mathbb{N}$ \begin{equation}\label{Teq9} \|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N}(\omega)\|_{V} \leq C(\omega) (\Delta t)^\vartheta, \end{equation} for every $\omega \in \Omega$. \end{lemma} Secondly for the difference between $Y_{m}^{N}$ and the full discretization in time \begin{lemma} \label{lem:bou2} Under the assumptions of Theorem \ref{maintheorem} there exists a finite random variable $C:\Omega\rightarrow [0,\infty)$ such that for all $m\in\{0,1,...,M\}$ and every $M,N\in \mathbb{N}$ \begin{equation}\label{error2} \|Y_{m}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V} \leq C(\omega) \sum_{k=0}^{m-1} \|X_{k\Delta t}^{N}(\omega)-Y_{k}^{N,M}(\omega)\|_{V}. \end{equation} \end{lemma} \subsection{Proof of Lemma \ref{lem:bou1}} For estimating the first error term stated in (\ref{Teq9}) we have \begin{equation}\label{Teq10} \begin{split} X_{m\Delta t}^{N}(\omega)-Y_{m}^{N}(\omega)=& \sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds \\& -\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-k\Delta t}F(X_{k\Delta t}^{N}(\omega))ds \\& +\int_{(m-1)\Delta t}^{m\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds \\& -\int_{(m-1)\Delta t}^{m\Delta t}P_{N}S_{\Delta t}F(X_{k\Delta t}^{N}(\omega))ds. \end{split} \end{equation} At first we obtain the bound for the last two integrals in (\ref{Teq10}). For the first one, we get \begin{equation*}\label{Teq11} \begin{split} \Big\|\int_{(m-1)\Delta t}^{m\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds\Big\|_{V} &\leq \int_{(m-1)\Delta t}^{m\Delta t}\|P_{N}S_{m\Delta t-s}\|_{L(W,V)}\cdot\|F(X_s^{N}(\omega))\|_{W}ds\\ &\leq \int_{(m-1)\Delta t}^{m\Delta t} (m\Delta t-s)^{-\alpha}ds\sup_{0\leq s\leq t}\|F(X_s^{N}(\omega))\|_{W}\\ &\leq C R(\omega) (\Delta t)^{1-\alpha}. \end{split} \end{equation*} For the second term we obtain similarly \begin{equation*}\label{Teq12} \begin{split} \Big\|\int_{(m-1)\Delta t}^{m\Delta t}P_{N}S_{\Delta t}F(X_{k\Delta t}^{N}(\omega))ds\Big\|_{V} &\leq \sup_{0\leq s\leq t}\|F(X_s^{N}(\omega))\|_{W}\int_{(m-1)\Delta t}^{m\Delta t} (\Delta t)^{-\alpha}ds\\ &\leq R(\omega) (\Delta t)^{1-\alpha}. \end{split} \end{equation*} Therefore we have \begin{equation}\label{Teq100} \begin{split} \|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N}(\omega)\|_{V}\leq & \Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}F(X_s^{N}(\omega))ds \\& -\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-k\Delta t}F(X_{k\Delta t}^{N}(\omega))ds\Big\|_{V} \\& +R(\omega)(\Delta t)^{1-\alpha}. \end{split} \end{equation} Now we insert the OU-process. Define \[Z_{s,k\Delta t}^N(\omega)=O_s^{N}(\omega)-O_{k\Delta t}^{N}(\omega) \] Thus for every $m\in \{0,1,...,M\}$ we have, \begin{equation}\label{Serror} \begin{split} &\Big\|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N}(\omega)\Big\|_{V}\\ &\leq \Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}\Big(F(X_s^{N}(\omega))-F\big(X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega)\big)\Big)ds\Big\|_{V}\\ &+\Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}\Big(F(X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega))-F(X_{k\Delta t}^{N}(\omega))\Big)ds\Big\|_{V}\\ &+\Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}\Big(P_{N}S_{m\Delta t-s}-P_{N}S_{m\Delta t-k\Delta t}\Big)F(X_{k\Delta t}^{N}(\omega))ds\Big\|_{V}\\ &+R(\omega) (\Delta t)^{1-\alpha}.\\ \end{split} \end{equation} For the first term in (\ref{Serror}) by using (\ref{drift condi}) and Lemma \ref{lemma8} we conclude \begin{equation}\label{Teq15} \begin{split} & \Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}\Big(F(X_s^{N}(\omega))-F\big(X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega)\big)\Big)ds\Big\|_{V}\\& \leq L\sum_{k=0}^{m-2} \int_{k\Delta t}^{(k+1)\Delta t}(m\Delta t-s)^{-\alpha} \Big\|X_s^{N}(\omega)-(X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega))\Big\|_{V}\\ &\qquad\qquad\qquad\qquad\qquad\cdot\Big(1+\|X_s^{N}(\omega)\|_{V}^{p}+\|X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega)\|_{V}^{p}\Big)ds\\& \leq C(\omega)\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}(m\Delta t-s)^{-\alpha}(s-k\Delta t)^{\vartheta}\Big(1+2R^{p}(\omega)+(s-k\Delta t)^{p\vartheta}\Big)ds\\& \leq C(\omega) (\Delta t)^{\vartheta}. \end{split} \end{equation} For the second term in (\ref{Serror}) by (\ref{drift condi}) we derive \begin{equation}\label{Teq16} \begin{split} &\Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-s}\Big(F\big(X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega)\big)-F(X_{k\Delta t}^{N}(\omega))\Big)ds\Big\|_{V}\\ &\leq L\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}\Big\|P_{N}S_{m\Delta t-s}\Big\|_{L(W,V)}\Big\|X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega)-X_{k\Delta t}^{N}(\omega)\Big\|_{V}\\ &\qquad \qquad \qquad \qquad \cdot \Big(1+\Big\|X_{k\Delta t}^{N}(\omega)+Z_{s,k\Delta t}^N(\omega)\Big\|_{V}^{p}+\|X_{k\Delta t}^{N}(\omega)\|_{V}^{p}\Big) \\ &\leq C(\omega) \sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}(m\Delta t-s)^{-\alpha}(s-k\Delta t)^{\vartheta}\Big(1+2R^{p}(\omega)+(s-k\Delta t)^{p\vartheta}\Big)\\& \leq C(\omega)(\Delta t)^{\vartheta}. \end{split} \end{equation} Finally, for the third term in (\ref{Serror}) we drive \begin{equation}\label{Teq18} \begin{split} &\Big\|\sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}\left(P_{N}S_{m\Delta t-s}-P_{N}S_{m\Delta t-k\Delta t}\right)F(X_{k\Delta t}^{N}(\omega))ds\Big\|_{V}\\ & \leq \sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}\Big\|P_{N}S_{m\Delta t-k\Delta t}(S_{k\Delta t-s}-I)F(X_{k\Delta t}^{N}(\omega))\Big\|_{V}ds \\& \leq \sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}\|P_{N}S_{\frac{m\Delta t-k\Delta t}{2}}\|_{L(W,V)} \cdot\|A^{\theta}S_{\frac{m\Delta t-k\Delta t}{2}}\|_{L(V,V)} \\& \qquad \qquad \qquad \qquad \qquad \cdot\|A^{-\theta}(S_{k\Delta t-s}-I)\|_{L(V,V)} \cdot\|F(X_{k\Delta t}^{N}(\omega))\|_{W}ds\\ &\leq C(\omega) \sum_{k=0}^{m-2}\int_{k\Delta t}^{(k+1)\Delta t}(m\Delta t-k\Delta t)^{-\alpha-\theta}(k\Delta t-s)^{\theta}\|F(X_{k\Delta t}^{N}(\omega))\|_{W}ds\\& \leq C(\omega)(\Delta t)^{\theta}\leq C(\omega)(\Delta t)^{\vartheta}, \end{split} \end{equation} where we used (\ref{e:SGtime}) from the assumption on the semigroup. Hence, from (\ref{Teq15}), (\ref{Teq16}) and (\ref{Teq18}) we get \begin{equation}\label{Teq19} \begin{split} &\|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N}(\omega)\|_{V} \leq C(\omega) (\Delta t)^{\vartheta}. \end{split} \end{equation} \subsection{Proof of Lemma \ref{lem:bou2}} Now, for the second error term in (\ref{error2}) because $Y_{m}^{N,M}:\Omega\rightarrow V$ satisfies \begin{equation}\label{Teq21} Y_{m}^{N,M}(\omega)=\sum_{k=0}^{m-1}\int_{k\Delta t}^{(k+1)\Delta t}P_{N} S_{m\Delta t-k\Delta t}F(Y_{k}^{N,M}(\omega))ds+P_{N}O_{m\Delta t}(\omega). \end{equation} and by using (\ref{Numeric data}), we can estimate \begin{equation}\label{Teq22} \begin{split} &\|Y_{m}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V}=\Big\|\sum_{k=0}^{m-1}\int_{k\Delta t}^{(k+1)\Delta t}P_{N}S_{m\Delta t-k\Delta t}(F(X_{k\Delta t}^{N}(\omega))-F(Y_{k}^{N,M}(\omega)))ds\Big\|_{V}\\ &\leq L\sum_{k=0}^{m-1}\int_{k\Delta t}^{(k+1)\Delta t}(m\Delta t-k\Delta t)^{-\alpha}\Big\|X_{k\Delta t}^{N}(\omega)-Y_{k}^{N,M}(\omega)\Big\|_{V}\Big(1+\Big\|X_{k\Delta t}^{N}(\omega)\Big\|^{p}_{V}+\Big\|Y_{k}^{N,M}(\omega)\Big\|^{p}_{V}\Big)ds\\ &\leq C(\omega)\sum_{k=0}^{m-1}\Delta t(m\Delta t-k\Delta t)^{-\alpha}\Big\|X_{k\Delta t}^{N}(\omega)-Y_{k}^{N,M}(\omega)\Big\|_{V}. \end{split} \end{equation} Therefore we have \begin{equation} \|Y_{m}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V} \leq C(\omega) \sum_{k=0}^{m-1} \|X_{k\Delta t}^{N}(\omega)-Y_{k}^{N,M}(\omega)\|_{V}. \end{equation} \end{proof} Now from Lemma \ref{lem:bou1} with Lemma \ref{lem:bou2}, we get \begin{equation}\label{Teq24} \begin{split} \|X_{m\Delta t}^{N}&(\omega)-Y_{m}^{N,M}(\omega)\|_{V} \leq C\Big(R(\omega), \theta,T,L\Big)(\Delta t)^{\vartheta}\\ &+ C(\omega) \sum_{k=0}^{m-1} \|X_{k\Delta t}^{N}(\omega)-Y_{k}^{N,M}(\omega)\|_{V}. \end{split} \end{equation} By the discrete Gronwall Lemma we finally conclude \begin{equation*}\label{Teq27} \|X_{m\Delta t}^{N}(\omega)-Y_{m}^{N,M}(\omega)\|_{V} \leq C \Big(R(\omega),\theta,T,L\Big)(\Delta t)^{\vartheta}. \end{equation*} \subsection{Main results -- Full Discretization} Combining Theorem \ref{maintheorem} for the time discretization and Theorem \ref{theoremBJ} for the spatial discretization, yields the following result on the full discretization \begin{theorem}\label{maintheorem2} Let Assumptions 1-5 be true. Let $X:[0,T]\times \Omega\rightarrow V$ be the solution of the SPDE (\ref{eqBJ1}) and $Y_{m}^{N,M}:\Omega\rightarrow V$, $m\in \{0,1,...,M\}, M,N\in \mathbb{N}$ the numerical solution given by (\ref{Teq1}). Fix $\vartheta \in (0,\min\{\theta,1-\alpha\})$, then there exists a finite random variable $C:\Omega\rightarrow [0,\infty)$ such that \begin{equation}\label{Teq31} \|X_{m\Delta t}(\omega)-Y_{m}^{N,M}(\omega)\|_{V}\leq C(\omega)\left(N^{-\gamma}+(\Delta t)^{\vartheta}\right) \end{equation} for all $m\in\{0,1,...,M\}$ and every $M,N\in \mathbb{N}$. \end{theorem} For simplicity of presentation we supposed in Theorem \ref{maintheorem2} both the full discretization (\ref{Numeric data}) and the Galerkin approximation (\ref{Teq1}) to be uniformly bounded. Following the proofs, it is easy to verify that it is sufficient to assume only one of those assumptions. Let us comment in more detail on the extension of the approximation result in that case. Let us focus on the case where the uniform bound (\ref{Numeric data}) for the full discretization is not satisfied. First it is easy to verify that the following minor modification of our main result is true. Its proof follows directly, by observing, that the proof of the main theorem never uses the supremum over $M$ or $N$. \begin{theorem}\label{maintheorem_modify} Let Assumptions 1-4 be true. Fix $\vartheta \in (0,\min\{\theta,1-\alpha\})$ and fix a non-negative random constant $K(\omega)$. Let $X:[0,T]\times \Omega\rightarrow V$ be the solution of the SPDE (\ref{eqBJ1}) and $Y_{m}^{N,M}:\Omega\rightarrow V$, $m\in \{0,1,...,M\}, M,N\in \mathbb{N}$ the numerical solution given by (\ref{Teq1}). Then there exists a finite random variable $C:\Omega\rightarrow [0,\infty)$, depending on $K$, but independent of $M$ and $N$, such that the following is true: If for one choice of $N,M\in\mathbb{N}$ we have \begin{equation} \sup_{0\leq m\leq M} \|Y_{m}^{N,M}\|_{V}\leq K(\omega) \end{equation} then \begin{equation} \|X_{m\Delta t}(\omega)-Y_{m}^{N,M}(\omega)\|_{V}\leq C(\omega)\left(N^{-\gamma}+(\Delta t)^{\vartheta}\right) \end{equation} for all $m\in\{0,1,...,M\}$. \end{theorem} To proceed, note that in the proofs we can always bound every occurrence of $\|Y_{m}^{N,M}\|^p_{V}$ by the bounded $\|X_{t}^{N}\|_{V}$ and the error \[e^{N,M}_t=\sup_{m\Delta t \leq t} \|Y_{m}^{N,M} - X_{m \Delta t}^{N}\|_{V}\;. \] I.e., we use for $m \leq t/\Delta t$ \[\|Y_{m}^{N,M}\|^p_{V} \leq C_p \|X_{m\Delta t}^{N}\|_{V}^p + C_p (e^{N,M}_t)^p\;. \] If we now assume a-priori that $e^{N,M}_t \leq 1$, which is easily true, for sufficiently small $t>0$, then we can proceed completely analogous as in the proofs of Theorem \ref{maintheorem2}. By Theorem \ref{maintheorem_modify} this implies now that for the error, probably with a different $C(\omega)$ that \[e^{N,M}_t \leq C(\omega) (\Delta t)^\vartheta \;. \] As the right hand-side is independent of $M$ and $N$, we can a-posteriori conclude, that as long as $C(\omega) (\Delta t)^\vartheta \leq 1$ our initial guess on $e^{N,M}$ was true, and we finally derive the following theorem: \begin{theorem}\label{maintheorem_modify2} Let Assumptions 1-4 be true. Let $X:[0,T]\times \Omega\rightarrow V$ be the solution of the SPDE (\ref{eqBJ1}) and $Y_{m}^{N,M}:\Omega\rightarrow V$, $m\in \{0,1,...,M\}, M,N\in \mathbb{N}$ the numerical solution given by (\ref{Teq1}). Then there exists a finite random variable $C:\Omega\rightarrow [0,\infty)$ such that the error estimate (\ref{Teq31}) holds provided $0<\Delta t< C(\omega)^{-\vartheta}$. \end{theorem} The case when the uniform bound (\ref{e:Galbou}) on the spectral Galerkin approximation fails, is verified in a similar way, by bounding in the whole proof $\|X_{t}^{N}\|_{V}^p$ by $\|Y_{m}^{N,M}\|^p_{V}$ and $e^{N,M}_t$. \section{Numerical results} In this section we consider examples for the numerical solution of stochastic equation by the method given in (\ref{Teq1}). Let $ V = W = C([0,\pi])$ be the $\mathbb{R}$-Banach space of continuous functions from $[0,\pi]$ to $\mathbb{R}$ equipped with the norm $\|v\|_{V}=\|v\|_{W}:= \sup_{x\in [0,\pi]} |v(x)|$ for every $v\in V=W$ , where $|.|$ is the absolute value of a real number. Moreover, consider as orthonormal $L^2$-basis the smooth eigenfunctions \begin{equation*} e_{k}:[0,\pi]\rightarrow \mathbb{R},~~~ e_{k}(x)=\sqrt{2/\pi}\sin(kx), \qquad\text{for every } x\in (0,\pi). \end{equation*} Denote the Laplacian with Dirichlet boundary conditions on $[0,\pi]$ by $A$, such that $Ae_k = - k^2e_k$. Moreover, define the operators $P_N$ as the $L^2$-orthogonal projections onto the span of the first $N$ eigenfunctions $e_k$. We define the mapping $S:[0,T]\rightarrow L(V)$ by \begin{equation} (S_{t})v(x)=\sum_{i\in \mathbb{N}}e^{-\lambda_{i}t}\int e_{i}(s) v(s)ds\cdot e_{i}(x), \end{equation} where $\lambda_{i}=-i^{2}$. It is well known that $A$ generates the analytic semigroup $(S_t)_{t\geq0}$ on $V$. See \cite{Ref2013}. From Lemma 4.1 in \cite{Ref1} and Lemma 1 in \cite{Ref1362} we recall that (\ref{S1}) is satisfied for $\gamma \in (0,\frac{3}{2})$ and $\alpha \in (\frac{1}{4}+\frac{\gamma}{2},1)$. Moreover, from \cite{Ref2013} we know that (\ref{e:SGtime}) is satisfied for any $\theta\in (0,1)$.\\ Assume that the OU-process $O:[0,T]\times \Omega\rightarrow V$ is as defined in the example in Section \ref{suse:OU}. Therefore $O$ satisfies Assumption 3, for all $\theta \in (0,\min\{\frac{1}{2},\frac{\rho}{2}\})$ and $\gamma \in (0,\rho)$. The covariance operator $Q$ is given as a convolution operator \begin{equation}\label{MathQ} \langle Q e_{k},e_{l} \rangle =\int_0^{\pi}\int_0^{\pi}e_{k}(x)e_{l}(y)q(x-y)dy dx. \end{equation} We obtain our numerical result with two kernels \begin{equation}\label{cov1exam1} q_1(x-y)=\frac1h\max\{0,1- \frac1{h^2} |x-y|\} \end{equation} and \begin{equation}\label{cov2exam1} q_2(x-y)=\max\{0,1 - \frac1h |x-y|\}. \end{equation} In Figures \ref{CovMatrix1}, \ref{CovMatrix2} we plotted the Covariance Matrix for $h=0.1$ and $h=0.01$ with kernel (\ref{cov1exam1}) and kernel (\ref{cov2exam1}). By some numerical calculations we can show that the condition on $Q$ from (\ref{MatQ}) is satisfied for any $\rho\in (0,\frac{1}{2})$, as we can calculate explicitly the Fourier-coefficients of $(x,y)\mapsto q(x-y)$ and check for summability. For simplicity fix the smooth deterministic initial condition \[ \xi(x)=\frac{\sin x}{\sqrt{2}}+\frac{3\sqrt{2}}{5}\sin(3x), \quad\text{ for all }x\in [0,\pi]. \] Now we consider two types of nonlinearity, globally Lipschitz and locally Lipschitz, as given by the following examples. \\ \textbf{Example 1} Consider for the nonlinearity the Nemytskii operator $F:V\rightarrow V$ given by $(F(v))(x) = f(v(x))$ for every $x \in [0,\pi]$ and every $v \in V$, where $f: \mathbb{R}\rightarrow \mathbb{R}$ is given by \begin{equation} f(y)=5\cdot\frac{1-y}{1+y^{2}}. \end{equation} This generates a globally Lipschitz nonlinearity. Thus Assumption 2 is true.\\ The stochastic equation (\ref{intr1}) now reads as \begin{equation}\label{eq4exam1} dX_{t}=\Big[\frac{\partial ^{2}}{\partial x^{2}}X_{t}+5\frac{1-X_{t}}{1+X_{t}^{2}}\Big]dt+dW_{t}, ~~X_0(x)=\frac{\sin x}{\sqrt{2}}+\frac{3\sqrt{2}}{5}\sin(3x), \end{equation} with Dirichlet boundary conditions $X_{t}(0)=X_{t}(\pi)=0$ for $t\in [0,1] $. The finite dimensional SDE (\ref{S4}) reduces to \begin{equation} dX_{t}^{N} =\Big[\frac{\partial ^{2}}{\partial x^{2}}X_{t}^{N} +5P_{N}\frac{1-X_{t}^{N}}{1+(X_{t}^{N})^{2}}\Big]dt +d P_{N}W_{t},~~X_0^{N}(x)=\frac{\sin x}{\sqrt{2}}+\frac{3\sqrt{2}}{5}\sin(3x), \end{equation} with $X_{t}^{N}(0)=X_{t}^{N}(\pi)=0$ for $t\in [0,1]$ and $x\in [0,\pi]$, and all $N \in \mathbb{N}.$ Now in our simple example we can verify rigorously that the numerical data is uniformly bounded. We derive \begin{equation} \begin{split} \|X_{t}^{N}(\omega)\|_{V}=&\Big\|\int_0^{t}P_{N}S_{t-s}F(X_s^{N}(\omega))ds+P_{N}(O_{t}(\omega))\Big\|_{V}\\ & \leq \int_0^{t}\|P_{N}S_{t-s}\|_{L(V,V)}\|F(X_s^{N}(\omega))\|_{V}ds+\|P_{N}(O_{t}(\omega))\|_{V}\\ & \leq C \int_0^{t}\|P_{N}S_{t-s}\|_{L(V,V)}(1+\|X_s^{N}(\omega)\|_{V})ds+C, \end{split} \end{equation} from Gronwall inequality we conclude \begin{equation} \sup_{N\in \mathbb{N}}\sup_{0\leq s \leq T}\|X_{s}^{N}(\omega)\|_{V} <\infty. \end{equation} In a similar way we conclude \begin{equation} \sup_{N,M\in \mathbb{N}}\sup_{0\leq m \leq M}\|Y_{m}^{N,M}(\omega)\|_{V} <\infty. \end{equation} Theorem \ref{maintheorem2} by using this fact that here $\theta \in (0,\min\{\frac{1}{2},\frac{\rho}{2}\})$, yields the existence of a unique solution $X:[0,\pi]\times \Omega \rightarrow C^{0}([0,\pi])$ of the SPDE (\ref{eq4exam1}) such that \begin{equation} \sup_{0\leq x \leq \pi}|X_{m\Delta t}(\omega,x)-Y_{m}^{N,M}(\omega,x)|\leq C(\omega)\left(N^{-\gamma}+(\Delta t)^{\vartheta}\right) \end{equation} for $m=1,...,M,~~ M=\frac{1}{\Delta t}$, such that $\gamma \in (0,\frac{1}{2})$, $\vartheta \in(0,\frac{1}{4})$.\\ \textbf{Example 2} Consider for the nonlinearity the Nemytskii operator $F:V\rightarrow V$ given by $(F(v))(x) = f(v(x))$ for every $x \in [0,\pi]$ and every $v \in V$, where $f:\mathbb{R}\rightarrow \mathbb{R}$ is given by \begin{equation} f(y)=-y^{3}. \end{equation} This generates a locally Lipschitz nonlinearity which satisfies Assumption \ref{ass:nonlin}. The stochastic equation (\ref{intr1}) now reads as \begin{equation}\label{eq4exam2} dX_{t}=\Big[\frac{\partial ^{2}}{\partial x^{2}}X_{t}-X_{t}^{3}\Big]dt+dW_{t}, ~~X_0(x)=\frac{\sin x}{\sqrt{2}}+\frac{3\sqrt{2}}{5}\sin(3x), \end{equation} with Dirichlet boundary conditions $X_{t}(0)=X_{t}(\pi)=0$ for $t\in [0,1] $. The finite dimensional SDE (\ref{S4}) reduces to \begin{equation}\label{eq5exam1} dX_{t}^{N} =\Big[\frac{\partial ^{2}}{\partial x^{2}}X_{t}^{N} -P_{N}(X_{t}^{N})^{3}\Big]dt +d P_{N}W_{t},~~X_0^{N}(x)=\frac{\sin x}{\sqrt{2}}+\frac{3\sqrt{2}}{5}\sin(3x), \end{equation} with $X_{t}^{N}(0)=X_{t}^{N}(\pi)=0$ for $t\in [0,1]$ and $x\in [0,\pi]$, and all $N \in \mathbb{N}.$ Now by using Theorem \ref{maintheorem_modify2} it remains to verify (\ref{e:Galbou}) from Assumption 4. This is straightforward by using first the estimates in $L^2$, and we sketch only the main ideas here. Define $y^N_t=X_t^N-P_{N}O(t)$. Thus \[ \partial_t y^N_t = \frac{\partial ^{2}}{\partial x^{2}}y^N_t -(y^N_t +P_{N}O_t)^{3}\;. \] Hence, \begin{equation} \tfrac12\partial_t\|y_{t}^{N}\|^2_{L^2} = -\| \partial_x y_{t}^{N}\|^2 - \tfrac12 \|y_{t}^{N}\|^4_{L^4} + C \|P_{N}O_t\|^4_{L^4}\;. \end{equation} This gives a random bound in $L^2([0,T], H^1) \cap L^\infty([0,T], L^2)\cap L^4([0,T], L^4)$. Using Agmon inequality yields \[ \|y^N_t\|_{L^4([0,T],V)}^2 \leq C \|y^N_t\|_{L^2([0,T],H^1)} \cdot \|y^N_t\|_{L^\infty([0,T],L^2)}\;. \] This is sufficient to verify the bound in $V$ from the mild formulation, as \[ \|y^N_t\|_{V} \leq C \int_0^{t}\|P_{N}S_{t-s}\|_{L(V,V)}\cdot \|y^N_s+P_N O_s\|_{V}^3 ds\;. \] Now from our main results for the unique solution $X:[0,\pi]\times \Omega \rightarrow C^{0}([0,\pi])$ of the SPDE (\ref{eq4exam2}) we obtain for sufficiently small $\Delta t$ \begin{equation} \sup_{0\leq x \leq \pi}|X_{m\Delta t}(\omega,x)-Y_{m}^{N,M}(\omega,x)|\leq C(\omega)\left(N^{-\gamma}+(\Delta t)^{\vartheta}\right) \end{equation} for $m=1,...,M,~~ M=\frac{1}{\Delta t}$, such that $\gamma \in (0,\frac{1}{2})$, $\vartheta \in(0,\frac{1}{4})$.\\ Let us now explain briefly how we implement our numerical results. The main part is generating the Brownian motions $X=(X_1,X_2,\cdots,X_N)$ that are correlated such that $X\sim \textit{N}(0,\Sigma)$, which $Cov(X_i,X_j)=\Sigma_{ij}$. For this assume $C$ is a $n\times m$ Matrix and let $Z=(Z_{1},\cdots,Z_{N})^{T}$, with $Z_{i}\sim \textit{N}(0,1)$, for $i=1,\cdots,N.$ Then obviously $C^{T}Z\sim N(0,C^{T}C)$. Therefore our aim clearly reduces to finding $C$ such that $C^{T}C=\Sigma$, which can for instance be achieved by Cholesky. By using $\Delta t=\frac{T}{N^{2}}$, the solutions $X_{t}^{N}(\omega,x)$ of the finite dimensional SODEs (\ref{eq5exam1}) converge uniformly in $t\in [0,1]$ and $x\in [0,\pi]$ to the solution $X_{t}(\omega,x)$ of the stochastic evolution equation (\ref{eq4exam2}) with the rate $\frac{1}{2}$, as $N$ goes to infinity for all $\omega \in \Omega.$ In Figure \ref{Figpatherr} the path-wise approximation error \begin{equation}\label{patherr} \sup_{0\leq x \leq \pi}\sup_{0\leq m \leq M}|X_{m\Delta t}(\omega,x)-Y_{m}^{N,M}(\omega,x)|\ \end{equation} is plotted against $N$, for $N \in \{16,32,\cdots,256\}$. As a replacement for the true unknown solution, we use a numerical approximation for $N$ sufficiently large. Figure \ref{Figpatherr} confirms that, as we expected from Theorem \ref{maintheorem_modify2}, the order of convergence is $\frac{1}{2}.$ Obviously, these are only two examples, but all out of a few hundred calculated examples behave similarly. Even their mean seem to behave with the same order of the error. Nevertheless, we did not calculate sufficiently many realizations to estimate the mean satisfactory, nor did we proof in the general setting, that the mean converges. Finally, as an example in Figures \ref{Evolution}, $X_{t}(\omega)$, are plotted for $t \in [0,T]$ for $T\in \{\frac{3}{200},0.2,1\},$ for $h=0.1$, with convolution operator (\ref{MathQ}) with kernel (\ref{cov1exam1}) and (\ref{cov2exam1}). \begin{figure \centering \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{q1h01.png}} \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{q2h01.png}} \hfill % \caption[Covariance Matrix ]{Covariance Matrix $<Qe_{k},e_{l}>_{k,l}$ for $k,l \in \{1,2,\cdots,100\}$, for $h=0.1$ by (a) kernel (\ref{cov1exam1}) and (b) kernel (\ref{cov2exam1})} \label{CovMatrix1} \end{figure} \begin{figure \centering \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{q1h001.png}} \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{q2h001.png}} \hfill % \caption[Covariance Matrix ]{Covariance Matrix $<Qe_{k},e_{l}>_{k,l}$ for $k,l \in \{1,2,\cdots,100\}$, for $h=0.01$ (a) kernel (\ref{cov1exam1}) and (b) kernel (\ref{cov2exam1})} \label{CovMatrix2} \end{figure} \begin{figure \centering \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{Patherr1.png}} \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{Patherr2.png}} \hfill % \caption[Pathwise approximation error]{Pathwise approximation error (\ref{patherr}) against $N$ for $N \in \{16,32,...,256\}$ with convolution operator with kernel (\ref{cov1exam1}) for (a) $h=0.1$ and (b) $h=0.01$, for one random $\omega \in \Omega$.} \label{Figpatherr} \end{figure} \begin{figure \centering \hfill % {\includegraphics[width=0.48\textwidth]{Evt3200h01q1.png}} \hfill % {\includegraphics[width=0.48\textwidth]{Evt3200h01q2.png}} \hfill % \end{figure} \begin{figure \centering \hfill % {\includegraphics[width=0.48\textwidth]{Evt02h01q1.png}} \hfill % {\includegraphics[width=0.48\textwidth]{Evt02h01q2.png}} \hfill % \end{figure} \begin{figure \centering \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{Evt1h01q1.png}} \hfill % \subfloat[]{\includegraphics[width=0.48\textwidth]{Evt1h01q2.png}} \hfill % \caption[Stochastic Evolution equation]{ $X_{t}(\omega,x)$, $ x \in [0,\pi],~t\in (0,T)$ for $T\in \{3/200,0.2,1\}$, given by (\ref{eq4exam1}) for $h=0.1$ with the covariance operator by (a) kernel (\ref{cov1exam1}) and (b) kernel (\ref{cov2exam1}), for one random $\omega \in \Omega$.} \label{Evolution} \end{figure}
\section{Zero-sum Game and Uncoupled Dynamics} The notions of a zero-sum matrix game and a minimax equilibrium are arguably the most basic and important notions of game theory. The tight connection between linear programming and minimax equilibrium suggests that there might be simple dynamics that can lead the two players of the game to eventually converge to the equilibrium value. Existence of such simple or natural dynamics is of interest in behavioral economics, where one asks whether agents can discover static solution concepts of the game iteratively and without extensive communication. More formally, let $A\in [-1,1]^{n\times m}$ be a matrix with bounded entries. The two players aim to find a pair of near-optimal mixed strategies $(\bar{f},\bar{x})\in \Delta_n \times \Delta_m$ such that $\bar{f}^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A \bar{x}$ is close to the minimax value $\min_{f\in\Delta_n} \max_{x\in\Delta_m} f^\ensuremath{{\scriptscriptstyle\mathsf{T}}} Ax$, where $\Delta_n$ is the probability simplex over $n$ actions. Of course, this is a particular form of the saddle point problem considered in the previous section, with $\phi(f,x) = f^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A x$. It is well-known (and follows immediately from \eqref{eq:sandwich_minimax}) that the players can compute near-optimal strategies by simply playing no-regret algorithms \cite{freund1999adaptive}. More precisely, on round $t$, the players I and II ``predict'' the mixed strategies $f_t$ and $x_t$ and observe $Ax_t$ and $f_t^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A$, respectively. While black-box regret minimization algorithms, such as Exponential Weights, immediately yield $\mathcal{O}(T^{-1/2})$ convergence rates, Daskalakis et al \cite{daskalakis2011near} asked whether faster methods exist. To make the problem well-posed, it is required that the two players are \emph{strongly uncoupled}: neither $A$ nor the number of available actions of the opponent is known to either player, no ``funny bit arithmetic'' is allowed, and memory storage of each player allows only for constant number of payoff vectors. The authors of \cite{daskalakis2011near} exhibited a near-optimal algorithm that, if used by both players, yields a pair of mixed strategies that constitutes an $\mathcal{O}\left(\frac{\log(m+n)(\log T + (\log(m+n))^{3/2})}{T}\right)$-approximate minimax equilibrium. Furthermore, the method has a regret bound of the same order as Exponential Weights when faced with an arbitrary sequence. The algorithm in \cite{daskalakis2011near} is an application of the excessive gap technique of Nesterov, and requires careful choreography and interleaving of rounds between the two non-communicating players. The authors, therefore, asked whether a simple algorithm (e.g. a modification of Exponential Weights) can in fact achieve the same result. We answer this in the affirmative. While a direct application of Mirror Prox does not yield the result (and also does not provide strong decoupling), below we show that a modification of Optimistic Mirror Descent achieves the goal. Furthermore, by choosing the step size adaptively, the same method guarantees the typical $\mathcal{O}(T^{-1/2})$ regret if not faced with a compliant player, thus ensuring robustness. In Section~\ref{sec:full_info}, we analyze the ``first-order information'' version of the problem, as described above: upon playing the respective mixed strategies $f_t$ and $x_t$ on round $t$, Player I observes $Ax_t$ and Player II observes $f_t^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A$. Then, in Section~\ref{sec:partial_info}, we consider an interesting extension to partial information, whereby the players submit their moves $f_t, x_t$ but only observe the real value $f_t^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A x_t$. Recall that in both cases the matrix $A$ is not known to the players. \subsection{First-Order Information} \label{sec:full_info} Consider the following simple algorithm. Initialize $f_0=g'_0\in \Delta_n$ and $x_0=y'_0\in\Delta_m$ to be uniform distributions, set $\beta= 1/T^2$ and proceed as follows: \begin{center} \framebox[.99\columnwidth][c]{ \begin{minipage}{.95\columnwidth} {\tt On round $t$, Player I performs \begin{align*} &\text{Play} ~~~~~ f_t \text{ and observe } A x_t\\ &\text{Update} ~~~ g_{t}(i) \propto g'_{t-1}(i)\exp\{-\eta_{t} [Ax_{t}]_i \}, ~~~ g'_{t} = \left(1-\beta\right) g_{t} + \left(\beta/n\right){\mathbf 1}_n\\ & ~~~~~~~~~ f_{t+1} (i) \propto g'_{t}(i)\exp\{-\eta_{t+1} [Ax_{t}]_i \} \end{align*} while simultaneously Player II performs \begin{align*} &\text{Play} ~~~~~ x_t \text{ and observe } f_t^\top A \\ &\text{Update} ~~~ y_{t}(i) \propto y'_{t-1}(i)\exp\{-\eta'_{t} [f_{t}^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A]_i \}, ~~~ y'_{t} = \left(1-\beta\right) y_{t} + \left(\beta/m\right){\mathbf 1}_m\\ & ~~~~~~~~~x_{t+1} (i) \propto y'_{t}(i)\exp\{-\eta'_{t+1} [f_{t}^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A]_i \} \end{align*} } \vspace{-4mm} \end{minipage} } \end{center} Here, ${\mathbf 1}_n\in \mathbb{R}^n$ is a vector of all ones and both $[b]_i$ and $b(i)$ refer to the $i$-th coordinate of a vector $b$. Other than the ``mixing in'' of the uniform distribution, the algorithm for both players is simply the Optimistic Mirror Descent with the (negative) entropy function. In fact, the step of mixing in the uniform distribution is only needed when some coordinate of $g_t$ (resp., $y_t$) is smaller than $1/(nT^2)$. Furthermore, this step is also not needed if none of the players deviate from the prescribed method. In such a case, the resulting algorithm is simply the constant step-size Exponential Weights $f_{t} (i) \propto \exp\{-\eta \sum_{s=1}^{t-2} [Ax_{s-1}]_i + 2\eta[Ax_{t-1}]_i \}$, but with a factor $2$ in front of the \emph{latest} loss vector! \begin{proposition} \label{prop:bilinear_optimal} Let $A\in [-1,1]^{n\times m}$, $\F=\Delta_n$, $\cX=\Delta_m$. If both players use above algorithm with, respectively, $M_t^1 = Ax_{t-1}$ and $M_t^2 = f_{t-1}^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A$, and the adaptive step sizes $$\textstyle \eta_t = \min\left\{ \log(nT)\left(\sqrt{\sum_{i=1}^{t-1} \norm{Ax_i - Ax_{i-1}}_*^2} + \sqrt{\sum_{i=1}^{t-2} \norm{Ax_i - Ax_{i-1}}_*^2} \right)^{-1} , \frac{1}{11}\right\} $$ and $$\textstyle \eta'_t = \min\left\{ \log(mT)\left(\sqrt{\sum_{i=1}^{t-1} \norm{f_i^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A - f_{i-1}^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A}_*^2} + \sqrt{\sum_{i=1}^{t-2} \norm{f_i^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A - f_{i-1}^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A}_*^2} \right)^{-1} , \frac{1}{11}\right\} $$ respectively, then the pair $(\bar{f}_T,\bar{x}_T)$ is an $O\left(\frac{\log m +\log n+ \log T}{T}\right)$-approximate minimax equilibrium. Furthermore, if only one player (say, Player I) follows the above algorithm, her regret against any sequence $x_1,\ldots,x_T$ of plays is \begin{align} \label{eq:intermediate_regret_game} \mathcal{O}\left(\frac{\log (nT)}{T}\left(\sqrt{\sum_{t=1}^T \norm{Ax_t - Ax_{t-1}}_*^2} + 1\right)\right) \ . \end{align} In particular, this implies the worst-case regret of $\mathcal{O}\left( \frac{\log(nT)}{\sqrt{T}}\right)$ in the general setting of online linear optimization. \end{proposition} We remark that \eqref{eq:intermediate_regret_game} can give intermediate rates for regret in the case that the second player deviates from the prescribed strategy but produces ``stable'' moves. For instance, if the second player employs a mirror descent algorithm (or Follow the Regularized Leader / Exponential Weights method) with step size $\eta$, one can typically show stability $\|x_t-x_{t-1}\|= \mathcal{O}(\eta)$. In this case, \eqref{eq:intermediate_regret_game} yields the rate $\mathcal{O}\left(\frac{\eta\log T}{\sqrt{T}}\right)$ for the first player. A typical setting of $\eta\propto T^{-1/2}$ for the second player still ensures the $\mathcal{O}(\log T/T)$ regret for the first player. Let us finish with a technical remark. The reason for the extra step of ``mixing in'' the uniform distribution stems from the goal of having an adaptive and robust method that still attains $\mathcal{O}(T^{-1/2})$ regret if the other player deviates from using the algorithm. If one is only interested in the dynamics when both players cooperate, this step is not necessary, and in this case the extraneous $\log T$ factor disappears from the above bound, leading to the $O\left(\frac{\log n + \log m}{T}\right)$ convergence. On the technical side, the need for the extra step is the following. The adaptive step size result of Corollary~\ref{cor:adaptive_step_size} involves the term $R_{\max}^2 \geq \sup_{g} \cD_{\cR_{1}}(f^*,g)$ which is potentially infinite for the negative entropy function $\cR_1$. It is possible that the doubling trick or the analysis of Auer et al \cite{auer2002adaptive} (who encountered the same problem for the Exponential Weights algorithm) can remove the extra $\log T$ factor while still preserving the regret minimization property. We also remark that $R_{\max}$ is small when $\cR_1$ is instead the $p$-norm; hence, the use of this regularizer avoids the extraneous logarithmic in $T$ factor while still preserving the logarithmic dependence on $n$ and $m$. However, projection onto the simplex under the $p$-norm is not as elegant as the Exponential Weights update. \input{partial} \section{Introduction} Recently, no-regret algorithms have received increasing attention in a variety of communities, including theoretical computer science, optimization, and game theory \cite{PLG,arora2012multiplicative}. The wide applicability of these algorithms is arguably due to the black-box regret guarantees that hold for arbitrary sequences. However, such regret guarantees can be loose if the sequence being encountered is not ``worst-case''. The reduction in ``arbitrariness'' of the sequence can arise from the particular structure of the problem at hand, and should be exploited. For instance, in some applications of online methods, the sequence comes from an additional computation done by the learner, thus being far from arbitrary. One way to formally capture the partially benign nature of data is through a notion of predictable sequences \cite{RakSri13pred}. We exhibit applications of this idea in several domains. First, we show that the Mirror Prox method \cite{nemirovski2004prox}, designed for optimizing non-smooth structured saddle-point problems, can be viewed as an instance of the predictable sequence approach. Predictability in this case is due precisely to smoothness of the inner optimization part and the saddle-point structure of the problem. We extend the results to H\"older-smooth functions, interpolating between the case of well-predictable gradients and ``unpredictable'' gradients. Second, we address the question raised in \cite{daskalakis2011near} about existence of ``simple'' algorithms that converge at the rate of $\tilde{\mathcal O}(T^{-1})$ when employed in an uncoupled manner by players in a zero-sum finite matrix game, yet maintain the usual $\mathcal{O}(T^{-1/2})$ rate against arbitrary sequences. We give a positive answer and exhibit a fully adaptive algorithm that does not require the prior knowledge of whether the other player is collaborating. Here, the additional predictability comes from the fact that both players attempt to converge to the minimax value. We also tackle a partial information version of the problem where the player has only access to the real-valued payoff of the mixed actions played by the two players on each round rather than the entire vector. Our third application is to convex programming: optimization of a linear function subject to convex constraints. This problem often arises in theoretical computer science, and we show that the idea of predictable sequences can be used here too. We provide a simple algorithm for $\epsilon$-approximate Max Flow for a graph with $d$ edges with time complexity $\tilde{\mathcal{O}}(d^{3/2}/\epsilon)$, a performance previously obtained through a relatively involved procedure \cite{goldberg1998beyond}. \section{Approximate Smooth Convex Programming} In this section we show how one can use the structured optimization results from Section \ref{sec:structopt} for approximately solving convex programming problems. Specifically consider the optimization problem \begin{align} \argmax{f \in \G} ~~~&~~~ c^\top f \label{eq:linopt}\\ \textrm{s.t. } ~~~&~~~\forall i \in [d], ~~G_i(f) \le 1\notag \end{align} where $\G$ is a convex set and each $G_i$ is an $H$-smooth convex function. Let the optimal value of the above optimization problem be given by $F^* > 0$, and without loss of generality assume $F^*$ is known (one typically performs binary search if it is not known). Define the sets $\F = \{f : f \in \G , c^\top f = F^*\}$ and $\cX = \Delta_d$. The convex programming problem in \eqref{eq:linopt} can now be reformulated as \begin{align}\label{eq:stroptcvx} \argmin{f \in \F} \max_{i \in [d]} G_i(f) = \argmin{f \in \F} \sup_{x \in \cX} \sum_{i=1}^d x(i) G_i(f) \ . \end{align} This problem is in the saddle-point form, as studied earlier in the paper. We may think of the first player as aiming to minimize the above expression over $\F$, while the second player maximizes over a mixture of constraints with the aim of violating at least one of them. \begin{lemma}\label{lem:cvxprog} Fix $\gamma, \epsilon>0$. Assume there exists $f_0 \in \G$ such that $c^\top f_0 \ge 0$ and for every $i \in [d]$, $G_i(f_0) \le 1 - \gamma$. Suppose each $G_i$ is $1$-Lipschitz over $\F$. Consider the solution $$ \hat{f}_T = (1 - \alpha) \bar{f}_T + \alpha f_0 $$ where $\alpha = \frac{\epsilon}{\epsilon + \gamma}$ and $\bar{f}_T = \frac{1}{T}\sum_{t=1}^T f_t \in \F$ is the average of the trajectory of the procedure in Lemma~\ref{lem:saddle_predictable} for the optimization problem \eqref{eq:stroptcvx}. Let $\cR_1(\cdot) = \frac{1}{2}\norm{\cdot}_2^2$ and $\cR_2$ be the entropy function. Further let $B$ be a known constant such that $B \ge \norm{f^* - g_0}_2$ where $g_0 \in \F$ is some initialization and $f^* \in \F$ is the (unknown) solution to the optimization problem. Set $\eta = \argmin{\eta \le H^{-1}}\left\{ \frac{ B^2}{\eta} + \frac{\eta \log d}{1- \eta H} \right\}$, $\eta' = \frac{1}{\eta} - H$, $M^1_t = \sum_{i=1}^d y_{t-1}(i) \nabla G_i(g_{t-1})$ and $M^2_t = (G_1(g_{t-1}) , \ldots,G_{d}(g_{t-1}))$. Let number of iterations $T$ be such that $$ T > \frac{1}{\epsilon} \inf_{\eta \le H^{-1}}\left\{ \frac{ B^2}{\eta} + \frac{\eta \log d}{1- \eta H} \right\ $$ We then have that $\hat{f}_T \in \G$ satisfies all $d$ constraints and is $\frac{\epsilon}{\gamma}$-approximate, that is $$ c^\top \hat{f}_T \ge \left(1 - \frac{\epsilon}{\gamma}\right)F^* \ . $$ \end{lemma} Lemma~\ref{lem:cvxprog} tells us that using the predictable sequences approach for the two players, one can obtain an $\frac{\epsilon}{\gamma}$-approximate solution to the smooth convex programming problem in number of iterations at most order $1/\epsilon$. If $T_1$ (reps. $T_2$) is the time complexity for single update of the predictable sequence algorithm of Player I (resp. Player 2), then time complexity of the overall procedure is $\mathcal{O}\left(\frac{T_1 + T_2}{\epsilon}\right)$ \subsection{Application to Max-Flow} We now apply the above result to the problem of finding Max Flow between a source and a sink in a network, such that the capacity constraint on each edge is satisfied. For simplicity, consider a network where each edge has capacity $1$ (the method can be easily extended to the case of varying capacity). Suppose the number of edges $d$ in the network is the same order as number of vertices in the network. The Max Flow problem can be seen as an instance of a convex (linear) programming problem, and we apply the proposed algorithm for structured optimization to obtain an approximate solution. For the Max Flow problem, the sets $\G$ and $\F$ are given by sets of linear equalities. Further, if we use Euclidean norm squared as regularizer for the flow player, then projection step can be performed in $\mathcal{O}(d)$ time using conjugate gradient method. This is because we are simply minimizing Euclidean norm squared subject to equality constraints which is well conditioned. Hence $T_1= \mathcal{O}(d)$. Similarly, the Exponential Weights update has time complexity $\mathcal{O}(d)$ as there are order $d$ constraints, and so overall time complexity to produce $\epsilon$ approximate solution is given by $\mathcal{O}(n d )$, where $n$ is the number of iterations of the proposed procedure. Once again, we shall assume that we know the value of the maximum flow $F^*$ (for, otherwise, we can use binary search to obtain it). \begin{corollary}\label{cor:maxflow} Applying the procedure for smooth convex programming from Lemma \ref{lem:cvxprog} to the Max Flow problem with $f_0 = \mbf{0} \in \G$ the $0$ flow, the time complexity to compute an $\epsilon$-approximate Max Flow is bounded by $$ \mathcal{O}\left(\frac{d^{3/2} \sqrt{\log d}}{\epsilon} \right) \ . $$ \end{corollary} This time complexity matches the known result from \cite{goldberg1998beyond}, but with a much simpler procedure (gradient descent for the flow player and Exponential Weights for the constraints). It would be interesting to see whether the techniques presented here can be used to improve the dependence on $d$ to $d^{4/3}$ or better while maintaining the $1/\epsilon$ dependence. While the result of \cite{christiano2011electrical} has the improved $d^{4/3}$ dependence, the complexity in terms of $\epsilon$ is much worse. \section{Discussion} We close this paper with a discussion. As we showed, the notion of using extra information about the sequence is a powerful tool with applications in optimization, convex programming, game theory, to name a few. All the applications considered in this paper, however, used some notion of smoothness for constructing the predictable process $M_t$. An interesting direction of further research is to isolate more general conditions under which the next gradient is predictable, perhaps even when the functions are not smooth in any sense. For instance one could use techniques from bundle methods to further restrict the set of possible gradients the function being optimized can have at various points in the feasible set. This could then be used to solve for the right predictable sequence to use so as to optimize the bounds. Using this notion of selecting predictable sequences one can hope to derive adaptive optimization procedures that in practice can provide rapid convergence. \vspace{3mm} {\bf Acknowledgements:} We thank Vianney Perchet for insightful discussions. We gratefully acknowledge the support of NSF under grants CAREER DMS-0954737 and CCF-1116928, as well as Dean's Research Fund. \section*{Proofs} \input{appendix} \end{document} \subsection{Partial Information} \label{sec:partial_info} We now turn to the partial (or, zero-th order) information model. Recall that the matrix $A$ is not known to the players, yet we are interested in finding $\epsilon$-optimal minimax strategies. On each round, the two players choose mixed strategies $f_t \in \Delta_n$ and $x_t \in \Delta_m$, respectively, and observe $f_t^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A x_t$. Now the question is, how many such observations do we need to get to an $\epsilon$-optimal minimax strategy? Can this be done while still ensuring the usual no-regret rate? The specific setting we consider below requires that on each round $t$, the two players play four times, and that these four plays are $\delta$-close to each other (that is, $\|f_t^i-f_t^j\|_1\leq \delta$ for $i,j\in\{1,\ldots,4\}$). Interestingly, up to logarithmic factors, the fast rate of the previous section is possible even in this scenario, but we do require the knowledge of the number of actions of the opposing player (or, an upper bound on this number). We leave it as an open problem the question of whether one can attain the $1/T$-type rate with only one play per round. \begin{minipage}[b]{0.51\linewidth} \frameit{ \small {\bf Player I} \\ $u_1,\ldots,u_{n-1}$ : orthonormal basis of $\Delta_n$\\ Initialize $g_1 , f_1 = \frac{1}{n}{\mathbf 1}_n$; Draw $i_0 \sim \mrm{Unif}([n-1])$ \\ At time $t=1$ to $T$\\ \hphantom{.}\hspace{3mm} {\bf Play $f_t$}\\ \hphantom{.}\hspace{3mm} Draw $i_t \sim \mrm{Unif}([n-1])$\\ \hphantom{.}\hspace{3mm} {\bf Observe : }\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $r^+_t = (f_t + \delta u_{i_{t-1}})^\top A x_t$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $r^-_{t} = (f_t - \delta u_{i_{t-1}})^\top A x_t$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\bar{r}^+_{t} = (f_t + \delta u_{i_{t}})^\top A x_t$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\bar{r}^-_{t} = (f_t - \delta u_{i_{t}})^\top A x_t$ \\ \hphantom{.}\hspace{3mm} {\bf Build estimates :}\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\hat{a}_t = \frac{n}{2 \delta}\left(r^+_{t} - r^-_{t}\right)u_{i_{t-1}}$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\bar{a}_t = \frac{n}{2 \delta}\left(\bar{r}^+_{t} - \bar{r}^-_{t}\right)u_{i_t}$\\ \hphantom{.}\hspace{3mm} {\bf Update : }\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $g_{t}(i) \propto g'_{t-1}(i) \exp\{-\eta_{t} \hat{a}_t(i) \}$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $g'_{t} = \left(1-\beta\right) g_{t} + (\beta/n){\mathbf 1} $\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $f_{t+1} (i) \propto g'_{t}(i) \exp\{-\eta_{t+1} \bar{a}_t(i) \}$\\ End } \end{minipage} \begin{minipage}[b]{0.51\linewidth} \frameit{ \small {\bf Player II} \\ $v_1,\ldots,v_{m-1}$ : orthonormal basis of $\Delta_m$\\ Initialize $y_1 , x_1 = \frac{1}{m}{\mathbf 1}_m$; Draw $j_0 \sim \mrm{Unif}([m-1])$ \\ At time $t=1$ to $T$\\ \hphantom{.}\hspace{3mm} {\bf Play $x_t$}\\ \hphantom{.}\hspace{3mm} Draw $j_t \sim \mrm{Unif}([m-1])$\\ \hphantom{.}\hspace{3mm} {\bf Observe : }\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $s^+_t = - f_t^\top A (x_t + \delta v_{j_{t-1}})$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $s^-_{t} = - f_t^\top A (x_t - \delta v_{j_{t-1}})$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\bar{s}^+_{t} = - f_t^\top A (x_t + \delta v_{j_{t}})$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\bar{s}^-_{t} = - f_t^\top A (x_t - \delta v_{j_{t}})$ \\ \hphantom{.}\hspace{3mm} {\bf Build estimates :}\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\hat{b}_t = \frac{m}{2 \delta}\left(s^+_{t} - s^-_{t}\right)v_{j_{t-1}}$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $\bar{b}_t = \frac{m}{2 \delta}\left(\bar{s}^+_{t} - \bar{s}^-_{t}\right)v_{j_t}$\\ \hphantom{.}\hspace{3mm} {\bf Update : }\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $y_{t}(i) \propto y'_{t-1}(i) \exp\{-\eta'_{t} \hat{b}_t(i) \}$\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $y'_{t} = \left(1-\beta\right) y_{t} + (\beta/m){\mathbf 1} $\\ \hphantom{.}\hspace{3mm} \hphantom{.}\hspace{3mm} $x_{t+1} (i) \propto y'_{t}(i) \exp\{-\eta'_{t+1} \bar{b}_t(i) \}$\\ End } \end{minipage} \begin{lemma}\label{lem:partial} Let $A\in [-1,1]^{n\times m}$, $\F=\Delta_n$, $\cX=\Delta_m$, let $\delta$ be small enough (e.g. exponentially small in $m,n,T$), and let $\beta=1/T^2$. If both players use above algorithms with the adaptive step sizes $$\textstyle \eta_t = \min\left\{\sqrt{\log(nT)} \frac{\sqrt{\sum_{i=1}^{t-1} \norm{\hat{a}_i - \bar{a}_{i-1}}_*^2} - \sqrt{\sum_{i=1}^{t-2} \norm{\hat{a}_i - \bar{a}_{i-1}}_*^2}}{\norm{\hat{a}_{t-1} - \bar{a}_{t-2}}_*^2} , \frac{1}{28 m\ \sqrt{\log(m T)}}\right\} $$ and $$\textstyle \eta'_t = \min\left\{\sqrt{\log(mT)} \frac{\sqrt{\sum_{i=1}^{t-1} \norm{\hat{b}_i - \bar{b}_{i-1}}_*^2} - \sqrt{\sum_{i=1}^{t-2} \norm{\hat{b}_i - \bar{b}_{i-1}}_*^2}}{\norm{\hat{b}_{t-1} - \bar{b}_{t-2}}_*^2} , \frac{1}{28 n\ \sqrt{\log(n T)}}\right\} $$ respectively, then the pair $(\bar{f}_T,\bar{x}_T)$ is an $$\mathcal{O}\left( \frac{\left( m \log(n T)\sqrt{\log(m T)} + n \log(m T) \sqrt{\log(n T)}\right) }{T}\right)$$ -approximate minimax equilibrium. Furthermore, if only one player (say, Player I) follows the above algorithm, her regret against any sequence $x_1,\ldots,x_T$ of plays is bounded by \begin{align*} \mathcal{O}\left(\frac{m \sqrt{\log(m T)} \log(n T) + n \sqrt{ \log(n T) \sum_{t=1}^T \norm{x_t - x_{t-1}}^2 }}{T}\right) \end{align*} \end{lemma} We leave it as an open problem to find an algorithm that attains the $1/T$-type rate when both players only observe the value $e_i^\ensuremath{{\scriptscriptstyle\mathsf{T}}} A e_j = A_{i,j}$ upon drawing \emph{pure} actions $i,j$ from their respective mixed strategies $f_t,x_t$. We hypothesize a rate better than $T^{-1/2}$ is not possible in this scenario. \section{Online Learning with Predictable Gradient Sequences} Let us describe the online convex optimization (OCO) problem and the basic algorithm studied in \cite{Chiangetal12,RakSri13pred}. Let $\F$ be a convex set of moves of the learner. On round $t=1,\ldots,T$, the learner makes a prediction $f_t\in\F$ and observes a convex function $G_t$ on $\F$. The objective is to keep \emph{regret} $$\frac{1}{T}\sum_{t=1}^T G_t(f_t)-G_t(f^*)$$ small for any $f^*\in\F$. Let $\cR$ be a $1$-strongly convex function w.r.t. some norm $\|\cdot\|$ on $\F$, and let $g_0 = \arg\min_{g\in\F} \cR(g)$. Suppose that at the beginning of every round $t$, the learner has access to $M_t$, a vector computable based on the past observations or side information. In this paper we study the \textbf{Optimistic Mirror Descent} algorithm, defined by the interleaved sequence \begin{align} f_{t} &= \argmin{f \in \F}~ \eta_t \inner{f,M_{t}} + \cD_\cR(f,g_{t-1}) \ , \ ~~~~ g_{t} = \argmin{g\in\F}~ \eta_t\inner{g,\nabla G_t(f_t)} + \cD_\cR(g,g_{t-1}) \end{align} where $\cD_\cR$ is the Bregman Divergence with respect to $\cR$ and $\{\eta_t\}$ is a sequence of step sizes that can be chosen adaptively based on the sequence observed so far. The method adheres to the OCO protocol since $M_t$ is available at the beginning of round $t$, and $\nabla G_t(f_t)$ becomes available after the prediction $f_t$ is made. The sequence $\{f_t\}$ will be called primary, while $\{g_t\}$ -- secondary. This method was proposed in \cite{Chiangetal12} for $M_t=\nabla G_{t-1} (f_{t-1})$, and the following lemma is a straightforward extension of the result in \cite{RakSri13pred} for general $M_t$: \begin{lemma} \label{lem:two-step-MD} Let $\F$ be a convex set in a Banach space $\cB$. Let $\cR:\cB\to\mathbb{R}$ be a $1$-strongly convex function on $\F$ with respect to some norm $\|\cdot\|$, and let $\|\cdot\|_*$ denote the dual norm. For any fixed step-size $\eta$, the Optimistic Mirror Descent Algorithm yields, for any $f^*\in\F$, \begin{align} \label{eq:main_bound} \sum_{t=1}^T G_t(f_t)-G_t(f^*) &\leq \sum_{t=1}^T \inner{f_t-f^*, \nabla_t} \le \eta^{-1} R^2 + \sum_{t=1}^T \norm{\nabla_t - M_t}_*\norm{g_{t} - f_t} - \frac{1}{2\eta} \sum_{t=1}^T \left(\norm{g_{t} - f_t}^2 + \norm{g_{t-1} - f_t}^2 \right) \end{align} where $R\geq 0$ is such that $\cD_\cR(f^*,g_0)\leq R^2$ and $\nabla_t = \nabla G_t(f_t)$. \end{lemma} When applying the lemma, we will often use the simple fact that \begin{align} \label{eq:alpha_split} \norm{\nabla_t - M_t}_*\norm{g_{t} - f_t} = \inf_{\rho>0}\left\{ \frac{\rho}{2} \norm{\nabla_t - M_t}_*^2 + \frac{1}{2\rho} \norm{g_{t} - f_t}^2 \right\} \ . \end{align} In particular, by setting $\rho=\eta$, we obtain the (unnormalized) regret bound of $\eta^{-1}R^2 + (\eta/2) \sum_{t=1}^T \norm{\nabla_t - M_t}_*^2$, which is $R\sqrt{2\sum_{t=1}^T \norm{\nabla_t - M_t}_*^2}$ by choosing $\eta$ optimally. Since this choice is not known ahead of time, one may either employ the doubling trick, or choose the step size adaptively: \begin{corollary} \label{cor:adaptive_step_size} Consider step size $$ \eta_t = R_{\max}\min\left\{\left(\sqrt{\sum_{i=1}^{t-1} \norm{\nabla_i - M_i}_*^2} + \sqrt{\sum_{i=1}^{t-2} \norm{\nabla_i - M_i}_*^2} \right)^{-1} , 1 \right\} \ $$ with $R_{\max}^2 = \sup_{f,g\in \F} \cD_{\cR}(f,g)$. Then regret of the Optimistic Mirror Descent algorithm is upper bounded by $$ 3.5R_{\max} \frac{\sqrt{\sum_{t=1}^T \norm{\nabla_t - M_t}_*^2} +1}{T} \ .$$ \end{corollary} These results indicate that tighter regret bounds are possible if one can guess the next gradient $\nabla_t$ by computing $M_t$. One such case arises in \emph{offline} optimization of a smooth function, whereby the previous gradient turns out to be a good proxy for the next one. More precisely, suppose we aim to optimize a function $G(f)$ whose gradients are Lipschitz continuous: $\|\nabla G(f)-\nabla G(g)\|_*\leq H\|f-g\|$ for some $H>0$. In this optimization setting, no guessing of $M_t$ is needed: we may simply query the oracle for the gradient and set $M_t=\nabla G(g_{t-1})$. The Optimistic Mirror Descent then becomes \begin{align*} f_{t} &= \argmin{f \in \F}~ \eta_t \inner{f,\nabla G(g_{t-1})} + \cD_\cR(f,g_{t-1}) \ , \ ~~~~ g_{t} = \argmin{g\in\F}~ \eta_t\inner{g,\nabla G(f_t)} + \cD_\cR(g,g_{t-1}) \end{align*} which can be recognized as the Mirror Prox method, due to Nemirovski \cite{nemirovski2004prox}. By smoothness, $\|\nabla G(f_t)-M_t\|_* = \|\nabla G(f_t)-\nabla G(g_{t-1})\|_*\leq H\|f_t-g_{t-1}\|$. Lemma~\ref{lem:two-step-MD} with Eq.~\eqref{eq:alpha_split} and $\rho=\eta = 1/H$ immediately yields a bound $$\sum_{t=1}^T G(f_t)-G(f^*) \le H R^2,$$ which implies that the average $\bar{f}_T = \frac{1}{T}\sum_{t=1}^T f_t$ satisfies $G(\bar{f}_T)-G(f^*)\leq HR^2/T$, a known bound for Mirror Prox. We now extend this result to arbitrary $\alpha$-H\"older smooth functions, that is convex functions $G$ such that $\|\nabla G(f)-\nabla G(g)\|_*\leq H\|f-g\|^\alpha$ for all $f,g\in \F$. \begin{lemma} \label{lem:Holder_smooth} Let $\F$ be a convex set in a Banach space $\cB$ and let $\cR:\cB\to\mathbb{R}$ be a $1$-strongly convex function on $\F$ with respect to some norm $\|\cdot\|$. Let $G$ be a convex $\alpha$-H\"older smooth function with constant $H>0$ and $\alpha\in [0,1]$. Then the average $\bar{f}_T = \frac{1}{T}\sum_{t=1}^T f_t$ of the trajectory given by Optimistic Mirror Descent Algorithm enjoys \begin{align*} G(\bar{f}_T)- \inf_{f \in \F}G(f) \le \frac{8 H R^{1 + \alpha} }{T^{\frac{1 + \alpha}{2}}} \end{align*} where $R\geq 0$ is such that $\sup_{f\in\F} \cD_\cR(f,g_0)\leq R$. \end{lemma} This result provides a smooth interpolation between the $T^{-1/2}$ rate at $\alpha=0$ (that is, no predictability of the gradient is possible) and the $T^{-1}$ rate when the smoothness structure allows for a dramatic speed up with a very simple modification of the original Mirror Descent. \section{Structured Optimization}\label{sec:structopt} In this section we consider the structured optimization problem $$ \argmin{f \in \F} G(f) $$ where $G(f)$ is of the form $G(f) = \sup_{x \in \X}\phi(f,x)$ with $\phi(\cdot,x)$ convex for every $x \in \X$ and $\phi(f,\cdot)$ concave for every $f \in \F$. Both $\F$ and $\X$ are assumed to be convex sets. While $G$ itself need not be smooth, it has been recognized that the structure can be exploited to improve rates of optimization if the function $\phi$ is smooth \cite{nesterov2005smooth}. From the point of view of online learning, we will see that the optimization problem of the saddle point type can be solved by playing two online convex optimization algorithms against each other (henceforth called Players I and II). Specifically, assume that Player I produces a sequence $f_1,\ldots,f_T$ by using a regret-minimization algorithm, such that \begin{align} \label{eq:reg1} \frac{1}{T} \sum_{t=1}^T \phi(f_t,x_t) - \inf_{f \in \F} \frac{1}{T} \sum_{t=1}^T \phi(f,x_t) \le \mrm{Rate}^1(x_1,\ldots,x_T) \end{align} and Player II produces $x_1,\ldots,x_T$ with \begin{align} \label{eq:reg2} \frac{1}{T} \sum_{t=1}^T \left(- \phi(f_t,x_t) \right) - \inf_{x \in \X} \frac{1}{T} \sum_{t=1}^T \left(-\phi(f_t,x)\right) \le \mrm{Rate}^2(f_1,\ldots,f_T) \ . \end{align} By a standard argument (see e.g. \cite{freund1999adaptive}), $$ \inf_{f} \frac{1}{T} \sum_{t=1}^T \phi(f,x_t) \le \inf_{f} \phi\left(f,\bar{x}_T\right) \le \sup_{x} \inf_{f} \phi\left(f,x\right) \leq \inf_{f} \sup_{x} \phi(f,x) \leq \sup_{x} \phi\left(\bar{f}_T,x\right) \leq \sup_{x} \frac{1}{T} \sum_{t=1}^T \phi(f_t,x) $$ where $\bar{f}_T = \frac{1}{T} \sum_{t=1}^T f_t$ and $\bar{x}_T = \frac{1}{T} \sum_{t=1}^T x_t$. By adding \eqref{eq:reg1} and \eqref{eq:reg2}, we have \begin{align} \label{eq:sandwich_minimax} \sup_{x \in \X} \frac{1}{T} \sum_{t=1}^T \phi(f_t,x) - \inf_{f \in \F} \frac{1}{T} \sum_{t=1}^T \phi(f,x_t) \leq \mrm{Rate}^1(x_1,\ldots,x_T) +\mrm{Rate}^2(f_1,\ldots,f_T) \end{align} which sandwiches the previous sequence of inequalities up to the sum of regret rates and implies near-optimality of $\bar{f}_T$ and $\bar{x}_T$. \begin{lemma} \label{lem:saddle_predictable} Suppose both players employ the Optimistic Mirror Descent algorithm with, respectively, predictable sequences $M^1_t$ and $M^2_t$, $1$-strongly convex functions $\cR_1$ on $\F$ (w.r.t. $\|\cdot\|_\F$) and $\cR_2$ on $\cX$ (w.r.t. $\|\cdot\|_\X$), and fixed learning rates $\eta$ and $\eta'$. Let $\{f_t\}$ and $\{x_t\}$ denote the primary sequences of the players while let $\{g_t\},\{y_t\}$ denote the secondary. Then for any $\alpha, \beta > 0$, \begin{align} \label{eq:minmaxopt} &\sup_{x \in \X} \phi\left(\bar{f}_T,x\right) - \inf_{f \in \F} \sup_{x \in \X} \phi(f,x) \\ & \leq \frac{ R_{1}^2}{\eta} + \frac{\alpha}{2} \sum_{t=1}^T \|\nabla_{f} \phi(f_t,x_t) - M^1_t\|_{\F^*}^2 + \frac{1}{2 \alpha} \sum_{t=1}^T \norm{g_{t} - f_t}_{\F}^2 -\frac{1}{2 \eta} \sum_{t=1}^T \left(\norm{g_{t} - f_t}_{\F}^2 + \norm{g_{t-1} - f_t}_{\F}^2 \right) \notag\\ & ~+ \frac{ R_{2}^2}{\eta'} + \frac{\beta}{2} \sum_{t=1}^T \|\nabla_{x} \phi(f_t,x_t) - M^2_t\|_{\X^*}^2 + \frac{1}{2 \beta} \sum_{t=1}^T \norm{y_{t} - x_t}_{\X}^2 -\frac{1}{2 \eta'} \sum_{t=1}^T \left(\norm{y_{t} - x_t}_{\X}^2 + \norm{y_{t-1} - x_t}_{\X}^2 \right) \notag \end{align} where $R_1$ and $R_2$ are such that $\cD_{\cR_1}(f^*,g_0)\leq R_{1}^2$ and $\cD_{\cR_2}(x^*,y_0)\leq R_{2}^2$, and $\bar{f}_T = \frac{1}{T} \sum_{t=1}^T f_t$. \end{lemma} The proof of Lemma~\ref{lem:saddle_predictable} is immediate from Lemma~\ref{lem:two-step-MD}. We obtain the following corollary: \begin{corollary} \label{cor:holder_saddle} Suppose $\phi:\F\times\X\mapsto \mathbb{R}$ is H\"older smooth in the following sense: \begin{center} ~~~~~~~~$\|\nabla_f \phi(f,x) - \nabla_f \phi(g,x)\|_{\F^*} \leq H_1\|f-g\|^{\alpha}_\F,~~~ \|\nabla_f \phi(f,x) - \nabla_f \phi(f,y)\|_{\F^*} \leq H_2\|x-y\|^{\alpha'}_\X$\\ and ~~$\|\nabla_x \phi(f,x) - \nabla_x \phi(g,x)\|_{\X^*} \leq H_4\|f-g\|^{\beta}_\F,~~~ \|\nabla_x \phi(f,x) - \nabla_x \phi(f,y)\|_{\X^*} \leq H_3\|x-y\|^{\beta'}_\X$. \end{center} Let $\gamma = \min\{\alpha, \alpha', \beta, \beta'\}$, $H = \max\{H_1,H_2,H_3,H_4\}$. Suppose both players employ Optimistic Mirror Descent with $M^1_t = \nabla_{f} \phi(g_{t-1},y_{t-1})$ and $M^2_t = \nabla_{x} \phi(g_{t-1},y_{t-1})$, where $\{g_t\}$ and $\{y_t\}$ are the secondary sequences updated by the two algorithms, and with step sizes $\eta = \eta' = (R_{1}^2 + R_{2}^2)^{\frac{1 - \gamma}{2}} (2 H)^{-1} \left(\frac{T}{2}\right)^{\frac{\gamma - 1}{2}}$. Then \begin{align} \sup_{x \in \X}~ & \phi\left(\bar{f}_T,x\right) - \inf_{f \in \F} \sup_{x \in \X} \phi(f,x) \le \frac{4 H (R_{1}^2 + R_{2}^2)^{\frac{1 + \gamma}{2}}}{ T^{\frac{1 + \gamma}{2}}} \end{align} \end{corollary} As revealed in the proof of this corollary, the negative terms in \eqref{eq:minmaxopt}, that come from an upper bound on regret of Player I, in fact contribute to cancellations with positive terms in regret of Player II, and vice versa. Such a coupling of the upper bounds on regret of the two players can be seen as leading to faster rates under the appropriate assumptions, and this idea will be exploited to a great extent in the proofs of the next section.
\section{Introduction} The aim of this paper is to give a self-contained, accessible and `elementary' proof of of the following theorem, which we call the Landau-Ingham Tauberian theorem: \begin{theorem} \label{theor-LI} Let $f:[1,\infty)\rightarrow\7R$ be non-negative and non-decreasing and assume that \begin{equation} F(x):=\sum_{n\le x}f\left(\frac{x}{n}\right) \quad \mathrm{satisfies} \ \quad\ F(x)=Ax\log x+Bx+C\frac{x}{\log x}+o\left(\frac{x}{\log x}\right). \label{eq-Kar}\end{equation} Then $f(x)=Ax+o(x)$, equivalently $f(x)\sim Ax$. \end{theorem} The interest of this theorem derives from the fact that, while ostensibly it is a result firmly located in classical real analysis, the prime number theorem (PNT) $\pi(x)\sim\frac{x}{\log x}$ can be deduced from it by a few lines of Chebychev-style reasoning. (Cf.\ the Appendix.) Versions of Theorem \ref{theor-LI} were proven by Landau \cite[\S 160]{landau} as early as 1909, Ingham \cite[Theorem 1]{ingham}, Gordon \cite{gordon} and Ellison \cite[Theorem 3.1]{ellison}, but none of these proofs was from scratch. Landau used as input the identity $\sum_n\frac{\mu(n)\log n}{n}=-1$. But the latter easily implies $M(x)=\sum_{n\le x}\mu(n)=o(x)$ which (as also shown by Landau) is equivalent to the PNT. Actually, $\sum_n\frac{\mu(n)\log n}{n}=-1$ is `stronger' than the PNT in the sense that it cannot be deduced from the latter (other than by elementarily reproving the PNT with a sufficiently strong remainder estimate). In this sense, Gordon's version of Theorem \ref{theor-LI} is an improvement, in that he uses as input exactly the PNT (in the form $\psi(x)\sim x$) and thereby shows that Theorem \ref{theor-LI} is not `stronger' than the PNT. Ellison's version assumes $M(x)=o(x)$ (and an $O(x^\beta)$ remainder with $\beta<1$ in (\ref{eq-Kar})). It is thus clear that none of these approaches provides a proof of the PNT. Ingham's proof, on the other hand, departs from the information that $\zeta(1+it)\ne 0$ (which can be deduced from the PNT, but also be proven {\it ab initio}). Thus his proof is not `elementary', but arguably it is one of the nicer and more conceptual deductions of the PNT from $\zeta(1+it)\ne 0$ -- though certainly not the simplest (which is \cite{zagier}) given that the proof requires Wiener's $L^1$-Tauberian theorem. Our proof of Theorem \ref{theor-LI} will essentially follow the elementary Selberg-style proof given by Karamata \footnote{Note des \'editeurs : Jovan Karamata, n\'e pr\`es de Belgrade en 1902 et mort \`a Gen\`eve en 1967, fut professeur \`a Gen\`eve d\`es 1951 et directeur de L'Enseignement Math\'ematique de 1954 \`a 1967. Voir M.\ Tomi\'c, \emph{Jovan Karamata (1902-1967)}, Enseignement Math.\ \ (2) {\bf 15}, 1--20 (1969).} \cite{karamata} under the assumption that $f$ is the summatory function of an arithmetic function, i.e.\ constant between successive integers. We will remove this assumption. For the proof of the PNT, this generality is not needed, but from an analysis perspective it seems desirable, and it brings us fairly close to Ingham's version of the theorem, which differed only in having $o(x)$ instead of $C\frac{x}{\log x}+o(\frac{x}{\log x})$ in the hypothesis. Unfortunately, Karamata's paper \cite{karamata} seems to be essentially forgotten: There are so few references to it that we can discuss them all. It is mentioned in \cite{EI} by Erd\"os and Ingham and in the book \cite{ellison} of Ellison and Mend\`es-France. (Considering that the latter authors know Karamata's work, one may find it surprising that for their elementary proof of the PNT they chose the somewhat roundabout route of giving a Selberg-style proof of $M(x)=o(x)$, using this to prove a weak version of Theorem \ref{theor-LI}, from which then $\psi(x)\sim x$ is deduced.) Even the two books \cite{BGT,korevaar} on Tauberian theory only briefly mention Karamata's \cite{karamata} (or just the survey paper \cite{karamata2}) but then discuss in detail only Ingham's proof. Finally, \cite{karamata,karamata2} are cited in the recent historical article \cite{nik}, but its emphasis is on other matters. We close by noting that Karamata is not even mentioned in the only other paper pursuing an elementary proof of a Landau-Ingham theorem, namely Balog's \cite{balog}, where a version of Theorem \ref{theor-LI} with a (fairly weak) error term in the conclusion is proven. Our reason for advertising Karamata's approach is that, in our view, it is the conceptually cleanest and simplest of the Selberg-Erd\"os style proofs of the PNT, cf.\ \cite{selberg,erdos} and followers, e.g.\ \cite{postn, nev, kalecki, levinson, schwarz, pollack}. For $f=\psi$ and $f(x)=M(x)+\lfloor x\rfloor$, Theorem \ref{theor-LI} readily implies $\psi(x)=x+o(x)$ and $M(x)=o(x)$. Making these substitutions in advance, the proof simplifies only marginally, but it becomes less transparent (in particular for $f=\psi$) due to an abundance of non-linear expressions. By contrast, Theorem \ref{theor-LI} is linear w.r.t.\ $f$ and $F$. To be sure, also the proof given below has a non-linear core, cf.\ (\ref{eq-selb}) and Proposition \ref{prop-selb}, but by putting the latter into evidence, the logic of the proof becomes clearer. One is actually led to believe that the non-linear component of the proof is inevitable, as is also suggested by Theorem 2 in Erd\"os' \cite{erdos2}, to wit \[ a_k\ge 0\ \forall k\ge 1\ \wedge\ \sum_{k=1}^N ka_k+\sum_{k+l\le N}a_ka_l=N^2+O(1)\ \ \Rightarrow\ \ \sum_{k=1}^N a_k=N+O(1),\] from which the PNT can be deduced with little effort. (Cf.\ \cite{HT} for more in this direction.) Another respect in which \cite{karamata} is superior to most of the later papers, including V.\ Nevanlinna's \cite{nev} (whose approach is adopted by several books \cite{schwarz,pollack}), concerns the Tauberian deduction of the final result from a Selberg-style integral inequality. In \cite{karamata}, this is achieved by a theorem attributed to Erd\"os (Theorem \ref{theor-2} below) with clearly identified, obviously minimal hypotheses and an elegant proof. This advantage over other approaches like \cite{nev}, which tend to use further information about the discontinuities of the function under consideration, is essential for our generalization to arbitrary non-decreasing functions. However, we will have to adapt the proof (not least in order to work around an obscure issue). In our exposition we make a point of avoiding the explicit summations over (pairs of) primes littering many elementary proofs, almost obtaining a proof of the PNT free of primes! This is achieved by defining the M\"obius and von Mangoldt functions $\mu$ and $\Lambda$ in terms of the functional identities they satisfy and using their explicit computation only to show that they are bounded and non-negative, respectively. Some of the proofs are formulated in terms of parametric Stieltjes integrals, typically of the form $\int f(x/t)dg(t)$ and integration by parts. We also do this in situations where $f$ and $g$ may both be discontinuous. Since our functions will always have bounded variation, thus at most countably many discontinuities, this can be justified by observing that the resulting identities hold for all $x$ outside a countable set. Alternatively, we can replace $f(x)$ at every point of discontinuity by $(f(x+0)+f(x-0))/2$ without changing the asymptotics. For such functions, integration by parts always holds in the theory of Lebesgue-Stieltjes integration, cf.\ \cite{hewitt,HS}. The proof of Theorem \ref{theor-LI} exhibited below is, including all preliminaries, just 12 pages long, and the author hopes that this helps dispelling the prejudice that the elementary proofs of the PNT are (conceptionally and/or technically) difficult. Indeed he thinks that this is the most satisfactory of the elementary (and in fact of all) proofs of the PNT in that, besides not invoking complex analysis or Riemann's $\zeta$-function, it minimizes number theoretic reasoning to a very well circumscribed minimum. One may certainly dispute that this is desirable, but we will argue elsewhere that it is. The author is of course aware of the fact that the more direct elementary proofs of the PNT give better control of the remainder term. (Cf.\ the review \cite{diamond} and the very recent paper \cite{koukou}, which provides a ``a new and largely elementary proof of the best result known on the counting function of primes in arithmetic progressions''.) It is not clear whether this is necessarily so. \\ \noindent{\it Acknowledgments.} The author would like to thank the referees for constructive comments that led to several improvements, in particular a better proof of Corollary \ref{coro-E2}. \section{First steps and strategy} \begin{prop} \label{prop-estim} Let $f:[1,\infty)\rightarrow\7R$ be non-negative and non-decreasing and assume that $F(x)=\sum_{n\le x}f(x/n)$ satisfies $F(x)=Ax\log x+Bx+o(x)$. Then \begin{itemize} \item[(i)] $f(x)=O(x)$. \item[(ii)] $\displaystyle \int_{1-0}^x \frac{df(t)}{t}=A\log x+O(1)$. \item[(iii)] $\displaystyle\int_1^x\frac{f(t)-At}{t^2}dt=O(1)$. \end{itemize} \end{prop} {\noindent\it Proof. } (i) Following Ingham \cite{ingham}, we define $f$ to be $0$ on $[0,1)$ and compute \begin{eqnarray*} f(x)-f\left(\frac{x}{2}\right)+f\left(\frac{x}{3}\right)-\cdots &=& F(x)-2F\left(\frac{x}{2}\right)\\ &=& Ax\log x+Bx-2\left(A\frac{x}{2}\log\frac{x}{2}+B\frac{x}{2}\right)+o(x)=Ax\log 2+o(x). \end{eqnarray*} With positivity and monotonicity of $f$, this gives $f(x)-f(x/2)\le Kx$ for some $K>0$. Adding these inequalities for $x, \frac{x}{2}, \frac{x}{4},\ldots$, we find $f(x)\le 2Kx$. Together with $f\ge 0$, this gives (i). (ii) We compute \begin{eqnarray*} F(x) &=& \sum_{n\le x} f\left(\frac{x}{n}\right) =\int_{1-0}^x f\left(\frac{x}{t}\right) d\lfloor t\rfloor \\ &=& \left[\lfloor t\rfloor f\left(\frac{x}{t}\right)\right]_{t=1-0}^{t=x} -\int_{1-0}^x\lfloor t\rfloor\,df\left(\frac{x}{t}\right)\\ &=& \lfloor x\rfloor f(1) -\int_{1-0}^x t\,df\left(\frac{x}{t}\right) +\int_{1-0}^x (t-\lfloor t\rfloor)\,df\left(\frac{x}{t}\right) \\ &=& O(x) + \int_{1-0}^x \frac{x}{u}df(u) +\int_{1-0}^x (t-\lfloor t\rfloor)\,df\left(\frac{x}{t}\right). \\ \end{eqnarray*} In view of $0\le t-\lfloor t\rfloor<1$ and the weak monotonicity of $f$, the last integral is bounded by $|\int_1^x df(x/t)|=f(x)-f(1)$, which is $O(x)$ by (i). Using the hypothesis about $F$, we have \[ Ax\log x+Bx+o(x)=O(x)+x\int_{1-0}^x\frac{df(t)}{t}+O(x),\] and division by $x$ proves the claim. (iii) Integrating by parts, we have \begin{eqnarray*} \int_1^x\frac{f(t)-At}{t^2}dt &=& -\frac{f(x)}{x}+\int_{1-0}^x \frac{df(t)}{t}-\int_1^x\frac{A}{t}dt \\ &=& O(1)+(A\log x+O(1))-A\log x=O(1),\end{eqnarray*} where we used (i) and (ii). \qed \begin{remark} 1. The proposition can be proven under the weaker assumption $F(x)=Ax\log x+O(x)$, but we don't bother since later we will need the stronger hypothesis anyway. 2. Theorem \ref{theor-LI}, which we ultimately want to prove, implies a strong form of (iii): $\int_1^\infty\frac{f(t)-At}{t^2}dt=B-\gamma A$, cf.\ \cite{ingham}. Conversely, existence of the improper integral already implies $f(x)\sim Ax$, cf.\ \cite{zagier}. 3. Putting $f=\psi$ and using (\ref{eq-ch}), the above proofs of (i) and (ii) reduce to those of Chebychev and Mertens, respectively. \hfill{$\Box$}\medskip\end{remark} The following two theorems will be proven in Sections \ref{s-th1} and \ref{s-th2}, respectively: \begin{theorem} \label{theor-1} Let $f,F$ be as in Theorem \ref{theor-LI}. Then $g(x)=f(x)-Ax$ satisfies \begin{equation} \frac{|g(x)|}{x} \le \frac{1}{\log x}\int_1^x \frac{|g(t)|}{t^2}\,dt +o(1)\ \ \ \mathrm{as}\ \ x\rightarrow\infty. \label{eq-ineq}\end{equation} (Here $f(x)\le g(x)+o(1)$ means that $f(x)\le g(x)+h(x)\ \forall x$, where $h(x)\rightarrow 0$ as $x\rightarrow\infty$.) \end{theorem} \begin{theorem} \label{theor-2} For $g:[1,\infty)\rightarrow\7R$, assume that there are $M, M'\ge 0$ such that \begin{equation} x\mapsto g(x)+Mx\ \ \mbox{is non-decreasing}, \label{et1}\end{equation} \begin{equation} \left| \int_1^x \frac{g(t)}{t^2} dt\right| \le M' \quad\forall x\ge 1. \label{et2}\end{equation} Then \begin{equation} S:=\limsup_{x\rightarrow\infty}\frac{|g(x)|}{x}<\infty, \label{et3}\end{equation} and when $S>0$ we have \begin{equation} \limsup_{x\rightarrow\infty} \frac{1}{\log x}\int_1^x\frac{|g(t)|}{t^2}dt<S. \label{et4}\end{equation} \end{theorem} \begin{remark} \label{rem-th2} 1. Note that (\ref{et1}) implies that $g$ is Riemann integrable over finite intervals. 2. In our application, (\ref{et3}) already follows from Proposition \ref{prop-estim} so that we do not need the corresponding part of the proof of Theorem \ref{theor-2}. It will be proven nevertheless in order to give Theorem \ref{theor-2} an independent existence. \hfill{$\Box$}\medskip\end{remark} \noindent{\it Proof of Theorem \ref{theor-LI} assuming Theorems \ref{theor-1} and \ref{theor-2}.} Since $f$ is nondecreasing, it is clear that $g(x)=f(x)-Ax$ satisfies (\ref{et1}) with $M=A$, and (\ref{et2}) is implied by Proposition \ref{prop-estim}(iii). Now $S=\limsup|g(x)|/x$ is finite, by either Proposition \ref{prop-estim}(i) or the first conclusion of Theorem \ref{theor-2}. Furthermore, $S>0$ would imply (\ref{et4}). But combining this with the result (\ref{eq-ineq}) of Theorem \ref{theor-1}, we would have the absurdity \[ S=\limsup_{x\rightarrow\infty}\frac{|g(x)|}{x}\le \limsup_{x\rightarrow\infty}\frac{1}{\log x}\int_1^x \frac{|g(t)|}{t^2}\,dt<S.\] Thus $S=0$ holds, which is equivalent to $\frac{g(x)}{x}=\frac{f(x)-Ax}{x}\rightarrow 0$, as was to be proven. \qed The next two sections are dedicated to the proofs of Theorems \ref{theor-1} and \ref{theor-2}. The statements of both results are free of number theory, and this is also the case for the proof of the second. The proof of Theorem \ref{theor-1}, however, uses a very modest amount of number theory, but nothing beyond M\"obius inversion and the divisibility theory of $\7N$ up to the fundamental theorem of arithmetic. \section{Proof of Theorem \ref{theor-1}}\label{s-th1} \subsection{Arithmetic}\label{ss-mobius} The aim of this subsection is to collect the basic arithmetic results that will be needed. We note that this is very little. We begin by noting that $(\7N,\cdot,1)$ is an abelian monoid. Given $n,m\in\7N$, we call $m$ a divisor of $n$ if there is an $r\in\7N$ such that $mr=n$, in which case we write $m|n$. In view of the additive structure of the semiring $\7N$, it is clear that the monoid $\7N$ has cancellation ($ab=ac\Rightarrow b=c$), so the quotient $r$ above is unique, and that the set of divisors of any $n$ is finite. Calling a function $f:\7N\rightarrow\7R$ an arithmetic function, the facts just stated allow us to define: \begin{defin} If $f,g:\7N\rightarrow\7R$ are arithmetic functions, their Dirichlet convolution $f\star g$ denotes the function \[ (f\star g)(n)=\sum_{d|n}f(d)g\left(\frac{n}{d}\right)=\sum_{a,b\atop ab=n} f(a)g(b).\] \end{defin} It is easy to see that Dirichlet convolution is commutative and associative. It has a unit given by the function $\delta$ defined by $\delta(1)=1$ and $\delta(n)=0$ if $n\ne 1$. By $\11$ we denote the constant function $\11(n)=1$. Clearly, $(f\star\11)(n)=\sum_{d|n}f(d)$. \begin{lemma} There is a unique arithmetic function $\mu$, called the M\"obius function, such that $\mu\star\11=\delta$. \end{lemma} {\noindent\it Proof. } $\mu$ must satisfy $\sum_{d|n}\mu(d)=\delta(n)$. Taking $n=1$ we see that $\mu(1)=1$. For $n>1$ we have $\sum_{d|n}\mu(d)=0$, which is equivalent to \[ \mu(n)=-\sum_{d|n \atop d<n}\mu(d). \] This uniquely determines $\mu(n)\in\7Z$ inductively in terms of $\mu(m)$ with $m<n$. \qed \begin{prop} \label{prop-mu1} \begin{itemize} \item[(i)] $\mu$ is multiplicative, i.e.\ $\mu(nm)=\mu(n)\mu(m)$ whenever $(n,m)=1$. \item[(ii)] If $p$ is a prime then $\mu(p)=-1$, and $\mu(p^k)=0$ if $k\ge 2$. \item[(iii)] $\mu(n)=O(1)$, i.e.\ $\mu$ is bounded. \end{itemize} \end{prop} {\noindent\it Proof. } (i) Since $\mu(1)=1$, $\mu(nm)=\mu(n)\mu(m)$ clearly holds if $n=1$ or $m=1$. Assume, by way of induction, that $\mu(uv)=\mu(u)\mu(v)$ holds whenever $(u,v)=1$ and $uv<nm$, and let $n\ne 1\ne m$ be relatively prime. Then every divisor of $nm$ is of the form $st$ with $s|n, t|m$, so that \begin{eqnarray*} 0 &=& \sum_{d|nm} \mu(d)=\mu(nm)+\sum_{s|n, t|m\atop st<nm}\mu(st) =\mu(nm)+\sum_{s|n, t|m\atop st<nm}\mu(s)\mu(t) \\ &=& \mu(nm)+\sum_{s|n}\mu(s)\sum_{t|m}\mu(t)-\mu(n)\mu(m)=\mu(nm)-\mu(n)\mu(m), \end{eqnarray*} which is the inductive step. (ii) For $k\ge 1$, we have $\mu(p^k)=-\sum_{i=0}^{k-1}\mu(p^i)$, inductively implying $\mu(p)=-1$ and $\mu(p^k)=0$ if $k\ge 2$. Thus $\mu(p^k)\in\{0,-1\}$, which together with multiplicativity (i) gives $\mu(n)\in\{-1,0,1\}$ for all $n$, thus (iii). \qed \begin{prop} \label{prop-Lambda1} \begin{itemize} \item[(i)] The arithmetic function $\Lambda:=\log\star\mu$ is the unique solution of $\Lambda\star\11=\log$. \item[(ii)] $\Lambda(n)=-\sum_{d|n}\mu(d)\log d$. In particular, $\Lambda(1)=0$. \item[(iii)] $\Lambda(n)=\log p$ if $n=p^k$ where $p$ is prime and $k\ge 1$, and $\Lambda(n)=0$ otherwise. \item[(iv)] $\Lambda(n)\ge 0$. \end{itemize} \end{prop} {\noindent\it Proof. } (i) Existence: $\log\star\mu\star\11=\log\star\delta=\log$. Uniqueness: If $\Lambda_1\star\11=\log=\Lambda_2\star\11$ then $\Lambda_1=\Lambda_1\star\delta=\Lambda_1\star\11\star\mu=\Lambda_2\star\11\star\mu=\Lambda_2\star\delta=\Lambda_2$. (ii) $\Lambda(n) =\sum_{d|n}\mu(d)\log\frac{n}{d}=\sum_{d|n}\mu(d)(\log n-\log d) =\log n\sum_{d|n}\mu(d)-\sum_{d|n}\mu(d)\log d$. Now use $\sum_{d|n}\mu(d)=\delta(n)$. $\Lambda(1)=0$ is obvious. (iii) Using (ii), we have $\Lambda(p^k)=-\sum_{l=0}^k \mu(p^l) l\log p$, which together with Proposition \ref{prop-mu1}(ii) implies $\Lambda(p^k)=\log p\ \forall k\ge 1$. If $n,m>1$ and $(n,m)=1$ then by the multiplicativity of $\mu$, \begin{eqnarray*} \Lambda(nm) &=&-\sum_{s|n}\sum_{t|m}\mu(st)\log(st) =-\sum_{s|n}\sum_{t|m}\mu(s)\mu(t)(\log s+\log t)\\ &=&\sum_{s|n}\mu(s)\log s\sum_{t|m}\mu(t)+\sum_{t|m}\mu(t)\log t\sum_{s|n}\mu(s)=0. \end{eqnarray*} (iv) Obvious consequence of (iii). \qed \begin{remark} The only properties of $\mu$ and $\Lambda$ needed for the proof of Theorem \ref{theor-1} are the defining ones ($\mu\star\11=\delta,\ \Lambda\star\11=\log$), the trivial consequence (ii) in Proposition \ref{prop-Lambda1}, and the boundedness of $\mu$ and the non-negativity of $\Lambda$. In particular, the explicit computations of $\mu(n)$ and $\Lambda(n)$ in terms of the prime factorization of $n$ were only needed to prove the latter two properties. (Of course, the said properties of the functions $\mu$ and $\Lambda$ would be obvious if one defined them by the explicit formulae proven above, but this would be ad hoc and ugly, and one would still need to use the fundamental theorem of arithmetic for proving that $\mu\star\11=\delta$ and $\Lambda\star\11=\log$.) Note that prime numbers will play no r\^ole whatsoever before we turn to the actual proof of the prime number theorem in the Appendix, where the computation of $\Lambda(n)$ will be used again. \hfill{$\Box$}\medskip\end{remark} \subsection{The (weighted) M\"obius transform} \begin{defin} Given a function $f:[1,\infty)\rightarrow\7R$, its `M\"obius transform' is defined by \[ F(x)=\sum_{n\le x}f\left(\frac{x}{n}\right).\] \end{defin} \begin{lemma} \label{lem-mobius}The M\"obius transform $f\mapsto F$ is invertible, the inverse M\"obius transform being given by \[ f(x)=\sum_{n\le x} \mu(n)F\left(\frac{x}{n}\right).\] \end{lemma} {\noindent\it Proof. } We compute \begin{eqnarray*} \sum_{n\le x}\mu(n) F\left(\frac{x}{n}\right) &=& \sum_{n\le x}\mu(n) \sum_{m\le x/n}f\left(\frac{x}{nm}\right) =\sum_{nm\le x}\mu(n)f\left(\frac{x}{nm}\right) \\ &=& \sum_{r\le x} f\left(\frac{x}{r}\right) \sum_{s|r}\mu(s) =\sum_{r\le x} f\left(\frac{x}{r}\right) \delta(r)=f(x), \end{eqnarray*} where we used the defining property $\sum_{d|n}\mu(d)=\delta(n)$ of $\mu$. \qed \begin{remark} Since the point of Theorem \ref{theor-LI} is to deduce information about $f$ from information concerning its M\"obius transform $F$, it is tempting to appeal to Lemma \ref{lem-mobius} directly. However, in order for this to succeed, we would need control over $M(x)=\sum_{n\le x}\mu(n)$, at least as good as $M(x)=o(x)$. But then one is back in Ellison's approach mentioned in the introduction. The essential idea of the Selberg-Erd\"os approach to the PNT, not entirely transparent in the early papers but clarified soon after \cite{TI}, is to consider weighted M\"obius inversion formulae as follows. \hfill{$\Box$}\medskip\end{remark} \begin{lemma} \label{lem-TI1} Let $f:[1,\infty)\rightarrow\7R$ be arbitrary and $F(x)=\sum_{n\le x}f(x/n)$. Then \begin{equation} f(x)\log x+\sum_{n\le x}\Lambda(n)f\left(\frac{x}{n}\right) =\sum_{n\le x}\mu(n)\log\frac{x}{n}\,F\left(\frac{x}{n}\right). \label{eq-TI}\end{equation} \end{lemma} {\noindent\it Proof. } We compute \[ \sum_{n\le x}\mu(n)\log\frac{x}{n}F\left(\frac{x}{n}\right) =\log x \sum_{n\le x}\mu(n)F\left(\frac{x}{n}\right)-\sum_{n\le x}\mu(n)\log n\,F\left(\frac{x}{n}\right).\] By Lemma \ref{lem-mobius}, the first term equals $f(x)\log x$, whereas for the second we have \begin{eqnarray*} \sum_{n\le x}\mu(n)\log n\,F\left(\frac{x}{n}\right) &=& \sum_{n\le x}\mu(n)\log n\,\sum_{m\le x/n} f\left(\frac{x}{nm}\right) =\sum_{nm\le x}\mu(n)\log n\, f\left(\frac{x}{nm}\right) \\ &=& \sum_{s\le x}\left( \sum_{n|s}\mu(n)\log n\right) \,f\left(\frac{x}{s}\right) =-\sum_{s\le x} \Lambda(s) \,f\left(\frac{x}{s}\right), \end{eqnarray*} the last equality being Proposition \ref{prop-Lambda1}(ii). Putting everything together, we obtain (\ref{eq-TI}). \qed \begin{remark} 1. Eq.\ (\ref{eq-TI}) is known as the `Tatuzawa-Iseki formula', cf.\ \cite[(8)]{TI} (and \cite[p.\ 24]{karamata}). 2. Without the factor $\log(x/n)$ on the right hand side, (\ref{eq-TI}) reduces to M\"obius inversion. Thus (\ref{eq-TI}) is a sort of weighted M\"obius inversion formula. The presence of the sum involving $f(x/n)$ is very much wanted, since it will allow us to obtain the integral inequality (\ref{eq-ineq}) involving all $f(t), t\in[1,x]$. In order to do so, we must get rid of the explicit appearance of the function $\Lambda(n)$, which is very irregular and about which we know little. This requires some preparations. \hfill{$\Box$}\medskip\end{remark} \begin{lemma} \label{lem-TI2} For any arithmetic function $f:\7N\rightarrow\7R$ we have \[ f(n)\log n+\sum_{d|n}\Lambda(d)f\left(\frac{n}{d}\right) =\sum_{d|n}\mu(d)\,\log\frac{n}{d}\,\sum_{m|(n/d)}f(m). \] In particular, we have Selberg's identity: \begin{equation} \Lambda(n)\log n+\sum_{d|n}\Lambda(d)\Lambda\left(\frac{n}{d}\right) =\sum_{d|n}\mu(d)\log^2\frac{n}{d}. \label{eq-selb}\end{equation} \end{lemma} {\noindent\it Proof. } If $f$ is an arithmetic function, i.e.\ defined only on $\7N$, we extend it to $\7R$ as being $0$ on $\7R\backslash\7N$. With this extension, \[ F(n)=\sum_{m\le n}f\left(\frac{n}{m}\right) =\sum_{m|n}f\left(\frac{n}{m}\right) =\sum_{m|n}f(m),\] so that (\ref{eq-TI}) becomes the claimed identity. Taking $f(n)=\Lambda(n)$ and using $\sum_{d|n}\Lambda(d)=\log n$, Selberg's formula follows. \qed \subsection{Preliminary estimates} \begin{lemma} The following elementary estimates hold as $x\rightarrow\infty$: \begin{eqnarray} \sum_{n\le x} \frac{1}{n} &=& \log x+\gamma+O\left(\frac{1}{x}\right), \label{s1}\\ \sum_{n\le x} \frac{\log n}{n} &=& \frac{\log^2x}{2}+c+O\left(\frac{1+\log x}{x}\right), \label{s5}\\ \sum_{n\le x} \log n &=& x\log x-x+O(\log x), \label{s2}\\ \sum_{n\le x} \log\frac{x}{n} &=& x+O(\log x), \label{s2b}\\ \sum_{n\le x} \log^2 n &=& x(\log^2x-2\log x+1)+ O(\log^2x),\label{s3}\\ \sum_{n\le x} \log^2\frac{x}{n} &=& x+O(\log^2x). \label{s3b} \end{eqnarray} Here, $\gamma$ is Euler's constant and $c>0$. \end{lemma} {\noindent\it Proof. } (\ref{s1}): We have \[ \sum_{n=1}^N\frac{1}{n}-\int_1^N\frac{dt}{t}=\int_{1-0}^N \frac{d(\lfloor t\rfloor -t)}{t} =\left[\frac{\lfloor t\rfloor -t}{t}\right]_1^N+\int_1^N \frac{t-\lfloor t\rfloor}{t^2}dt. \] Since $0\le t-\lfloor t\rfloor<1$, the integral on the r.h.s.\ converges as $N\rightarrow\infty$ to some number $\gamma$ (Euler's constant) strictly between 0 and $1=\int_1^\infty dt/t^2$. Thus \[ \sum_{n=1}^N\frac{1}{n}=\int_1^N\frac{dt}{t}+\gamma-\int_N^\infty \frac{t-\lfloor t\rfloor}{t^2}dt =\log N+\gamma+O\left(\frac{1}{N}\right). \] (\ref{s5}): Similarly to the proof of (\ref{s1}), we have \[ \sum_{n=1}^N\frac{\log n}{n}-\int_1^N\frac{\log t}{t}dt=\int_{1-0}^N \frac{\log t}{t}\,d(\lfloor t\rfloor -t) =\left[\frac{(\lfloor t\rfloor -t)\log t}{t}\right]_1^N+\int_1^N \frac{(t-\lfloor t\rfloor)\log t}{t^2}dt. \] The final integral converges to some $c>0$ as $N\rightarrow\infty$ since $(\log t)/t^2=O(t^{-2+\varepsilon})$. Using \[ \int_1^x \frac{\log t}{t}dt= \frac{\log^2 x}{2},\quad\quad \int_N^\infty\frac{\log t}{t^2}dt=-\left[\frac{\log t}{t}\right]_N^\infty+\int_N^\infty\frac{dt}{t^2} =\frac{1+\log N}{N} \] we have \[ \sum_{n=1}^N\frac{\log n}{n}=\int_1^x \frac{\log t}{t}dt+c-\int_N^\infty \frac{(t-\lfloor t\rfloor)\log t}{t^2}dt =\frac{\log^2 x}{2} + c + O\left(\frac{1+\log x}{x}\right). \] (\ref{s2}): By monotonicity, we have \[ \int_1^x \log t\,dt \le \sum_{n\le x} \log n \le \int_1^{x+1} \log t\,dt. \] Combining this with $\int_1^x\log t\,dt=x\log x-x+1$, (\ref{s2}) follows. (\ref{s2b}): Using (\ref{s2}), we have \[ \sum_{n\le x} \log\frac{x}{n}= \lfloor x\rfloor\log x-\sum_{n\le x} \log n = (x+O(1)) \log x-(x\log x-x+O(\log x))=x+O(\log x). \] (\ref{s3}): By monotonicity, \[ \int_1^x \log^2t\,dt \le \sum_{n\le x} \log^2 n \le \int_1^{x+1} \log^2t\,dt. \] Now, \[ \int_1^x\log^2t\,dt=\int_0^{\log x} e^u u^2du=[e^u(u^2-2u+1)]_0^{\log x}=x(\log^2x-2\log x+1)-1.\] Combining these two facts, (\ref{s3}) follows. (\ref{s3b}): Using (\ref{s2}) and (\ref{s3}), we compute \begin{flalign*} \quad\sum_{n\le x} \log^2\frac{x}{n} &= \sum_{n\le x} (\log x-\log n)^2 & \\ \quad &= \lfloor x\rfloor\log^2x-2\log x(x\log x-x+O(\log x))+x(\log^2x-2\log x+1)+ O(\log^2x) & \\ \quad &= x+O(\log^2x) & \blacksquare \quad\end{flalign*} \vspace{.2cm} \begin{prop} \label{prop-mu2} The following estimates involving the M\"obius function hold as $x\rightarrow\infty$: \begin{eqnarray} \sum_{n\le x} \frac{\mu(n)}{n} &=& O(1), \label{mu1}\\ \sum_{n\le x} \frac{\mu(n)}{n}\log\frac{x}{n} &=& O(1), \label{mu2}\\ \sum_{n\le x} \frac{\mu(n)}{n}\log^2\frac{x}{n} &=& 2\log x+ O(1). \label{mu3} \end{eqnarray} \end{prop} {\noindent\it Proof. } (\ref{mu1}): If $f(x)=1$ then $F(x)=\lfloor x\rfloor$. M\"obius inversion (Lemma \ref{lem-mobius}) gives \begin{equation} 1=\sum_{n\le x}\mu(n)\left\lfloor\frac{x}{n}\right\rfloor=\sum_{n\le x}\mu(n)\left(\frac{x}{n}+O(1)\right) =x\sum_{n\le x}\frac{\mu(n)}{n}+\sum_{n\le x}O(1),\label{eq-mu}\end{equation} where we used $\mu(n)=O(1)$ (Proposition \ref{prop-mu1}(iii)). In view of $\sum_{n\le x}O(1)=O(x)$, we have $\sum_{n\le x}\mu(n)/n=O(x)/x=O(1)$. (\ref{mu2}): If $f(x)=x$ then $F(x)=\sum_{n\le x}x/n=x\log x+\gamma x+O(1)$ by (\ref{s1}). By M\"obius inversion, \[ x=\sum_{n\le x}\mu(n)\left( \frac{x}{n}\log \frac{x}{n}+\gamma \frac{x}{n}+O(1)\right) =x \sum_{n\le x}\frac{\mu(n)}{n}\log \frac{x}{n}+xO(1)+O(x), \] where we used (\ref{mu1}) and Proposition \ref{prop-mu1}(iii). From this we easily read off (\ref{mu2}). (\ref{mu3}): If $f(x)=x\log x$ then \begin{eqnarray*} F(x) &=& \sum_{n\le x} \frac{x}{n}\log \frac{x}{n}=\sum_{n\le x} \frac{x}{n}(\log x-\log n) =x\log x\sum_{n\le x}\frac{1}{n}-x\sum_{n\le x}\frac{\log n}{n} \\ &=& x\log x\left(\log x+\gamma+O\left(\frac{1}{x}\right)\right) -x\left(\frac{\log^2x}{2}+c+O\left(\frac{1+\log x}{x}\right)\right) \\ &=& \frac{1}{2}x\log^2x+\gamma x\log x-cx+O(1+\log x), \end{eqnarray*} by (\ref{s1}) and (\ref{s5}). Now M\"obius inversion gives \begin{eqnarray*} x\log x &=& \sum_{n\le x}\mu(n)\left(\frac{x}{2n}\log^2\frac{x}{n} +\gamma\frac{x}{n}\log \frac{x}{n}-c\frac{x}{n}+O(1+\log \frac{x}{n})\right)\\ &=& \frac{x}{2}\sum_{n\le x}\frac{\mu(n)}{n}\log^2\frac{x}{n} +xO(1)+xO(1)+O(x), \end{eqnarray*} where we used (\ref{mu1}), (\ref{mu2}) and (\ref{s2b}), and division by $x/2$ gives (\ref{mu3}). \qed \subsection{Conclusion} \begin{prop} [Selberg, Erd\"os-Karamata \cite{EK}] \label{prop-selb} Defining \[ K(1)=0,\quad\quad K(n)=\frac{1}{\log n}\sum_{d|n}\Lambda(d)\Lambda\left(\frac{n}{d}\right) \ \ \mathrm{if}\ \ n\ge 2,\label{eq-K}\] we have $K(n)\ge 0$ and \begin{equation} \sum_{n\le x} (\Lambda(n)+K(n))=2x+O\left(\frac{x}{\log x}\right). \label{eq-K2}\end{equation} \end{prop} {\noindent\it Proof. } The first claim is obvious in view of Proposition \ref{prop-Lambda1}(iv). We estimate \begin{eqnarray*} U(x) &:= &\sum_{n\le x}\sum_{d|n}\mu(d)\log^2\frac{n}{d} \ =\ \sum_{n\le x}\mu(n) \sum_{m\le x/n}\log^2 m \\ &=& \sum_{n\le x}\mu(n) \left( \frac{x}{n}(\log^2\frac{x}{n}-2\log\frac{x}{n}+1)+ O(\log^2\frac{x}{n})\right)\\ &=& x(2\log x+O(1))-2x O(1)+ xO(1)+ O(x)=2x\log x+O(x). \end{eqnarray*} Here we used (\ref{s3}), (\ref{mu3}), (\ref{mu2}), (\ref{mu1}), the fact $\mu(n)=O(1)$, and (\ref{s3b}). Comparing (\ref{eq-K2}) and (\ref{eq-selb}), we have \begin{eqnarray*} \sum_{n\le x} \Lambda(n)+K(n) &=& \sum_{2\le n\le x}\frac{1}{\log n} \sum_{d|n} \mu(d)\log^2\frac{n}{d}\\ &=& \int_{2-0}^x\frac{dU(t)}{\log t}\ =\ \left[\frac{U(t)}{\log t}\right]_2^x +\int_2^x \frac{U(t)}{t\log^2t}dt \\ &=& 2x+O\left(\frac{x}{\log x}\right)+ \int_2^x\frac{dt}{\log t}+O\left(\int_2^x\frac{dt}{\log^2t}\right). \end{eqnarray*} In view of the estimate \[ \int_2^x\frac{dt}{\log t}=\int_2^{\sqrt{x}}\frac{dt}{\log t}+\int_{\sqrt{x}}^x\frac{dt}{\log t} \le \frac{\sqrt{x}}{\log 2}+\frac{x}{\log\sqrt{x}}=O\left(\frac{x}{\log x}\right), \] we are done. \qed \begin{remark} In view of (\ref{eq-selb}), the above estimate $U(x)=2x\log x+O(x)$ is equivalent to \[ \sum_{n\le x} \Lambda(n)\log n+\sum_{ab\le x}\Lambda(a)\Lambda(b)=2x\log x+O(x),\] which is used in most Selberg-style proofs. (It would lead to (\ref{eq-balog}) with $k=2$.) \hfill{$\Box$}\medskip\end{remark} \begin{prop} \label{prop-G} If $g:[1,\infty)\rightarrow\7R$ is such that \begin{equation} G(x)=\sum_{n\le x}g\left(\frac{x}{n}\right) =Bx+C\frac{x}{\log x}+o\left(\frac{x}{\log x}\right) \label{eq-Kar2}\end{equation} then \begin{equation} g(x)\log x+\sum_{n\le x}\Lambda(n)\,g\left(\frac{x}{n}\right)= o(x\log x). \label{e2}\end{equation} \end{prop} {\noindent\it Proof. } In view of Lemma \ref{lem-TI1}, all we have to do is estimate \[ \sum_{n\le x}\mu(n)\log\frac{x}{n}\left( B\frac{x}{n}+C\frac{x}{n\log \frac{x}{n}} +o\left(\frac{x}{n\log \frac{x}{n}}\right) \right)=S_1+S_2+S_3. \] The three terms are \begin{eqnarray*} S_1 &=& Bx \sum_{n\le x}\frac{\mu(n)}{n}\log\frac{x}{n}=x O(1)=O(x), \\ S_2 &=& Cx \sum_{n\le x}\frac{\mu(n)}{n}=xO(1)=O(x), \\ S_3 &=& \sum_{n\le x}\mu(n) o\left(\frac{x}{n}\right) =\sum_{n\le x}o\left(\frac{x}{n}\right) =o\left(x\sum_{n\le x}\frac{1}{n}\right)=o(x\log x), \end{eqnarray*} where we used (\ref{mu2}), (\ref{mu1}), and $\mu(n)=O(1)$, respectively. \qed \noindent{\it Proof of Theorem \ref{theor-1}.} In view of $g(x)=f(x)-Ax$ and Proposition \ref{prop-estim} (i), (ii), we immediately have \begin{equation} g(x)=O(x), \quad\quad\quad \int_1^x\frac{dg(u)}{u}=O(1). \label{eq-g}\end{equation} Furthermore, since $f$ satisfies (\ref{eq-Kar}), and (\ref{s1}) gives $\sum_{n\le x} Ax/n=Ax\log x+A\gamma x+O(1)$, the M\"obius transform $G$ of $g(x)=f(x)-Ax$ satisfies (\ref{eq-Kar2}) (with a different $B$), so that Proposition \ref{prop-G} applies and (\ref{e2}) holds. Writing $N(x)=\sum_{n\le x}\Lambda(n)+K(n)$, by Proposition \ref{prop-selb} we have $N(x)=2x+\omega(x)$ with $\omega(x)=o(x)$. Now, \begin{eqnarray} \sum_{n\le x}(\Lambda(n)+K(n))g\left(\frac{x}{n}\right) &=& \int_{1-0}^xg\left(\frac{x}{t}\right)dN(t) \nonumber\\ &=& \left[ N(t)g\left(\frac{x}{t}\right)\right]_{1-0}^x-\int_1^x N(t)dg\left(\frac{x}{t}\right)\nonumber\\ &=& (N(x)g(1)-N(1-0)g(x)) +\int_1^x N\left(\frac{x}{u}\right) dg(u)\nonumber\\ &=& O(x) +\int_1^x\left(\frac{2x}{u}+o\left(\frac{2x}{u}\right)\right)dg(u) \nonumber\\ &=& O(x) +2x\int_1^x \frac{dg(u)}{u}+o\left(x\int_1^x \frac{dg(u)}{u}\right)\nonumber\\ &=& O(x)+O(x)+o(x)=O(x), \label{eq-x1} \end{eqnarray} where we used (\ref{eq-g}). On the other hand, \begin{eqnarray} \lefteqn{ \sum_{n\le x} (\Lambda(n)+K(n))\,\left|g\left(\frac{x}{n}\right)\right| = \int_{1-0}^x \left|g\left(\frac{x}{t}\right)\right| \, dN(t) } \nonumber\\ &=& 2\int_1^x \left|g\left(\frac{x}{t}\right)\right|\,dt+\int_{1-0}^x \left|g\left(\frac{x}{t}\right)\right|\,d\omega(t) \nonumber\\ &=& 2x\int_1^x \frac{|g(t)|}{t^2}dt -\int_{1-0}^x \omega(t)\, d\!\left|g\left(\frac{x}{t}\right)\right| +\left[g\left(\frac{x}{t}\right)\omega(t)\right]_{t=1-0}^{t=x}\nonumber\\ &=& 2x\int_1^x \frac{|g(t)|}{t^2}dt +\int_1^{x+0} \omega\left(\frac{x}{t}\right)\, d|g(t)| +g(1)\omega(x)-g(x+0)\omega(1-0).\label{eq-x2} \end{eqnarray} In view of $g(x)=O(x)$ and $\omega(x)=o(x)$, the sum of the last two terms is $O(x)$. Furthermore, \begin{eqnarray*} \int_1^{x+0} \omega\left(\frac{x}{t}\right)\,d|g(t)| &=& o\left( x\int_1^{x+0}\frac{|d|g(t)||}{t}\right) \le o\left( x\int_1^{x+0}\frac{|dg(t)|}{t}\right) \\ &\le& o\left( x\int_1^{x+0}\frac{df+Adt}{t}\right)=o(x\log x), \end{eqnarray*} where we used $g(x)=f(x)-Ax$ and $df=|df|$ (since $f$ is non-decreasing) to obtain $|dg|=|df-Adt|\le |df|+Adt=df+Adt$ and Proposition \ref{prop-estim}(ii). Plugging this into (\ref{eq-x2}), we have \begin{equation} \sum_{n\le x} (\Lambda(n)+K(n))\,\left|g\left(\frac{x}{n}\right)\right| = 2x\int_1^x \frac{|g(t)|}{t^2}dt+ o(x\log x). \label{eq-x3}\end{equation} After these preparations, we can conclude quickly: Subtracting (\ref{eq-x1}) from (\ref{e2}) we obtain \[ g(x)\log x= \sum_{n\le x} K(n) g\left(\frac{x}{n}\right) + o(x\log x).\] Taking absolute values of this and of (\ref{e2}) while observing that $\Lambda$ and $K$ are non-negative, we have the inequalities \[ |g(x)|\log x\le \sum_{n\le x} \Lambda(n) \left|g\left(\frac{x}{n}\right)\right| + o(x\log x), \quad\quad |g(x)|\log x\le \sum_{n\le x} K(n) \left|g\left(\frac{x}{n}\right)\right| + o(x\log x).\] Adding these inequalities and comparing with (\ref{eq-x3}) we have \[ 2|g(x)|\log x \le \sum_{n\le x}(\Lambda(n)+K(n)) \left|g\left(\frac{x}{n}\right)\right| + o(x\log x) = 2x\int_1^x \frac{|g(t)|}{t^2}dt+ o(x\log x), \] so that (\ref{eq-ineq}), and with it Theorem \ref{theor-1}, is obtained dividing by $2x\log x$. \qed \begin{remark} 1. We did not use the full strength of Proposition \ref{prop-selb}, but only an $o(x)$ remainder. 2. Inequality (\ref{eq-ineq}) is the special case $k=1$ of the more general integral inequality \begin{equation} \frac{|g(x)|}{x}\log^kx\le k\int_1^x\frac{|g(t)|\log^{k-1}t}{t^2}dt+O(\log^{k-c}x) \quad\forall k\in\7N \label{eq-balog}\end{equation} proven in \cite{balog}, assuming a $O\left(\frac{x}{\log^2x}\right)$ in (\ref{eq-Kar}) instead of $C\frac{x}{\log x}+o\left(\frac{x}{\log x}\right)$. \hfill{$\Box$}\medskip\end{remark} \section{Proof of Theorem \ref{theor-2}}\label{s-th2} The proof will be based on the following proposition, to be proven later: \begin{prop} \label{prop-E} If $s:[0,\infty)\rightarrow\7R$ satisfies \begin{equation} e^{t'}s(t')-e^t s(t)\ge -M(e^{t'}-e^t) \quad \forall t'\ge t\ge 0,\label{eq-s1}\end{equation} \begin{equation} \left|\int_0^x s(t)dt\right|\le M'\quad\forall x\ge 0, \label{eq-s2}\end{equation} and $S=\lim\sup|s(x)|>0$ then there exist numbers $0<S_1<S$ and $e,h>0$ such that \begin{equation} \mu(E_{x,h,S_1}) \ge e\quad \forall x\ge 0,\quad \mathrm{where}\quad E_{x,h,S_1}=\{ t\in [ x,x+h]\ | \ |s(t)|\le S_1 \},\label{eq-E}\end{equation} and $\mu$ denotes the Lebesgue measure. \end{prop} \noindent{\it Proof of Theorem \ref{theor-2} assuming Proposition \ref{prop-E}.} It is convenient to replace $g:[1,\infty)\rightarrow\7R$ by $s:[0,\infty)\rightarrow\7R,\ s(t)= e^{-t} g(e^t)$. Now $s$ is locally integrable, and the assumptions (\ref{et1}) and (\ref{et2}) become (\ref{eq-s1}) and (\ref{eq-s2}), respectively, whereas the conclusions (\ref{et3}) and (\ref{et4}) assume the form \begin{equation} S=\limsup_{t\rightarrow\infty}|s(t)|<\infty, \label{eq-s3}\end{equation} \begin{equation} S>0 \quad\Rightarrow\quad \limsup_{x\rightarrow\infty} \frac{1}{x}\int_0^x |s(t)|dt < S. \label{eq-s4}\end{equation} The proof of (\ref{eq-s3}) is easy: Dividing (\ref{eq-s1}) by $e^{t'}$ and integrating over $t'\in[t,t+h]$, where $h>0$, one obtains \[ \int_t^{t+h} s(t')dt' - s(t)(1-e^{-h})\ge -Mh+M(1-e^{-h}), \] and using $|\int_a^b s(t)dt|\le |\int_0^as(t)dt|+|\int_0^bs(t)dt|\le 2M'$ by (\ref{eq-s2}), we have the upper bound \[ s(t)\le \frac{2M'+M(e^{-h}-1+h)}{1-e^{-h}} . \] Similarly, dividing (\ref{eq-s1}) by $e^t$ and integrating over $t\in[t'-h,t']$, one obtains the lower bound \[ -\frac{2M'+M(e^h-1-h)}{e^h-1} \le s(t),\] thus (\ref{eq-s3}) holds. Assuming $S>0$, let $S_1,h,e$ as provided by Proposition \ref{prop-E}. For each $\widehat{S}>S$ there is $x_0$ such that $x\ge x_0\Rightarrow |s(x)|\le\widehat{S}$. Given $x\ge x_0$ and putting $N=\left\lfloor\frac{x-x_0}{h}\right\rfloor$, we have \begin{eqnarray*} \int_0^x |s(t)|dt &=& \int_0^{x_0}|s(t)|dt+ \sum_{n=1}^N \int_{x_0+(n-1)h}^{x_0+nh} |s(t)|dt +\int_{x_0+Nh}^x |s(t)|dt \\ &\le & 2M'+ N [ \widehat{S}(h-e)+S_1e] + 2M' \\ &=& \left(\frac{x-x_0}{h}+O(1)\right) h \left[\left(1-\frac{e}{h}\right)\widehat{S}+\frac{e}{h}S_1\right] +4M'\\ &=& x \left[\left(1-\frac{e}{h}\right)\widehat{S}+\frac{e}{h}S_1\right] +O(1). \end{eqnarray*} Thus \[ \limsup_{x\rightarrow\infty}\frac{1}{x}\int_0^x |s(t)|dt\le \left(1-\frac{e}{h}\right)\widehat{S}+\frac{e}{h}S_1. \] Since $S_1<S$ and since $\widehat{S}>S$ can be chosen arbitrarily close to $S$, (\ref{eq-s4}) holds and thus Theorem \ref{theor-2}. \qed In order to make plain how the assumptions (\ref{eq-s1}) and (\ref{eq-s2}) enter the proof of Proposition \ref{prop-E}, we prove two intermediate results that each use only one of the assumptions. For the first we need a ``geometrically obvious'' lemma of isoperimetric character: \begin{lemma} Let $t_1<t_2,\ C_1>C_2>0$ and $k:[t_1,t_2]\rightarrow\7R$ non-decreasing with $k(t_1)\ge C_1e^{t_1}$ and $k(t_2)\le C_2e^{t_2}$. Then \[ \mu\left(\{ t\in[t_1,t_2] \ | \ C_2e^t\le k(t)\le C_1 e^t\}\right) \ge \log\frac{C_1}{C_2}.\] \end{lemma} {\noindent\it Proof. } As a non-decreasing function, $k$ has left and right limits $k(t\pm 0)$ everywhere and $k(t-0)\le k(t)\le k(t+0)$. The assumptions imply $t_1\in A:=\{ t\in[t_1,t_2] \ | \ k(t)\ge C_1e^t\}$, thus we can define $T_1=\sup(A)$. Quite obviously we have $t>T_1\Rightarrow k(t)<C_1e^t$, which together with the non-decreasing property of $k$ and the continuity of the exponential function implies $k(T_1+0)\le C_1e^{T_1}$ (provided $T_1<t_2$). We have $T_1\in A$ if and only if $k(T_1)\ge C_1e^{T_1}$. If $T_1\not\in A$ then $T_1>t_1$, and every interval $(T_1-\varepsilon,T_1)$ (with $0<\varepsilon<T_1-t_1$) contains points $t$ such that $k(t)\ge C_1e^t$. This implies $k(T_1-0)\ge C_1e^{T_1}$. Now assume $T_1=t_2$. If $T_1\in A$ then $C_1e^{T_1}\le k(T_1)\le C_2e^{T_1}$. If $T_1\not\in A$ then $C_1e^{T_1}\le k(T_1-0)\le k(T_1)\le C_2e^{t_2}$. In both cases we arrive at a contradiction since $C_2<C_1$. Thus $T_1<t_2$. If $T_1\in A$ (in particular if $T_1=t_1$) then $C_1e^{T_1}\le k(T_1)\le k(T_1+0)\le C_1e^{T_1}$. Thus $k$ is continuous from the right at $T_1$ and $k(T_1)=C_1e^{T_1}$. If $T_1\not\in A$ then $T_1>t_1$ and $C_1e^{T_1}\le k(T_1-0)\le k(T_1+0)\le C_1e^{T_1}$. This implies $k(T_1)=C_1e^{T_1}$, thus the contradiction $T_1\in A$. Thus we always have $T_1\in A$, thus $k(T_1)=C_1e^{T_1}$. Now let $B=\{ t\in[T_1,t_2]\ | \ k(t)\le C_2e^t\}$. We have $t_2\in B$, thus $T_2=\inf(B)$ is defined and $T_2\ge T_1$. Arguing similarly as before we have $t<T_2\Rightarrow k(t)>C_2e^t$, implying $k(T_2-0)\ge C_2e^{T_2}$. And if $T_2<t_2$ and $T_2\not\in B$ then $k(T_2+0)\le C_2e^{T_2}$. If $T_2\in B$ (in particular if $T_2=t_2$) then $C_2e^{T_2}\le k(T_2-0)\le k(T_2)\le C_2e^{T_2}$, implying $k(T_2-0)=k(T_2)=C_2e^{T_2}$ so that $k$ is continuous from the left at $T_2$. If $T_2\not\in B$ then $T_2<t_2$ and $C_2e^{T_2}\le k(T_2-0)\le k(T_2+0)\le C_2e^{T_2}$, implying $k(T_2)=C_2e^{T_2}$ and thus a contradiction. Thus we always have $T_2\in B$, thus $k(T_2)=C_2e^{T_2}$. By the above results, we have $C_2e^t\le k(t)\le C_1 e^t\ \forall t\in[T_1,T_2]$ and thus \begin{equation} \mu\left(\{ t\in[t_1,t_2] \ | \ C_2e^t\le k(t)\le C_1 e^t\}\right) \ge T_2-T_1. \label{mu}\end{equation} Using once more that $k$ is non-decreasing, we have \[ C_1e^{T_1}= k(T_1)\le k(T_2)=C_2e^{T_2}, \] implying $T_2-T_1\ge \log\frac{C_1}{C_2}$, and combining this with (\ref{mu}) proves the claim. \qed \begin{coro} \label{coro-E2} Assume that $s:[0,\infty)\rightarrow\7R$ satisfies (\ref{eq-s1}) and $s(t_1)\ge S_1\ge S_2\ge s(t_2)$, where $S_2+M>0$. Then \[ \mu(\{t\in [t_1,t_2]\ | \ s(t)\in [S_2,S_1] \}) \ge\log\frac{S_1+M}{S_2+M}. \] \end{coro} {\noindent\it Proof. } We note that (\ref{eq-s1}) is equivalent to the statement that the function $k:t\mapsto e^t(s(t)+M)$ is non-decreasing. The assumption $s(t_1)\ge S_1\ge S_2\ge s(t_2)$ implies $k(t_1)\ge (S_1+M)e^{t_1}$ and $k(t_2)\le(S_2+M)e^{t_2}$. Now the claim follows directly by an application of the preceding lemma. \qed \begin{lemma} \label{lem-E1} Let $s:[0,\infty)\rightarrow\7R$ be integrable over bounded intervals, satisfying (\ref{eq-s2}). Let $e>0$ and $0<S_2<S_1$ be arbitrary, and assume \begin{equation} h\ge 2\left( e+\frac{M'}{S_1}+\frac{M'}{S_2}\right).\label{eq-h}\end{equation} Then every interval $[x,x+h]$ satisfies at least one of the following conditions: \begin{itemize} \item[(i)] $\displaystyle\mu(E_{x,h,S_1}) \ge e$, where $E_{x,h,S_1}$ is as in (\ref{eq-E}), \item[(ii)] there exist $t_1,t_2$ such that $x\le t_1<t_2\le x+h$ and $s(t_1)\ge S_1$ and $s(t_2)\le S_2$. \end{itemize} \end{lemma} {\noindent\it Proof. } It is enough to show that falsity of (i) implies (ii). Define \[ T=\sup\{ t\in[x,x+h]\ | \ s(t)\le S_2\}, \] with the understanding that $T=x$ if $s(t)>S_2$ for all $t\in[x,x+h]$. Then $s(t)>S_2\ \forall t\in(T,x+h]$, which implies \[ (x+h-T)S_2\le \int_{T}^{x+h} s(t)dt\le 2M' \] and therefore \begin{equation} x+h-T\ \le \frac{2M'}{S_2}. \label{eq-t2}\end{equation} We observe that (\ref{eq-t2}) with $T=x$ would contradict (\ref{eq-h}). Thus $x<T\le x+h$, so we can indeed find a $t_2\in[x,x+h]$ with $s(t_2)\le S_2$. Since we do not assume continuity of $s$, we cannot claim that we may take $t_2=T$, but by definition a $t_2$ can be found in $(T-\varepsilon,T]$ for every $\varepsilon>0$. Now we claim that there is a point $t_1\in[x,t_2]$ such that $s(t_1)\ge S_1$. Otherwise, we would have $s(t)<S_1$ for all $t\in[x,t_2]$. By definition, $|s(t)|\le S_1$ for $t\in E_{x,h,S_1}$, thus $|s|>S_1$ on the complement of $E_{x,h,S_1}$. Combined with $s(t)<S_1$ for $t\in[x,t_2]$, this means $s(t)<-S_1$ whenever $t\in[x,t_2]\backslash E_{x,h,S_1}$. Thus \begin{eqnarray*} \int_x^{t_2} s(t)dt &\le& S_1\mu([x,t_2]\cap E_{x,h,S_1})-S_1 \mu([x,t_2]\backslash E_{x,h,S_1}) \\ &=& -S_1(t_2-x) + 2S_1\mu([x,t_2]\cap E_{x,h,S_1}) \\ &=& S_1(x-t_2+2\mu([x,t_2]\cap E_{x,h,S_1})) \end{eqnarray*} In view of (\ref{eq-t2}) and $t_2>T-\varepsilon$ (with $\varepsilon>0$ arbitrary), we have $x-t_2<x-T+\varepsilon\le 2M'/S_2-h+\varepsilon$, thus we continue the preceding inequality as \begin{eqnarray*} \cdots & < & S_1\left(\frac{2M'}{S_2}-h+\varepsilon + 2\mu([x,t_2]\cap E_{x,h,S_1})\right) \end{eqnarray*} By our assumption that (i) is false, we have $\mu([x,t_2]\cap E_{x,h,S_1})\le\mu(E_{x,h,S_1})<e$. Thus choosing $\varepsilon$ such that $0<\varepsilon<2(e-\mu([x,t_2]\cap E_{x,h,S_1}))$, we have \begin{eqnarray*} \cdots &<& S_1\left(\frac{2M'}{S_2}-h+2e\right). \end{eqnarray*} Combining this with (\ref{eq-h}), we finally obtain $\int_x^{t_2} s(t)dt<-2M'$, which contradicts the assumption (\ref{eq-s2}). Thus there is a point $t_1\in[x,t_2]$ such that $s(t_1)\ge S_1$. In view of $s(t_1)\ge S_1>S_2\ge s(t_2)$, we have $t_1\ne t_2$, thus $t_1<t_2$. \qed \noindent{\it Proof of Proposition \ref{prop-E}.} Assuming that $S=\limsup |s(x)|>0$, choose $S_1, S_2$ such that $0<S_2<S_1<S$. Then $e:=\log\frac{S_1+M}{S_2+M}>0$. Let $h$ satisfy (\ref{eq-h}). Assume that there is an $x\ge 0$ such that $\mu(E_{x,h,S_1})<e$. Then Lemma \ref{lem-E1} implies the existence of $t_1,t_2$ such that $x\le t_1<t_2\le x+h$ and $s(t_1)\ge S_1, \ s(t_2)\le S_2$. But then Corollary \ref{coro-E2} gives $\mu([t_1,t_2]\cap s^{-1}([S_2,S_1])\ge \log\frac{S_1+M}{S_2+M}$. Since $[t_1,t_2]\cap s^{-1}([S_2,S_1])\subset E_{x,h,S_1}$, we have $\mu(E_{x,h,S_1})\ge\log\frac{S_1+M}{S_2+M}=e$, which is a contradiction. \qed \begin{remark} The author did not succeed in making full sense of the proof in \cite{karamata} corresponding to that of Corollary \ref{coro-E2}. It seems that there is a logical mistake in the reasoning, which is why we resorted to the above more topological approach. \hfill{$\Box$}\medskip\end{remark}
\section{Introduction} The development of a cosmological first-order phase transition via nucleation and expansion of bubbles provides an interesting scenario for the formation of cosmological objects such as magnetic fields \cite{gr01}, topological defects \cite{vs94}, baryon inhomogeneities \cite{w84,h95,ma05}, gravitational waves \cite{gw,lms12}, or the baryon asymmetry of the universe \cite{ckn93}. One of the relevant aspects of the dynamics of a first-order phase transition is the motion of the transition fronts. In most cases, stationary solutions exist, and the bubble growth reaches a terminal velocity shortly after the bubble nucleates. The velocity of bubble walls depends on the friction with the plasma \cite{dlhll,micro} and on hydrodynamics \cite{hidro}. Thus, for a given set of parameters, there can be one or more solutions. For instance, the wall may propagate as a supersonic detonation or as a subsonic deflagration. The cosmological consequences of a phase transition depend strongly on the wall velocity. For example, detonations favor the generation of gravitational waves, whereas weak deflagrations favor electroweak baryogenesis. It is well known that these hydrodynamic solutions may be unstable \cite{landau,link,hkllm,abney,r96}. Such instabilities would have important implications for the cosmological remnants of the phase transition. The standard approach to the stability of combustion or phase transition fronts is to consider small perturbations of the wall surface and the fluid \cite{landau}. In the cosmological context, the stability of deflagrations was first studied in Ref. \cite{link}, in the non-relativistic limit. This analysis was improved in Ref. \cite{hkllm} by considering relativistic velocities, and by taking into account the dependence of the velocity of phase transition fronts on temperature. The latter is the most important difference with previous analysis, since temperature fluctuations cause velocity fluctuations which may stabilize the wall. Numerical simulations \cite{fa03} agree with this stabilization. A very simple expression for the wall velocity was considered in Ref. \cite{hkllm}, namely, $v_w\propto T_c-T_+$, where $T_c$ is the critical temperature and $T_+$ is the temperature outside the bubble. Perturbing this equation gives an equation for the perturbations of the interface, which involves the temperature fluctuations $\delta T_+$. We wish to point out, however, that such a simple form of the wall velocity does not take into account (among other things) the dependence on temperature perturbations $\delta T_-$ behind the wall (which were otherwise considered in \cite{hkllm} as independent from $\delta T_+$). Furthermore, the results were applied to a specific case (the electroweak phase transition) without taking into account the fact that the temperature $T_+$ is higher than the nucleation temperature $T_N$, due to reheating in front of the wall. These issues are not relevant in the case of small wall velocity, small latent heat, and little supercooling, but can be important otherwise. The discussion on the electroweak phase transition in Ref. \cite{hkllm} was carried out for the minimal standard model (SM) with a Higgs mass $m_H=40GeV$. This gives a rather weak phase transition (although strong enough for baryogenesis), for which the aforementioned approximations are valid. The deflagration was found to be stable for wall velocities above a critical value $v_c=0.07$. Thus, arguing that the wall velocity is generally $v_w\gtrsim 0.1$, it was concluded that the deflagration is commonly stable. However, the wall velocity depends on the details of the specific model, and may be smaller. Furthermore, many extensions of the SM give stronger phase transitions, which would give higher values of $v_c$. In this paper we aim to perform a more complete calculation, and to investigate a wider range of parameters. The main improvements of our treatment will be to consider a more realistic equation for the wall velocity, which depends on the fluid fluctuations on both sides of the interface, and to take into account the effects of reheating, i.e., the fact that the temperature $T_+$ in front of the wall is not a boundary condition for the deflagration. In the limit of small wall velocities, little supercooling, and small latent heat, our results agree with those of Ref. \cite{hkllm}. Increasing the latent heat or the wall velocity does not change the qualitative picture. However, as the amount of supercooling is increased, the critical velocity increases much more quickly than what found in Ref. \cite{hkllm}, even though the reheating effect tends to stabilize the deflagration. As a consequence, for strong enough supercooling the wall propagation is unstable for any velocity below the speed of sound. moreover, the instability is stronger at the speed of sound. This is in contradiction with the results of Ref. \cite{hkllm}. The origin of the discrepancy is in the fact that, in our treatment, the wall velocity depends on fluid fluctuations on both sides of the wall. The paper is organized as follows. In the next section we review the stationary motion of a phase transition front. In Sec. \ref{staban} we consider linear perturbations of the interface and the fluid. We discuss the approaches and results of previous works, and then we derive the equations for the perturbations and find the general solution. In Sec. \ref{stab} we study the possible instabilities of the deflagration and compare our results with previous works. We find analytical approximations for the case of small velocity, small latent heat, and little supercooling. We also discuss how the reheating which occurs in front of the wall affects the stability of a deflagration. In Sec. \ref{numres} we use the bag equation of state to study the instability as a function of the relevant parameters. We explore numerically a wide region of parameter space. In Sec. \ref{conseq} we consider the dynamics of the instabilities in a cosmological phase transition, and in Sec. \ref{model} we discuss the results for the specific case of the electroweak phase transition. We also discuss briefly on some cosmological effects. Finally, in Sec. \ref{conclu} we summarize our conclusions. \section{Phase transition dynamics and stationary wall motion} \label{stationary} Cosmological phase transitions are generally a consequence of the high temperature behaviour of a theory with spontaneous symmetry breaking. Macroscopically, the system can be described by a relativistic fluid and a scalar field $\phi$ which acts as an order parameter. The free energy density $\mathcal{F}(\phi,T)$ has different minima $\phi_+$ and $\phi_-$ at high and low temperatures, respectively. These minima characterize two different phases. For instance, in the case of the electroweak phase transition, $\phi$ corresponds to the expectation value of the Higgs field, and we have $\phi_+=0$, $\phi_-\sim T_c\sim 100\mathrm{GeV}$. If the phase transition is first-order, there is a range of temperatures at which these two minima coexist separated by a barrier. Thus, the metastable phase is characterized by the free energy density $ \mathcal{F}_{+}(T)= \mathcal{F}(\phi_+,T)$, whereas the stable phase is characterized by $\mathcal{F}_{-}(T)= \mathcal{F}(\phi _{-},T)$. The pressure in each phase is given by $p_\pm=-\mathcal{F}_\pm$, the entropy density by $s_\pm=dp_\pm/dT$, and the energy density by $e_\pm=Ts_\pm-p_\pm$. The critical temperature $T_c$ is defined by $\mathcal{F}_{+}(T_{c})=\mathcal{F}_{-}(T_{c})$. The latent heat is defined as the energy discontinuity at $T=T_c$, and is given by $L =T_c[\mathcal{F}_-'(T_c)-\mathcal{F}_+'(T_c)]$. A first-order phase transition is characterized by the supercooling of the system (which remains in the metastable phase below $T_c$), followed by the nucleation and growth of bubbles of the stable phase at a temperature $T_N<T_c$ (see, e.g., \cite{gw81,ah92,m00}). The latent heat is released at the phase transition fronts, which are the bubble walls. We are interested in the motion of the latter. Therefore, we shall consider the hydrodynamics of two phases separated by a moving interface. The equations for the wall and the fluid variables can be derived from the conservation of the stress tensor for the scalar field and the fluid. These can be written in the form \bega \pa_\mu\left(-T\fr{\pa\cf}{\pa T}u^\mu u^\nu+g^{\mu\nu} \cf\right)+\pa_\mu\pa^\mu\phi\pa^\nu\phi=0\label{EM1.1},\\ \pa_\mu\pa^\mu\phi+\fr{\pa\cf}{\pa\phi}+\tilde\eta T_c u^\mu\pa_\mu\phi=0\label{EM1.2}, \ena with $u^\mu=(\ga,\ga\bv)$ the four velocity of the fluid and $g^{\mu\nu}$ the Minkowsky metric tensor. The terms in parenthesis in Eq. (\ref{EM1.1}) give the well known stress tensor of a relativistic fluid \bega T^{\mu\nu}=wu^\mu u^\nu-p g^{\mu\nu}, \label{Tmunu} \ena where $w$ is the enthalpy and $p$ the pressure. The last term in Eq. (\ref{EM1.1}) gives the transfer of energy between the plasma and the field. The motion of the latter is governed by Eq. (\ref{EM1.2}). The last term in this equation is a phenomenological damping term\footnote{Recently \cite{ekns10,hs13,ariel13}, different forms of the damping term have been proposed in order to account for the saturation of the friction force at ultra-relativistic velocities \cite{bm09}. Since we shall only deal with deflagrations, the damping term in Eq. (\ref{EM1.2}) is a good approximation \cite{ariel13,ms12}.}. The dimensionless coefficient $\tilde\eta$ can be obtained from microphysics calculations (in the general case, $\tilde\eta$ may depend on the field $\phi$). For the macroscopic treatment, it is a good approximation to consider an infinitely thin interface. The system of equations (\ref{EM1.1}) then gives the fluid equations on either of the phases (where the field is constant), as well as the connection between the solutions at each side of the interface. On the other hand, Eq. (\ref{EM1.2}) gives an equation for the interface itself and the forces acting on it. Due to the friction with the surrounding plasma, the bubble walls in general reach a terminal velocity\footnote{In Ref. \cite{bm09}, it was shown that, if the wall reaches ultra-relativistic velocities, it may enter a stage of continuous acceleration. However, models which allow such ultra-relativistic velocities will not allow, in general, deflagrations. We are not interested in such models.}. We shall now consider the stationary motion. In the next section we shall study perturbations of the stationary solutions. For simplicity, we shall assume a planar interface moving towards the positive $z$ axis. \subsection{Fluid equations} \label{hydrostat} For planar symmetry, the problem becomes $(1+1)$-dimensional, and we need only consider the $z$ component of the fluid velocity, which we shall denote $v(z,t)$. Within each phase the field is a constant, and Eq. (\ref{EM1.1}) just gives the conservation of energy and momentum of the fluid, $\pa_\mu T^{\mu\nu}=0$, with $T^{\mu\nu}$ given by Eq. (\ref{Tmunu}). The absence of a distance scale in these equations justifies to assume the \emph{similarity condition}, namely, that quantities depend only on the variable $\xi=z/t$. Thus, we have $\partial _t=-(\xi/t)d/d\xi$ and $\partial _z=(1/t)d/d\xi$. Furthermore, variations of thermodynamical quantities are related by the speed of sound $c_s^2=dp/de$. We have, e.g., \beg dp=dw/(1+c_s^{-2}). \label{pwcs} \en We may thus obtain an equation for $v(\xi)$ which depends only on the parameter $c_s$ \cite{landau} ($c_s$ depends on the equation of state (EOS) and will be in general a function of temperature). In the planar case this equation is very simple (see e.g. \cite{lm11}). The solutions are either $v=$ constant, or the particular solution $v_{\mathrm{rar}}(\xi)=(\xi-c_s)/(1-\xi c_s)$. The latter corresponds to a rarefaction wave. In this paper we shall be interested in the constant velocity solutions. For these, the temperature is also a constant. \subsection{Matching conditions} The fluid solutions on each side of the bubble wall can be linked by integrating Eq. (\ref{EM1.1}) across the wall. It is convenient to consider a reference frame moving with the wall, where all time derivatives vanish. We obtain two equations, $\partial_z T^{z0}=0$, $\partial_zT^{zz}=0$, and the integration gives simply \bega w_-v_-\ga^2_-&=&w_+v_+\ga^2_+,\label{EM3.1}\\ w_-v_-^2\ga^2_- +p_-&=&w_+v^2_+\ga^2_+ +p_+ , \label{EM3.2} \ena where $+$ and $-$ signs refer to variables just in front and just behind the wall, respectively. Notice that in this frame the fluid velocity is negative (Fig. \ref{figwall}). \begin{figure}[bth] \centering \epsfysize=2cm \leavevmode \epsfbox{wall.eps} \caption{Sketch of a deflagration in the wall frame.} \label{figwall} \end{figure} These equations have two branches of solutions, called detonations and deflagrations. Detonations are characterized by the relation $|v_+|>|v_-|$, whereas deflagrations are characterized by $|v_+|<|v_-|$. For detonations, the curve of $|v_+|$ vs $|v_-|$ has a minimum at the Jouguet point $|v_-|=c_{s-}$, where $|v_+|$ takes a value $v_J^{\mathrm{det}}>c_{s+}$ (hence, for detonations the incoming flow is supersonic). For deflagrations, the curve of $|v_+|$ vs $|v_-|$ has a maximum value $v_J^{\mathrm{def}}$ at the Jouguet point $c_{s-}$. In this case, $v_+$ is subsonic, $v_J^{\mathrm{def}}<c_{s+}$. Detonations are called weak if $v_-$ is supersonic as well as $v_+$, and deflagrations are called weak if $v_-$ is subsonic as well as $v_+$. If one of the velocities is supersonic and the other one subsonic, then the hydrodynamic process is called strong. \subsection{Fluid profiles} The profiles of the fluid velocity and temperature must be constructed from the solutions of the fluid equations on each side of the wall, using the matching conditions at the wall and the boundary conditions. The latter correspond to vanishing fluid velocity far behind and far in front of the wall. The value of the temperature $T_N$ far in front of the wall is also a boundary condition. We shall now describe briefly the possible fluid profiles. For details, see e.g. \cite{lm11}. Let us call $\tilde v _+$ and $\tilde v_-$ the values of the fluid velocity on each side of the phase discontinuity. For a detonation, the fluid velocity $\tilde v _+$ vanishes in front of the wall, which moves supersonically, i.e., we have $v_w=|v_+|\geq v_J^{\mathrm{det}}>c_{s+}$. Behind the wall, we have a non-vanishing velocity $\tilde v _-$, and the wall is followed by a rarefaction wave. It turns out that only weak detonations can fulfil the boundary conditions. Therefore, strong detonations are not possible. For a subsonic wall, the rarefaction solution $v_{\mathrm{rar}}(\xi)$ cannot be accommodated in the velocity profile. The fluid velocity vanishes behind the wall ($\tilde v _-=0$), and the hydrodynamic process is a weak deflagration, with $|v_-|=v_w<c_{s-}$. In front of the wall, the velocity is a constant up to a certain point where the velocity vanishes abruptly (see Fig. \ref{figdefla}). Such a discontinuity without change of phase is called a shock front. At the shock discontinuity, Eqs. (\ref{EM3.1}-\ref{EM3.2}) still apply, but now the enthalpy and pressure are related by the same EOS on both sides of the interface. These equations give the temperature $T_+$ as a function of the boundary condition $T_N$, as well as the velocity of the shock front. \begin{figure}[hbt] \centering \epsfysize=4cm \leavevmode \epsfbox{defla.eps} \caption{Sketch of fluid velocity and temperature profiles for a deflagration, in the reference frame of the bubble center.} \label{figdefla} \end{figure} Between weak deflagrations and detonations there is a velocity gap $c_{s-}<v_w< v_J^{\mathrm{det}}$. The ``traditional'' weak deflagration profile described above is in principle possible for a supersonic wall as well. In such a case, the hydrodynamic solution is a strong deflagration, with $|v_-|=v_w>c_{s-}$. Numerical calculations \cite{ikkl94} suggest that such a strong deflagration is unstable since, if set as initial condition, it evolves to other stationary solutions. A supersonic deflagration can also be constructed using the solution $v_{\mathrm{rar}}$ behind the wall (see Fig. \ref{figstrong}). Instead of $|v_-|=v_w$, in this case the matching conditions only require $|v_-|\geq c_{s-}$ (for details, see, e.g., \cite{lm11}). Therefore, this kind of solution is either a strong or a Jouguet deflagration. In either case, since the fluid velocity $\tilde{v}_-$ does not vanish, the wall velocity is given by the relativistic sum of $|v_-|$ and $\tilde{v}_-$, and is always supersonic. In order to avoid a strong deflagration, the only possibility is a Jouguet solution, $v_-=-c_{s-}$. As shown in Fig. \ref{figstrong} (right panel), in this case the rarefaction begins immediately at the wall. Thus, the strong deflagration shown in the left panel is an intermediate case between the traditional strong deflagration of Fig. \ref{figdefla} and the supersonic Jouguet deflagration. Notice that we have a family of solutions depending on two free parameters, namely $v_-$ and $\tilde{v}_-$. For $v_-=-v_w$ we obtain a reduced family of strong deflagrations, namely, those with the traditional profile, whereas for $v_-=-c_s$ we obtain the family of supersonic Jouguet deflagrations. Fixing the latter condition and varying the value of $\tilde{v}_-$, the velocity of the Jouguet deflagration fills the range of $v_w$ between the weak deflagration and the detonation. \begin{figure}[hbt] \centering \epsfysize=3.5cm \leavevmode \epsfbox{strong.eps} \caption{Sketch of fluid velocity profile for strong and Jouguet deflagrations.} \label{figstrong} \end{figure} \subsection{Equation for the interface} We shall now obtain an equation for the interface from Eq. (\ref{EM1.2}). In the reference frame of the wall, we multiply by $\phi'\equiv d\phi/dz$ and then integrate across the interface. We obtain \beg\label{ecfric} p_- - p_+ + \int_-^+ \left(-\fr{\pa\cf}{\pa T}\right)\fr{dT}{dz}dz +\tilde\eta T_c \int_-^+ \phi'^2 \ga v dz=0 , \en where we have used the relation $(\pa \cf / \pa \phi) d\phi/dz=d\cf / dz - (\pa\cf/\pa T)dT/dz$. In Eq. (\ref{ecfric}) we identify the force which drives the bubble expansion, \beg F_{\mathrm{dr}} = p_-(T_-)-p_+(T_+)+\int_-^+ \frac{\partial\mathcal{F}}{\pa T} \frac{d T}{dz}dz , \label{fdrexac} \en and the friction force, which explicitly depends on the velocity of the fluid with respect to the wall. Thus, Eq. (\ref{ecfric}) gives the balance between the forces that rule the stationary wall motion. While the integration of Eqs. (\ref{EM1.1}) across the wall was straightforward, in this case we shall need some approximations to avoid the dependence on the wall shape \cite{ms09}. To evaluate the last integral in Eq. (\ref{ecfric}), we notice that $\phi'(z)^2$ behaves approximately like a delta function which picks the value of $\gamma v$ at the center of the wall width, say $\gamma_0 v_0$. Thus, the integral gives $\gamma_0v_0\sigma$, where $\sigma\equiv\int\phi'^2dz$ is the surface tension. We shall approximate the value of $\gamma_0 v_0$ by the average $\lan \gamma v\ran\equiv \fr{1}{2}(\gamma_+v_+ +\gamma_-v_-)$. Thus, the last term in (\ref{ecfric}) can be written in the form $-\eta \lan \gamma v\ran$, with \beg \eta=\tilde\eta T_c\sigma , \en and the force balance reads \beg\label{solfric} {\eta} \lan \ga v\ran=-{F_{\mathrm{dr}}}. \en For the first integral in Eq. (\ref{ecfric}), we can use a linear approximation for the $z$ dependence of the entropy density $s=-\partial \mathcal{F}/\partial T$ inside the wall. We obtain $\fr{1}{2}(s_+ +s_-)(T_+ -T_-)$. This gives the following approximation for the driving force, \beg F_{\mathrm{dr}}=p_-(T_-)-p_+(T_+) + \lan {s}\ran(T_+-T_-). \label{fdr} \en In many cases, the free energy has even powers of the temperature, whereas the last term in Eq. (\ref{fdr}) introduces odd powers. A more convenient approximation for such cases can be obtained by noticing that $(\partial \mathcal{F}/\partial T)dT=(\partial \mathcal{F}/\partial T^2)dT^2$. Then, instead of using a linear approximation for $\partial \mathcal{F}/\partial T$, we may use a linear approximation for $\partial \mathcal{F}/\partial T^2$. We obtain \beg F_{\mathrm{dr}}=p_-(T_-)-p_+(T_+) +\left\langle\fr{dp}{dT^2}\right\rangle\left(T_+^2-T_-^2\right). \label{fdr2} \en The approximations (\ref{fdr}) and (\ref{fdr2}) are quantitatively very similar, though analytically different. The former involves the physical quantity $s$ and is useful for physical discussions. As we shall see, the latter gives cleaner analytical results. \section{Stability analysis} \label{staban} \subsection{Previous results and relevant scales} \label{scales} The stability of deflagrations in a cosmological phase transition was first studied by Link \cite{link} in the non-relativistic limit. The results are similar to those of a classical gas \cite{landau} (with the enthalpy taking the role of the mass density). According to this analysis, small wavelength perturbations of the phase transition front are stabilized by the surface tension $\sigma$, whereas large wavelength perturbations grow exponentially. The perturbations are of the form \beg e^{i\mathbf{k}\cdot \mathbf{x}^{\bot}+qz+\Omega t}, \en and the initial growth time is given by $\Omega^{-1}$. The essential features of the instability are easily seen if we consider the limit $T_\pm\simeq T_c$ in Link's result, so that the latent heat is given by $L\simeq w_+-w_-$. For small $L/w_+$ we have \beg \Omega\simeq \fr{L}{w_+}\fr{v_w}{2}\left(1-\fr{k}{k_c}\right)k, \label{omegalink} \en with \beg k_c={L}{v_w^2}/{\sigma} \en Thus, the critical wavelength above which perturbations are unstable is given by $\lambda_c=1/k_c$. Notice that, for non-relativistic velocities, Eqs. (\ref{EM3.1}-\ref{EM3.2}) give $p_+-p_-=Lv_+v_-$ which, for small $L/w_+$ becomes $p_+-p_-=Lv_w^2$. Therefore, the critical wavelength can be written as \beg \lambda_c=\frac{\sigma}{p_+(T_+)-p_-(T_-)}\equiv d_c. \label{dc} \en Physically, $d_c$ is the length scale over which the surface tension just balances the difference of the pressures on each side of the interface \cite{landau,link}. This result was improved by Huet et al. \cite{hkllm}. In the first place, relativistic velocities were considered in order to study fast deflagrations ($v_w\to c_s$). Most important, the dependence of the wall velocity on temperature was taken into account. A very simple form for the velocity was considered. \beg \gamma_w v_w(T_+)=\frac{p_-(T_+)-p_+(T_+)}{\eta} \approx \frac{L(1-T_+/T_c)}{\eta}. \label{vwsimple} \en where $\eta$ is a friction parameter, which can be assumed to be a constant. The last approximation in Eq. (\ref{vwsimple}) is valid in the non-relativistic limit and for $T_+$ close to $T_c$. The inclusion of Eq. (\ref{vwsimple}) into consideration takes into account that temperature fluctuations induce a change in the velocity of the interface. Instead of this, the treatment of Landau \cite{landau} just equated the normal velocity fluctuations $\delta v_\pm$ of the fluid to the velocity of the surface perturbation $\partial_0 \zeta$, while Link \cite{link} considered the energy flux $F=w v\gamma ^2$ to be proportional to the net blackbody energy flux, $F\propto g_-(\pi^2/30) (T_+^4-T_-^4)$, where $g_-$ is the effective number of degrees of freedom in the $-$ phase. Notice that these conditions lack important information, namely, that the interface is a phase transition front (and not, e.g., a burning front). This information is not present in either of the equations considered in \cite{landau,link}. In contrast, in Eq. (\ref{vwsimple}), the velocity of the interface depends on the pressure difference and vanishes at $T_+=T_c$, which is the essential feature of a phase transition. From Eq. (\ref{vwsimple}), an important parameter arises in the perturbation equations, namely, \beg \beta= T_c\left(-\frac{dv_w}{dT_+}\right)\gamma_w^2 v_w. \label{betasimple} \en The dependence of the wall velocity on temperature tends to stabilize the perturbations. In the non-relativistic limit and for $\Omega\ll k$, the result of Ref. \cite{hkllm} takes the form \beg {\Omega}\simeq \frac{\baL v_w}{2}\frac{\lambda-\lambda_c}{\lambda^2}\frac{(1-\beta)}{1+\beta \baL+d_b k}, \label{omegaellosnr} \en with \beg \lambda_c=d_c\left(1+\frac{2d_b/d_c}{\baL(1-\beta)}\right), \label{lambdacnrellos} \en where $\baL=L/w_+$, and a new length scale \beg d_b=\frac{\sigma}{p_-(T_+)-p_+(T_+)} \label{db} \en has appeared due to the introduction of Eq. (\ref{vwsimple}). Its physical significance is similar to that of $d_c$ [cf. Eq. (\ref{dc})]. It gives the length scale over which the surface tension just balances the pressure difference \emph{in equilibrium at $T=T_+$}. Thus, it gives the size scale of a critical bubble before it begins to grow. As remarked in Ref. \cite{hkllm}, the quantity $d_b$ gives also a time scale characteristic of the dynamics of growth, corresponding to the acceleration period before the bubble wall reaches a terminal velocity. Within the present approximations, the two scales $d_b$ and $d_c$ are related by \beg \frac{d_b}{d_c}=\frac{v_w^2}{1-T_+/T_c}, \label{dbdc} \en and Eq. (\ref{betasimple}) gives $\beta =v_w L/\eta$. We can eliminate the friction parameter and write \cite{hkllm} \beg \beta={v_w^2}/{v_c^2}, \en where \beg v_c^2\equiv(T_c-T_+)/T_c. \label{vc} \en The length scales are thus related by $d_b/d_c=\beta$. A few comments on these scales are worth. Notice that, in the equilibrium situation with $T_-=T_+$, we have $p_-(T_+)>p_+(T_+)$, whereas, for a real deflagration with $T_+>T_-$ (see Fig. \ref{figdefla}), the pressure balance is inverted, $p_-(T_-)<p_+(T_+)$ (this is why it is necessary to invert the order of $p_-$ and $p_+$ between the definitions of $d_c$ and $d_b$). It is evident that the pressure difference $p_-(T_-)-p_+(T_+)$, being negative, cannot be used as the driving force, and the equilibrium value $p_-(T_+)-p_+(T_+)$ in Eq. (\ref{vwsimple}) is a better approximation (although it does not take hydrodynamics into account). A still better approximation to the driving force is given by Eq. (\ref{fdr}). In practice, the relevant difference between the results of Refs. \cite{link} and \cite{hkllm} is the appearance of the quantity $(1-\beta)$ in Eqs. (\ref{omegaellosnr}-\ref{lambdacnrellos}) with respect to Eqs. (\ref{omegalink}-\ref{dc}). For $\beta\ll 1$, the results are essentially the same. However, for $\beta\approx 1$ the situation changes drastically. For $\beta<1$, perturbations on wavelengths $\lambda<\lambda_c$ are stable. Thus, as $\beta$ approaches 1, only very large wavelengths will be unstable, and with very large growth time $\Omega^{-1}$. For $\beta>1$, the behavior is inverted. Perturbations with $\lambda>\lambda_c$ are now {stable}. Besides, the second term in Eq. (\ref{lambdacnrellos}) dominates, and we have $\lambda_c<0$. This means that \emph{perturbations at all scales are stable for $\beta>1$}. Hence, $v_c$ is a critical velocity, above which the deflagration becomes stable. Notice that the second term in Eq. (\ref{lambdacnrellos}) is not a small correction, even for $\beta\ll 1$. In fact, due to the smallness of $\baL$, this term will generally dominate. As a consequence, the critical wavelength predicted by Huet et al. is generally quite larger than Link's result. A numerical study of the stability of planar walls was performed in Ref. \cite{fa03} for the QCD phase transition, for a value $\baL\approx 0.089$. For the case of deflagrations, two cases were considered, corresponding to $v_w=0.196$ (case A) and $v_w=0.048$ (case B). For each of these two runs of the numerical simulation, the planar wall was initialized with sinusoidal perturbations of wavelengths up to $\lambda=5\lambda_c^{\mathrm{Link}}$, where $\lambda_c^{\mathrm{Link}}$ is the critical wavelength obtained by Link and given approximately by Eq. (\ref{dc}). According to Link's results, these perturbations should be unstable. However, the perturbations decayed in the simulation, in agreement with the results of Huet et al. Indeed, the stability parameter is $\beta=0.516$ for case A and $\beta=0.722$ for case B \cite{fa90}. Hence, according to Eq. (\ref{lambdacnrellos}), we have $\lambda_c\approx 25\lambda_c^{\mathrm{Link}}$ and $\lambda_c\approx 60\lambda_c^{\mathrm{Link}}$, respectively. Perturbations of wavelengths $\lambda$ higher than these values of $\lambda_c$ should be unstable according to the results of Huet et al. Unfortunately, such perturbations were not considered in the simulations of Ref. \cite{fa03}. Although the treatment of Ref. \cite{hkllm} improved significantly upon previous stability analysis, some of the approximations used in this approach will not hold in the general situation. The most important are the following. In the first place, we remark that the simple expression (\ref{vwsimple}) for $v_w$ implicitly assumes the relations $T_-=T_+$ and $v_+=v_-=v_w$ [cf. Eqs. (\ref{solfric}),(\ref{fdr})], while for a deflagration we have $T_-<T_+$ and $ v_+<v_-=v_w$. Moreover, since the expression (\ref{vwsimple}) depends only on $T_+$ and $v_-$, it does not allow to perform \emph{independent perturbations} of variables on each side of the wall (such as $\delta T_-,\delta v_+$). These limitations of the surface equation constrain the validity of the treatment of Huet et al. This contrasts with their treatment of the fluid equations, where independent perturbations were considered for $-$ and $+$ variables. One expects that the approximation $\delta T_-=\delta T_+$ will be valid if $T_-\simeq T_+$. However, the latter is not the most general case. In the second place, the reheating in front of the wall was not taken into account. The value of $T_+$ was estimated from results on the amount of supercooling\footnote{The temperature at which bubbles nucleate and expand can be estimated using the bubble nucleation rate, which is calculated using the thermal instanton technique \cite{linde}.} for the electroweak phase transition \cite{dlhll} (i.e., the approximation $T_+\simeq T_N$ was used). Notice that, for a deflagration, the fluid is reheated in front of the wall (see Fig. \ref{figdefla}). As a consequence, the temperature $T_+$ in front of the phase transition front does not coincide with the nucleation temperature $T_N$. Depending on the wall velocity and the amount of latent heat released, the local reheating can be important. Our derivation of the perturbation equations will be similar to that of Ref. \cite{hkllm}. The main difference will be, essentially, that instead of considering Eq. (\ref{vwsimple}), we shall consider Eq. (\ref{solfric}), which depends explicitly on the two velocities $v_\pm$ and on the two temperatures $T_\pm$ (through the driving force). According to the approximation (\ref{fdr}), we have $F_{\mathrm{dr}}=p_-(T_-)-p_+(T_+) +\langle s\rangle (T_+-T_-)$. In the case of small supercooling (i.e., $T_c-T_{\pm}\ll T_c$) and small wall velocity, the pressure difference is $\mathcal{O}(Lv_w^2)$ and can be neglected in comparison with the term $\langle s\rangle (T_+-T_-)$. Besides, we can use the approximation \cite{ms09} \beg \label{tmatmenr} T_+-T_-=\fr{\Delta s(T_c)}{s_-(T_c)}(T_c-T_+) \en to obtain \beg F_{\mathrm{dr}}=({\langle w\rangle}/{w_-})\,{L(1-T_+/T_c)}, \label{fdrsmallsup} \en which, taking into account that $\langle v\rangle=-v_w\langle w\rangle/w_+$, gives \beg v_w=\frac{w_+}{w_-}\frac{L(1-T_+/T_c)}{\eta}.\label{vwsmallsup} \en This is similar to Eq. (\ref{vwsimple}), except for the factor $w_+/w_-$, which is $\approx 1$ for small latent heat. Thus, in the limit of small supercooling, small $v_w$, and small $L/w_+$, we obtain the approximation used in Ref. \cite{hkllm} for the wall velocity. In our treatment, the parameter $d_b$ will be replaced by \begin{equation} d=\frac{\sigma}{F_{\mathrm{dr}}} \label{d} \end{equation} For small supercooling and small latent heat, we have $d_b\approx d$. According to Eq. (\ref{dbdc}), the parameter $d_c$ is determined by $d_b$ and $v_c$. Similarly, our results will depend on $d$ and the critical velocity. \subsection{Linearized equations} We shall consider small perturbations of the fluid and the interface. The planar symmetry allows us to consider a single transverse direction $x^{\perp}$ instead of two directions $x,y$. The perturbation variables we shall use are the pressure fluctuation $\de p(x^{\pe},z,t)$, the variation of the velocity along the wall motion $\de v (x^{\pe},z,t)$, the transverse velocity $v^\pe(x^{\pe},z,t)$, and the variation of the wall position $\z(x^\pe,t)$. For the sake of clarity, in this subsection we shall denote the stationary solutions with a bar. Thus, we have \bega\label{defs.1} p&=&\bar{p}+\de p, \\ u^\mu&=&\bar{u}^\mu+\de u^\mu=(\bar{\ga},0,0,\bar{\ga}\bar{v})+ (\bar{\ga}^3\bar{v}\de v,\bar{\ga}v^\pe,\bar{\ga}\de v), \label{defs.2} \\ z_w&=&\bar{z}_w +\z. \label{defs.3} \ena To derive the equations for these four variables, we shall consider again Eqs. (\ref{EM1.1}-\ref{EM1.2}) for a field configuration corresponding to the perturbed wall. \subsubsection{Fluid equations} Away from the (perturbed) interface, the field is a constant as before, and Eq. (\ref{EM1.1}) gives the local conservation of energy and momentum $\pa_{\mu}T^{\mu\nu}=0$, where $T^{\mu\nu}$ is given by Eq. (\ref{Tmunu}). To obtain the fluid equations it is convenient to take the projections of the 4-divergence of the stress tensor along the directions of the fluid 4-velocity and orthogonal to it, $ u_\mu {T^{\mu\nu}}_{,\nu}=0\label{HD2}, \ (g_{\al\mu}-u_\al u_\mu){T^{\mu\nu}}_{,\nu}=0\label{HD3}. $ Taking into account the relation between enthalpy and pressure variations, Eq. (\ref{pwcs}), we obtain \bega c_s^2 w u^\nu_{,\nu}+u^\nu p_{,\nu}=0,\label{HD4}\\ w u_{\al,\nu}u^\nu-p_{,\al}+u_\al u^\nu p_{,\nu}=0. \label{HD5} \ena We are interested in the stability of the deflagration solution depicted in Fig. \ref{figdefla}, for which the fluid profile is constant on both sides of the wall. Therefore, we must consider perturbations from a constant solution (we shall not consider here perturbations of the shock front). To linear order in the perturbations we obtain, for the direction along the fluid velocity, \beg\label{HD6} c_s^2 \bar{w}(\bar{\ga}^2\bar{v}\de v{,_0}+\bar{\ga}^2\de v_{,z}+v^\pe_{,\pe})+\de p_{,0}+\bar{v}\de p_{,z}=0 \en and, for the orthogonal directions, \bega \bar{w}\bar{\ga} ^2(\de v_{,0}+\bar{v}\de v_{,z})+\bar{v}\de p_{,0}+\de p_{,z}=0, \label{HD7} \\ \bar{w} \bar{\ga}^2(v^\pe_{,0}+\bar{v}v^\pe_{,z})+\de p_{,\pe}=0. \label{HD8} \ena We have not specified the reference frame yet. We shall consider the frame which moves with the unperturbed wall. Therefore, $\bav$ corresponds to the incoming and outgoing flow velocities. \subsubsection{Matching conditions at the interface} To obtain matching conditions for the perturbations, we need to derive the matching conditions for the perturbed wall, and then subtract those for the stationary wall. We shall consider that the unperturbed wall is at $\bar{z}_w=0$ and the perturbed wall at $z_w=\z (x^\pe,t)$. Let us consider again Eqs. (\ref{EM1.1}), $\pa_\mu(wu^{\mu}u^{\nu}-g^{\mu\nu}p)=-\Box\phi\pa^{\nu}\phi$. Since there is a discontinuity at the bubble wall, if we integrate these equations in a small interval along the normal direction across the surface, only the normal derivatives will give a finite difference. Thus we can neglect all other derivatives. We obtain \beg \pa_n(w\ga^2 v^n)=0, \label{stat.1} \;\; \pa_n(w\ga^2 v^n v^{\perp})=0, \;\; \pa_n(w\ga^2 v^{n2}+p)=\Box \phi \ \pa_n\phi. \en For instance, for the unperturbed wall we have $n=z$, $\Box \phi=-\pa_z^2\phi$, and we re-obtain Eqs. (\ref{EM3.1}-\ref{EM3.2}), \bega \Delta ( \baw\bav\baga^2)&=&0,\label{EM3.1b}\\ \Delta (\baw\bav^2\baga^2 +\bap)&=&0, \label{EM3.2b} \ena together with the continuity of the transverse velocity, $\Delta \bav^{\perp}=0$. The velocity $\bav^{\perp}$, though, is set to zero by symmetry. If we use Eqs. (\ref{stat.1}) for the perturbed wall, we then have to express $\partial_n$ in terms of $\partial_z$, $\partial_{\perp}$ and $\partial_0$, in order to compare with the stationary equations. Alternatively, we may stay in the coordinate system of the unperturbed wall and just consider the fluid equations, \bega \pa_0 (w \ga^2 - p)+\pa_z(w\ga^2 v)+\pa_{\perp}(w\ga^2 v^{\perp})&=&-\Box \phi \ \pa_0\phi, \label{EMcomp.1}\\ \pa_0 (w \ga^2 v^{\perp})+\pa_z(w\ga^2 v v^{\perp})+\pa_{\perp}(w\ga^2 v^{\perp}+p)&=&\Box \phi \ \pa_{\perp}\phi, \label{EMcomp.2}\\ \pa_0 (w \ga^2 v)+\pa_z(w\ga^2 v^2+p)+\pa_{\perp}(w\ga^2 v v^{\perp})&=&\Box \phi \ \pa_z\phi, \label{EMcomp.3} \ena taking into account the discontinuity of the variables at the wall. We shall use this approach. We are going to integrate across the wall along the $z$ axis. Since we shall integrate in a vanishingly small interval, we may neglect any dependence on $x^\pe$ and $t$, other than the position of the discontinuity, i.e., we may assume (for this integration) step functions depending only on $z-\z (x^\pe,t)$. Hence, we have, e.g., $\pa_0 w=-\pa_0\z \pa_z w$, $\pa_{\perp} w=-\pa_{\perp}\z \pa_z w$, etc. This leaves only $z$ derivatives in the lhs of Eqs. (\ref{EMcomp.1}-\ref{EMcomp.3}). Keeping up to linear terms in the perturbations $\z$ and $v^{\perp}$, only terms proportional to $\pa_z^2\phi \ \pa_z\phi$ (which vanish after integration) remain in the rhs, except in Eq. (\ref{EMcomp.3}), where there is also a term proportional to $(\pa_z\phi)^2$. Performing the $z$ integration, the latter gives the surface tension $\sigma=\int (\pa_z\phi)^2dz$. We thus obtain \bega -\pa_0\z \Delta(w\ga^2 - p)+\Delta(w \ga^2 v)&=&0,\label{EM2.1}\\ \Delta(w\ga^2 v v^\pe)-\pa_\pe\z\Delta p &=&0,\label{EM2.2}\\ \Delta(w\ga^2 v^2 +p)&=&-\si(\pa_0^2-\pa_\pe^2)\z,\label{EM2.3} \ena where $\Delta$ applied to any function $f$ means $\Delta f=f_+-f_-$. In the last equation we have used the fact that, according to Eq. (\ref{EM2.1}), $\Delta(w \ga^2 v)=\mathcal{O}(\z)$. Now we replace $w=\baw+\delta w$, $v=\bav+\delta v$, etc., taking into account the unperturbed equations (\ref{EM3.1b}-\ref{EM3.2b}) and the relation $\delta w=(1+c_s^{-2})\delta p$. To first order in all the perturbations, we have \bega \Delta\left[\baw\baga^2(1+\bav^2)(-\pa_0\z+\baga^2\de v)+ (1+c_s^{-2})\baga^2\bav\de p\right]=0,\label{EM4.1}\\ \Delta(v^\pe+\bav\pa_\pe\z)=0,\label{EM4.2}\\ \sigma(\pa_0^2-\pa_\pe^2)\z+\Delta\left[2\baw\baga^4\bav\de v+ \left(1+(1+c_s^{-2})\baga^2\bav^2\right)\de p\right]=0. \label{EM4.3} \ena \subsubsection{Equation for the interface} Let us now consider the field equation (\ref{EM1.2}). The field varies in a region (the wall width) around the wall position $z_w=\zeta(x^{\perp},t)$. Hence, we may assume a field profile of the form\footnote{Notice that here we are considering a reference frame for which $\bar{z}_w=\dot{\bar{z}}_w=0$, and $\zeta$ is a small perturbation, so we have $\gamma_w=1+\mathcal{O}(\zeta^2)$.} $\phi(z,x^{\perp},t)=\phi[z-\z(x^{\perp},t)]$. To first order in $\z$ and $v^{\pe}$, we have $\pa_{\mu}\pa^{\mu}\phi=\phi'(\pa_{\pe}^2-\pa_0^2)\z-\phi''$ and $u^{\mu}\pa_{\mu}\phi=\gamma(-\pa_0\zeta+v)\phi'$. Multiplying Eq. (\ref{EM1.2}) by $\phi'(z-\z)$ and integrating in $z$ (as we did in Sec. \ref{stationary}), we obtain \beg\label{EM5} \si(\pa_0^2-\pa_\pe^2)\z= p_- - p_+ + \int sdT+\tilde\eta T_c\int \phi'^2 \ga(v-\pa_0\z)dz. \en This equation is similar to Eq. (\ref{ecfric}), except for the three terms depending on $\z$. The first one, $\si\pa_0^2\z$, takes into account the acceleration of a surface element of the wall which, according to Eq. (\ref{EM5}), is determined by the sum of all the forces acting on it. The second one, $-\si\pa_\pe^2\z$, gives the restoring force due to the curvature of the surface. Finally, the term $-\pa_0\z$ takes into account the fact that the friction force depends on the relative velocity $v_r$ between the fluid and the wall. We have $\gamma(v-\pa_0\z)=\ga_r v_r$. Approximating the integrals in (\ref{EM5}) as we did in Sec. \ref{stationary}, we obtain \beg\label{EM7} \si(\pa_0^2-\pa_\pe^2)\z- F_{\mathrm{dr}}-\eta \lan \ga(v-\pa_0\z)\ran=0. \en The various thermodynamical quantities (entropy, pressure, temperature), are related through the equation of state. Hence, we may consider the driving force as a function of $T_-$ and $T_+$. In the stationary case, Eq. (\ref{EM7}) gives Eq. (\ref{solfric}), \beg\label{EM8} \lan \baga\bav\ran=-{F_{\mathrm{dr}}(\bar{T}_+,\bar{T}_-)}/{\eta}, \en where $\bar{T}_-$ and $\bar{T}_+$ are related through Eqs. (\ref{EM3.1b}-\ref{EM3.2b}). For the perturbations we obtain \beg\label{EM9} \si(\pa_0^2-\pa_\pe^2)\z= \eta\lan\de(\ga v)-\baga\pa_0\z\ran +\fr{\pa F_{\mathrm{dr}}}{\pa T_+}\delta T_+ +\fr{\pa F_{\mathrm{dr}}}{\pa T_-}\delta T_-, \en The derivatives $\partial F_{\mathrm{dr}}/\partial T_\pm$ can be calculated by using either the approximation (\ref{fdr}) or the approximation (\ref{fdr2}). In terms of our perturbation variables $\delta p_\pm$, we have \beg \fr{\si}{\eta}(\pa_0^2-\pa_\pe^2)\z = \fr{1}{2}\left[\de (\gamma_+ v_+) + b_+\fr{\de p_+}{w_{+}} - \baga_+\pa_0\z \right] + \fr{1}{2}\left[\de (\gamma_- v_-) +b_-\fr{\de p_-}{w_-} - \baga_-\pa_0\z \right], \label{EM13} \en where \beg \label{betapm} \fr{b_{\pm}}{2}\equiv T_{\pm}\fr{1}{\eta}\fr{\pa F_{\mathrm{dr}}}{\pa T_{\pm}} . \en The parameters $\sigma$ and $\eta$ can be written in terms of the fluid velocity and the scale $d$ using Eqs. (\ref{EM8}) and (\ref{d}). Thus, we can write Eq. (\ref{EM13}) in a more concise form, \beg \label{EM17} \left\langle \baga\bav d (\pa_0^2-\pa_\pe^2)\z + \baga^3\de v +b\de p/w- \baga\pa_0\z\right\rangle=0, \en with \beg \label{betapm2} \fr{b_{\pm}}{2}= \lan - \baga\bav\ran \fr{T_{\pm}}{F_{\mathrm{dr}}} \fr{\pa F_{\mathrm{dr}}}{\pa T_{\pm}} =2\lan - \baga\bav\ran \fr{T^2_{\pm}}{F_{\mathrm{dr}}} \fr{\pa F_{\mathrm{dr}}}{\pa T^2_{\pm}} . \en The last equality is useful if $F_{\mathrm{dr}}$ is quadratic in temperature. To understand the meaning of these equations, it is useful to consider some simplifications used in previous works. In Ref. \cite{hkllm} hydrodynamics was neglected in the equation for the wall. This means that, in the first place, it was assumed that $\baT_-=\baT_+$ and $\bav_+=\bav_-=-v_w$. In this case we have $F_{\mathrm{dr}}=p_-(\baT_+)-p_+(\baT_+)$ and the scale $d$ becomes equal to $d_b$ defined in (\ref{db}). In the second place, only linearly-dependent perturbations were considered for the interface equation, $\de T_-=\de T_+, \de v_-=\de v_+$, such that even the perturbed force is of the form $F_{\mathrm{dr}}(T_+)=p_-(T_+)-p_+(T_+)$. This is the most sensitive simplification used in Ref. \cite{hkllm}, since perturbations in each phase may be quite different\footnote{This is more apparent in the case of detonations, for which perturbations can grow \emph{only} in the $-$ phase, and leads the authors of Ref. \cite{hkllm} to a wrong conclusion about the stability of detonations \cite{abney, stabdeto}.}. With these approximations, the two terms in the rhs of (\ref{EM13}) are almost identical. The sum of the two coefficients $b_{\pm}/2$ contains a \emph{total} derivative $dF_{\mathrm{dr}}/dT_+$ and gives \beg \lan b\ran= T_{+}\fr{d (\gamma_w v_w)}{d T_{+}} . \en We thus obtain \beg \label{ecbetaellos} v_w d_b(\pa_0^2-\pa_\pe^2)\z = \left(1-\beta\right)\gamma_+^2\de v_+ - \pa_0\z, \en where \beg \label{betaellos} \beta=- T_{+}\fr{\pa v_w}{\pa T_{+}}\fr{1}{w_+}\fr{\de p_+}{\de v_+}. \en In Eq. (\ref{betaellos}), the variation $\de p/\de v$ depends on the solution of the fluid equations (\ref{HD6}-\ref{HD8}) on the $+$ side of the wall. Equation (\ref{ecbetaellos}) is essentially\footnote{There is a discrepancy, namely, the relative factor of $\gamma_+^2$ between the terms $\de v_+$ and $\pa_0\z$. The origin of this is that, in their derivation, the authors of \cite{hkllm} considered, for $v_+\approx -v_w$, the relation $\de(v_++v_w)=\pa_0\z$. However, the correct relativistic velocity sum gives $\de(\gamma_+^2(v_++v_w))=\pa_0\z.$} the same as Eq (52) of Ref. \cite{hkllm}. As we have seen at the beginning of this section, the coefficient $\beta$ plays a relevant role in the hydrodynamic stability of the deflagration. In the realistic case, we see that $\beta$ will split into two parts, $\beta_\pm$, corresponding to perturbations on each side of the wall. We remark that, in the case of small supercooling, Eq. (\ref{fdrsmallsup}) (which depends only on $T_+$) is a good approximation for the stationary force $F_{\mathrm{dr}}(\bar{T}_+,\bar{T}_-)$, but not for the general form $F_{\mathrm{dr}}(T_+,T_-)$, and should not be used to obtain $b_{\pm}$. \subsection{Fourier modes of the perturbations} The fluid equations away from the wall, Eqs. (\ref{HD6}-\ref{HD8}), can be expressed in matrix form, \beg\label{HD9} \left[ \hC_0\pa_0+\hC_z\pa_z+\hC_\pe\pa_\pe \right]{\vec{U}}=0 \en where (from now on we remove the bars on unperturbed variables) \beg\label{HD10} \hC_0\equiv\left[\begin{array}{ccc} 1 & c_s^2w\ga^2 v & 0\\ v& w \ga^2 & 0\\ 0 & 0 &w\ga^2 \end{array}\right],\; \hC_z\equiv\left[\begin{array}{ccc} v & c_s^2w\ga^2 & 0\\ 1 & w\ga^2v & 0\\ 0 & 0 &w\ga^2v \end{array}\right],\; \hC_\pe\equiv\left[\begin{array}{ccc} 0 & 0 & c_s^2w\\ 0 & 0 & 0\\ 1 & 0 & 0 \end{array}\right], \en and ${\vec{U}}$ is the perturbation vector \beg {\vec{U}}\equiv\left[\begin{array}{c} \de p\\ \de v\\ v^\pe \end{array}\right]. \en To solve Eq. (\ref{HD9}) we may use the separation of variables method which, in this simple case, amounts to searching for solutions of the form \beg\label{Fou1} {\vec{U}}(t,z,x^\pe)=\vec{L}e^{\Omega t+qz+ikx^{\pe}}, \en where $\Omega, q$ and $ik$ are the eigenvalues of the operators $\pa_0,\pa_{z}$ and $\pa_{\perp}$, respectively. While $k$ is a real wavenumber, corresponding to Fourier modes along the wall, $\Omega$ and $q$ are in general complex numbers. The stationary solution will be unstable whenever $\mathrm{Re} (\Omega)>0$. Inserting the modes (\ref{Fou1}) into Eqs. (\ref{HD9}) we obtain the system of homogeneous equations \beg\label{Fou3} (\hC_0\Omega+\hC_z q+\hC_\pe ik)\vec{L}\equiv C\vec{L}=0. \en The determinant of the matrix $C$ must vanish so that the trivial solution is not the only one. This gives the dispersion relations \beg qv+\Omega=0 \en or \beg \label{Fou4} (q v+\Omega)^2-c_s^2(\Omega v+q)^2+c_s^2\ga^{-2}k^2=0, \en and we have three solutions, \bega {q}_1&=&-{\Omega}/{v} ,\label{Fou10}\\ q_{2,3}&=&\fr{ (1-c_s^2)v\Omega \pm c_s(1-v^2)\sqrt{\Omega^2+ ({c_s^2-v^2})\ga^2 k^2}}{c_s^2-v^2}. \label{Fou11} \ena The corresponding eigenvectors are \beg\label{Fou7} \vec{L}_1=\left[\begin{array}{c} 0\\ 1\\ \fr{iq_1}{k} \end{array} \right],\; \vec{L}_{2,3}=\left[\begin{array}{c} -w\ga^2\left(\fr{\Omega+q_{2,3}v}{\Omega v+q_{2,3}} \right)\\ 1\\ \fr{ik}{\Omega v+q_{2,3}} \end{array} \right] \en The eigenvector $\vec{L}_1$ corresponds to a special solution which describes isobaric perturbations ($\de p_1=0$) moving with the fluid (i.e., with $z,t$ dependence of the form $z-v t$). The general solution is a superposition of these modes. In particular, for given $k$ and $\Omega$, we must consider a perturbation vector of the form \beg {\vec{U}}(t,z,x^\pe)=\vec{A}(z)e^{(\Omega t+ikx^\pe)}, \en with \beg\label{Fou2} \vec{A}(z)=\sum_{j=1}^3 A_j \vec{L}_je^{q_j z}. \en The function $\vec{A}(z)$ must satisfy the boundary conditions at $z=\pm \infty$ and the junction conditions at the wall. Accordingly, the perturbation from the planar shape of the surface will be of the form \beg\label{Fou12} \z(t,x^\pe)=D e^{(\Omega t+ikx^\pe )}. \en For $\mathrm{Re} (\Omega)>0$ the perturbation grows exponentially with time, whereas for $\mathrm{Re} (\Omega)<0$ the perturbation decays exponentially. On the other hand, the condition that the source is the wall itself, and not something outside it, implies that the perturbations must decay away from the wall \cite{landau}. Therefore, instability of the front also requires $\mathrm{Re} (q)<0$ for $z>0$ ($+$ phase) and $\mathrm{Re} (q)>0$ for $z<0$ ($-$ phase). Since $v$ is negative, the special solution gives $\mathrm{Re} (q_1)>0$ for $\mathrm{Re}(\Omega)>0$. Hence, unstable perturbations will be associated to the presence of this mode in the $-$ phase (behind the wall). Conversely, for stable perturbations this mode will be in the $+$ phase. Regarding the other two solutions, we can write Eq. (\ref{Fou11}) in the form $(\Omega-a_+q)(\Omega-a_-q)=-K^2$, with \beg a_\pm=\pm\fr{c_{s}\mp v}{1\mp c_s v} \en and $K^2=c_s^2\ga^{-2}k^2/(1-c_s^2 v^2)$, which shows that, for real $q$ and $\Omega$, we have a hyperbola with asymptotes $q=\Omega/a_\pm$. For complex $q$ and $\Omega$, it can be shown that the points $\mathrm{Re}(q),\mathrm{Re}(\Omega)$ lie in the same region between these asymptotes. We show some examples in Fig. \ref{fighip}. We have three different cases, depending on the value of $v$. For $v$ supersonic (left panel), $\mathrm{Re}(q_2)$ and $\mathrm{Re}(q_3)$ have the same sign for a given $\Omega$. For $v$ subsonic (right panel), $\mathrm{Re}(q_2)$ and $\mathrm{Re}(q_3)$ have opposite sign. In the case $v=-c_s$ (central panel) there is only one solution (besides the special one), namely \beg\label{Fou6} q_2=\fr{c_s k^2}{2\Omega}+\fr{(1+c_s^2)}{2c_s}\Omega. \en In front of the wall we require $\mathrm{Re}(q)<0$, and the possible modes are those in the lower quadrants of Fig. \ref{fighip}. Conversely, the possible modes behind the wall are those in the upper quadrants. \begin{figure}[bth] \centering \epsfysize=5.5cm \leavevmode \epsfbox{hiperb.eps} \caption{The real part of the dispersion relations $q_2(\Omega)$ (in black) and $q_3(\Omega)$ (in blue) for different values of the imaginary part $\mathrm{Im}(\Omega)$. Gray dashed lines indicate the asymptotes of these solutions. We also show the special solution $q_1(\Omega)$ in a dotted red line.} \label{fighip} \end{figure} Deflagrations are characterized by $|v_+|<c_{s+}$, i.e., the incoming flow is always subsonic. Thus, for the $+$ phase, the right panel of Fig. \ref{fighip} applies. Furthermore, in this phase we require $\mathrm{Re}(q)<0$ and the dispersion relation corresponds to the lower (blue) curves of this panel, i.e., $q= q_3(v_+,\Omega)\equiv q_{3+}$. In the $+$ phase, the special solution $q_1$ must be included only for the stable case $\mathrm{Re}\Omega<0$. On the other hand, in the $-$ phase we require $\mathrm{Re}(q)>0$, but the outgoing flow may, in principle, be supersonic as well as subsonic. As a consequence, the stability analysis is quite different for weak, strong, or Jouguet deflagrations. \subsection{Solution of the perturbation equations} \label{solu} \subsubsection{Supersonic deflagrations} As we have seen, a supersonic deflagration can be either a strong or a Jouguet solution. Let us consider first the case of strong deflagrations. As already discussed by Landau \cite{landau} for a classical gas, strong deflagrations are absolutely unstable, either in the case of a combustion front (\S 131) or a condensation discontinuity (\S 132), due to the fact that such discontinuities are not \emph{evolutionary} (i.e., there are more free parameters than conditions). According to numerical calculations \cite{ikkl94}, strong deflagrations are unstable also in the case of a relativistic phase transition. It is interesting to verify this result analytically. Strong deflagrations are characterized by a supersonic outgoing velocity $v_-$, corresponding to the left panel of Fig. \ref{fighip}. For $\mathrm{Re}(\Omega) <0$, we see that all the solutions give $\mathrm{Re}(q)<0$. Therefore, we have no possible modes with $\mathrm{Re}(\Omega) <0$ in the $-$ phase. On the other hand, for $\mathrm{Re}(\Omega)>0$ we have $\mathrm{Re}(q_2)>0$, $\mathrm{Re}(q_3)>0$, and $\mathrm{Re}(q_1)>0$. We thus have three unstable modes in the $-$ phase. In the $+$ phase, as we have already discussed, we have one unstable mode for $\Omega>0$, namely, $q_3$, since $\mathrm{Re}(q_3)<0$ (right panel of Fig. \ref{fighip}). Hence, according to Eq. (\ref{Fou2}), for $z>0$ we must consider a perturbation of the form \beg \vec A (z)=A_{+}\vec{L}_{3+} e^{{q}_{3+} z} , \label{modomas} \en whereas for $z<0$ we have \beg \vec{A}(z)=A_-\vec{L}_{3-}e^{{q}_{3-} z}+ B\vec{L}_{2-}e^{{q}_{2-}z}+C\vec{L}_{1-}e^{{q}_{1-}z}. \label{modosstrong} \en Here, the $\pm$ signs in the $q_i$ and $\vec L_i$ mean that these quantities [which are given by (\ref{Fou10}-\ref{Fou7})] must be evaluated at $v_\pm$, respectively. We must impose the junction conditions (\ref{EM4.1}-\ref{EM4.3}) and the surface equation (\ref{EM17}) to the fluid variables and the surface deformation $\zeta$. This gives four equations for the five variables $A_+,A_-,B,C$, and $D$. Such a system of equations has infinite solutions. Hence, the strong deflagration is trivially unstable. Notice that our treatment applies to any of the strong deflagration profiles sketched in Figs. \ref{figdefla} and \ref{figstrong}, since we perturbed the fluid near the wall from constant velocity solutions. As we decrease the velocity and reach the Jouguet point ($|v_-|=c_{s-}$), one of the asymptotes becomes vertical and the hyperbola becomes single-valued (central panel of Fig. \ref{fighip}). In this case the solution $q_3$ disappears. Thus, we now have only two unstable modes in the $-$ phase, corresponding to $q_2$ and the special solution $q_1$. Hence, we must set $A_-=0$ in Eq. (\ref{modosstrong}). In the $+$ phase the situation is the same as before (since, for deflagrations, $v_+$ is always subsonic), and the unstable mode is given by Eq. (\ref{modomas}). As a consequence, the deflagration becomes evolutionary. As we have discussed in Sec. \ref{hydrostat}, for the supersonic Jouguet deflagration the fluid velocity profile develops a tail just behind the wall (right panel of Fig. \ref{figstrong}), and our treatment no longer applies. Although it would be interesting to study the stability of this kind of solution, it is quite a difficult task since the stationary velocity profile is a function of $z$ and $t$, namely, $v_\mathrm{rar}(z/t)$. Such a study is out of the scope of the present paper and we shall attempt it elsewhere. For the traditional deflagration profile, the Jouguet point just corresponds to the case $v_w=c_s$, which is the limit between strong and weak deflagrations. \subsubsection{Weak deflagrations} In the case of weak deflagrations, both velocities $v_-$ and $v_+$ are subsonic. We have $\mathrm{Re}(q_2)>0$ and $\mathrm{Re}(q_3)<0$ (right panel of Fig. \ref{fighip}). Hence, we must consider the mode with eigenvalue $q_3$ again in the $+$ phase and the mode with eigenvalue $q_2$ in the $-$ phase. Besides, for $\mathrm{Re}(\Omega)>0$, we have $\mathrm{Re}(q_1)>0$, and the special solution must be considered in the $-$ phase. Thus, in the unstable case, for $z>0$ the function $\vec A(z)$ is again of the form (\ref{modomas}), $\vec A(z)=A\vec{L}_{3+} e^{{q}_{3+} z}$, while for $z<0$ we have $ \vec{A}(z)= B\vec{L}_{2-}e^{{q}_{2-}z}+C\vec{L}_{1-}e^{{q}_{1-}z}$. The junction conditions (\ref{EM4.1}-\ref{EM4.3}), as well as the surface equation (\ref{EM17}), require evaluating these functions at the wall position $z=\z$. However, to first order in the perturbations, we just evaluate at $z=0$. We thus have, on each side of the interface (omitting a factor $e^{\Omega t+ik x^\pe}$), \beg\label{Weak2} \de v_+= A,\;\; \de p_+=-\baga_+^2\baw_+\fr{\Omega+{q}_{3+}\bav_+}{\Omega \bav_+ +{q}_{3+}}A,\;\; v^\pe_+=\fr{ik}{\Omega\bav_++{q}_{3+}}A,\;\; \en and \beg\label{Weak4} \de v_-=B+C,\; \de p_-=-\baga_-^2\baw_-\fr{\Omega+{q}_{2-}\bav_-}{\Omega \bav_- +{q}_{2-}}B, \; v^\pe_-=\fr{ik}{\Omega\bav_- +{q}_{2-}}B+\frac{i{q}_{1-}}{k}C. \en These quantities and those related to the corrugation of the wall, \beg \pa_0\z=\Omega D, \; \pa_{\perp}\z=ikD , \; \pa_0^2\z=\Omega^2 D, \; \pa_{\perp}^2\z=-k^2D \label{corrug} \en (omitting again a factor $e^{\Omega t+ik x^\pe}$), are related by Eqs. (\ref{EM4.1}-\ref{EM4.3}) and (\ref{EM17}). We thus have four equations for the four unknowns $A,B,C$ and $D$, i.e., the weak deflagration is evolutionary. It is interesting to consider also the case of a stable perturbation, $\mathrm{Re}(\Omega)<0$. In this case we have $q_1<0$ and the special solution must now be included in the $+$ phase instead of the $-$ phase. The form of the perturbations is similar to that of Eqs. (\ref{Weak2}-\ref{Weak4}), except that the variable $C$ appears in (\ref{Weak2}) instead of (\ref{Weak4}). As we shall see, the jump of the special mode from one side of the interface to the other as $\Omega$ changes sign will cause a discontinuity in the wavenumber $k$ as a function of $\Omega$. Inserting Eqs. (\ref{Weak2}-\ref{corrug}) in Eqs. (\ref{EM4.1}-\ref{EM4.3},\ref{EM17}), we obtain a homogeneous system of linear equations for the constants $A,B,C$ and $D$. Nontrivial solutions exist if the determinant of the matrix associated to this system vanishes. After some manipulations (e.g., multiplying the first column by the factor $Q_+$ defined below, etc.), this condition can be written in the form \begin{equation} \label{matriz} \left| \begin{array}{cccc} \frac{1}{R_+} & \frac{1}{R_-} & \frac{\hat\Omega}{v_-\gamma_-^2} & -(v_+-v_-) \\ v_-(1+\frac{v_+\W}{R_+}) & -v_+(1-\frac{v_-\W}{R_-}) & v_+(1+v_-^2) & \W (1-v_-v_+)(v_+-v_-) \\ 1+\frac{\W}{v_+R_+} & -1+\frac{\W}{v_-R_-} & 2 & \frac{F_{\mathrm{dr}}}{w_+}\frac{1}{v_+\ga_+^2}(\W^2+1)kd \\ \fr{\ga_{s+}^2}{2}(\ga_+Q_+-b_+P_+) & \fr{\ga_{s-}^2}{2}(\ga_-Q_--b_-P_-) & -\fr{\ga_-}{2} & -\langle\ga\rangle\W+\langle\ga v\rangle(\W^2+1)kd \\ \end{array}% \right| =0. \end{equation} where $\W\equiv \Omega/k$, and \begin{equation} R_{\pm}=\sqrt{\frac{\gamma_\pm^2}{\gamma_{s\pm}^2}+\frac{\hat{\Omega}^2}{c_{s\pm}^2}},\;\; P_\pm=v_\pm \mp \frac{\W}{R_\pm}, \;\; Q_\pm=1\mp \frac{v_\pm}{c^2_{s\pm}}\frac{\W}{R_\pm}, \end{equation} with $\gamma_{s\pm}\equiv 1/\sqrt{1-v_\pm^2/c_{s\pm}^2}$. Although the solution is, by symmetry, symmetric in $k$, to obtain these expressions we have assumed $k>0$\footnote{In particular, we have inserted a factor of $k$ inside a square root in the expressions for $q_{2,3}/k$ to obtain the quantities $R_{\pm}$.}. Thus, from now on we are using the notation $k=|k|$. A solution of Eq. (\ref{matriz}) is $\W=-\gamma_-v_-$. Indeed, for this value of $\W$ we have $R_-=\gamma_-$, $P_-=0$, and $Q_-=\gamma_{s-}^{-2}$. Hence, the second and third columns of the matrix are proportional, and the determinant vanishes. However, as explained in Ref. \cite{hkllm}, this solution is spurious and has no physical significance, as it leads to vanishing values of the variables. Finding analytical solutions for $\Omega(k)$ from Eq. (\ref{matriz}) is a difficult task. Notice, on the other hand, that the wavenumber $k$ appears only in the fourth column, in the matrix elements ${34}$ and ${44}$. Hence, we can readily find an expression for $k$ as a function of $\W$, \beg kd=\frac{(v_+ -v_-) \left[\det_{14}+\det_{24}(1-v_+v_-)\W\right]-\det_{44}\langle \ga\rangle \W}{(1+\W^2)\left[({F_{\mathrm{dr}}}/{w_+})\det_{34}/({v_+\ga_+^2})-\langle\ga v\rangle\det_{44}\right]}, \label{kd} \en where $\det_{ij}$ is the determinant of the $3\times 3$ matrix that results by removing the $i$-th row and the $j$-th column in Eq. (\ref{matriz}). We remark that, for $\mathrm{Re}(\Omega)<0$, we will have a different matrix, since the special mode must be considered in the $+$ phase instead of the $-$ phase. This amounts to changing, in the third column of the matrix, the indices $\pm\leftrightarrow \mp$ and the sign of the first three elements. \section{Stability of weak deflagrations} \label{stab} We shall now attempt to find all the unstable solutions for a given wavenumber. Notice that, for real $\Omega$, Eq. (\ref{kd}) facilitates to study the general properties of the solution. Moreover, one may obtain a plot of $\Omega$ vs $k$ by just inverting the graph of $kd(\W)$. However, it is not clear form Eq. (\ref{kd}) whether solutions with $\mathrm{Im}(\Omega)\neq 0$ are possible as well. \subsection{Small velocity limit} In order to understand the general behavior of the function $\Omega(k)$, it is convenient to consider first non-relativistic velocities, so that we can write down analytical expressions which are rather lengthy in the general case. We shall also use, in this subsection, the approximation of small supercooling, which is consistent with a small wall velocity and avoids considering a particular EOS. Thus, from Eq. (\ref{betapm2}) and the general driving force (\ref{fdr}) we obtain $b_\pm\simeq \langle v\rangle L/F_{\mathrm{dr}}$. Then we may use, for the stationary driving force, the approximation (\ref{fdrsmallsup}), \beg \label{fdrvcnr} F_{\mathrm{dr}}=\frac{\langle v \rangle}{v_+}Lv_c^2, \en where $v_c=\sqrt{1-T_+/T_c}$. We thus obtain \beg b_\pm ={v_+}/{v_c^2}. \label{bnr} \en Besides, with these approximations Eq. (\ref{EM3.1}) gives $v_--v_+= v_- L/w_+$ plus higher order in $(T_c-T)/T_c$. Hence, Eq. (\ref{matriz}) becomes \begin{equation} \label{matriznr} \left| \begin{array}{cccc} \frac{1}{R_+} & \frac{1}{R_-} & \frac{\hat\Omega}{v_-} & v_- \baL \\ v_-(1+\frac{v_+\W}{R_+}) & -v_+(1-\frac{v_-\W}{R_-}) & v_+ & -\W v_- \baL \\ 1+\frac{\W}{v_+R_+} & -1+\frac{\W}{v_-R_-} & 2 & (\W^2+1)kd\langle v \rangle\frac{v_c^2}{v_+^2}\baL \\ \fr{1}{2}(1-\beta_+) & \fr{1}{2}(1-\beta_-) & -\fr{1}{2} & (\W^2+1)kd\langle v\rangle-\W \\ \end{array}% \right| =0, \end{equation} where $\baL={L}/{w_+} $, and \beg \label{Rnr} R_{\pm}=\sqrt{1+\frac{\hat{\Omega}^2}{c_{s\pm}^2}},\; \beta_+=\frac{v_+(v_+-\frac{\W}{R_+})}{v_c^2},\; \beta_-=\frac{v_+(v_-+\frac{\W}{R_-})}{v_c^2}. \en \subsubsection{Case $|\W|\gg |v_\pm|$} Let us first look for solutions of large $|\W|$. We are interested in the instability case $\mathrm{Re} (\W)>0$, for which we have $R_\pm=\W/c_{s\pm}(1+\mathcal{O}(c_s^2/\W^2))$. Hence, to lowest order, the parameters appearing in Eq. (\ref{matriznr}) become \beg \label{Rapprox} \frac{1}{R_\pm}=\frac{c_{s\pm}}{\W}, \; \frac{\W}{R_\pm}=c_{s\pm},\; \beta_+=\frac{v_+(-c_{s+}+v_+)}{v_c^2},\; \beta_-=\frac{v_+(c_{s-}+v_-)}{v_c^2}. \en Notice that, in this limit, we have $\beta_+>0,\beta_-<0$. We can now easily calculate the $3\times 3$ determinants defined above. The factors of $\W$ cancel out in $\det_{14}$, whereas $\det_{24},\det_{34}$, and $\det_{44}$ are of the form $\det_{i4}=(\W/v_-)\det'_{i4}$, where the $\det'_{i4}$ are determinants of $\W$-independent $2\times 2$ matrices. Thus, we have \beg \label{omegagrande} \Omega d= \fr{1}{\langle v\rangle }\frac{\det'_{44}+\baL v_-\det'_{24}}{\det'_{44}- \baL\frac{v_c^2}{v_+^2}\det'_{34}}, \en From this expression we see that (in the limit of large $|\Omega/k|$) $\Omega$ is be real. More importantly, we can see that the rhs of Eq. (\ref{omegagrande}) is negative, which means that, in fact, we cannot have $|\W|$ large for $\Omega>0$\footnote{Notice that we are considering the case $\mathrm{Re} (\W)>0$; for $\mathrm{Re} (\W)<0$, the matrix in Eq. (\ref{matriznr}), as well as the approximations (\ref{Rapprox}), would have different forms.}. Indeed, consider for simplicity $c_{s+}=c_{s-}$. Using the relation $v_-=(1-\baL)v_+$ and dropping terms $\mathcal{O}(v^2/c_s^2)$, we obtain \begin{eqnarray} \det{}'_{44}&=&2c_s+\baL(1+c_s^2)|v_-| ,\\ v_-\det{}'_{24} &=& \frac{c_s}{2}\left[\frac{\baL}{1-\baL} +\frac{(2-\baL)|v_-|(c_s-\baL|v_-|)}{v_c^2}\right] ,\\ - \frac{v_c^2}{v_+^2}\det{}'_{34} &=& \frac{v_c^2}{v_+^2}\frac{2-\baL}{2}|v_-| +\frac{\baL(c-\baL|v_-|)}{2(1-\baL)} -|v_-|(1+c_s^2), \end{eqnarray} where we have used absolute values to make the signs clearer. We see immediately that $\det'_{44}$ and $v_-\det'_{24}$ are positive (since $\baL<1$). On the other hand, in the expression for $- ({v_c^2}/{v_+^2})\det{}'_{34} $, only the last term is negative. However, in Eq. (\ref{omegagrande}) this term cancels with the last term of $\det'_{44}$. Hence, since $\langle v\rangle<0$, the rhs of Eq. (\ref{omegagrande}) is negative. We conclude that there are no unstable modes with $|\W|\gg v_w$. \subsubsection{Case $0<|\W|\lesssim|v_\pm|$} Let us consider now the case of smaller $\W $, up to order $v_\pm$. To linear order in $v_+$, $v_-$, and $\hat{\Omega}$, we have $R_\pm=1$, and \beg \label{betanr} \beta_+=\frac{v_+}{v_c^2}(v_+-\W), \ \beta_-=\frac{v_+}{v_c^2}(v_-+\W). \en In order to compare with the results of Ref. \cite{hkllm}, we write down again the determinant in this limit, \beg \left|% \begin{array}{cccc} 1 & 1 & \fr{\W}{v_-} & v_- {\baL} \\ v_- & -v_+ & v_+ & -v_-\W {\baL} \\ 1+\fr{\W}{v_+} & -1+\fr{\W}{v_-} & 2 & kd \frac{v_c^2}{v_+^2}\langle v\rangle {\baL}\\ \fr{1}{2}(1-\be_+) & \fr{1}{2}(1-\be_-) & -\fr{1}{2} & kd \langle v\rangle -\W\\ \end{array}% \right|=0. \label{matrnr} \en As expected, the first three rows of the matrix (corresponding to the junction equations for the fluid perturbations) match\footnote{Taking into account that $d\,v_c^2/v_+^2=d_c$ and the notations $v_q=-v_+$, $v_h=-v_-$, $\baL\to\baL/2$.} those of Ref. \cite{hkllm}. All the differences appear in the forth row (which corresponds to the equation for the interface). Since we considered independent perturbations of variables in each phase, this row is more symmetric in our case. In the case of dependent perturbations, we would replace $\delta v_-\to\delta v_+$, $\delta p_-\to\delta p_+$ in Eq. (\ref{EM13}). We would thus obtain zeros in the elements ${42}$ and ${43}$ (corresponding to the perturbations $\de v_-$ and $\de p_-$) and a factor of $2$ in the element $41$, which would be just given by $1-\beta$ ($1-\eta$ in the notation of \cite{hkllm}). As we shall see next, for small velocities, the aforementioned differences do not introduce a significant qualitative variation with respect to the results of Ref. \cite{hkllm}. Roughly, the role of that single $\beta$ will be played by the average of $\be_+$ and $\be_-$. More important discrepancies will appear for higher velocities. The results will differ significantly also in the case $\Omega\leq 0$, even in the non-relativistic case. Indeed, since the special mode must be considered on either side of the wall according to the sign of $\Omega$, we cannot use Eq. (\ref{matrnr}) for $\Omega<0$. As a consequence, we shall find a discontinuity in the passage from $\Omega> 0$ to $\Omega<0$. Let us first consider the case $\mathrm{Re}(\Omega)>0$. Besides the aforementioned spurious solution $\W=-v_-$, we obtain a solution of the form \beg kd=N/D. \en The expressions for $N$ and $D$ are still rather lengthy, and we shall only write down the case of small $L/w_+$. To first order we have \begin{equation} \begin{split} kd & \left[ 1+\fr{\baL(1+\fr{\bar{L}}{2})}{2\beta}+\fr{\W}{v_w} \right] \ = \ \fr{\baL}{2}\left(1- \beta\right) \label{kdnrcubic} \\ & - \left[1+\fr{\baL (1-\baL)\beta}{2}\right]\fr{\W}{v_w} - \left[1+\fr{\baL(1+\beta)}{2}\right]\left[\fr{\W}{v_w}\right]^2 - \fr{\baL\beta(1+\baL)}{2}\left[\fr{\W}{v_w}\right]^3, \end{split} \end{equation} where we have defined the parameter \beg \beta=(1-\baL)\; \fr{v_w^2}{v_c^2}. \label{betanrfinal} \en In Eqs. (\ref{betanr}-\ref{betanrfinal}) we have neglected terms of order $v_w^2$, except in the ratio $v_w^2/v_c^2$, since $v_c$ may be small. Notice that $v_c^2=1-T_+/T_c$ gives a measure of the amount of supercooling. Hence, in the non-relativistic approximation, if we are interested in velocities $v_w\sim v_c$, the amount of supercooling must be small enough (of order $v_w^2$), i.e., $(T_c-T_+)/T_c=v_c^2\sim v_w^2$. In order to obtain all the possible values of $\Omega$ (with positive real part), we should invert the relation (\ref{kdnrcubic}), which amounts to finding the three roots of a cubic polynomial. Notice that the coefficient of the cubic term is in general suppressed with respect to the the linear and quadratic ones. If we neglect this term, we obtain a quadratic equation, with only two roots. The third root of the cubic equation must be large enough to make the cubic term comparable to the quadratic one, i.e., $\W/v_w\sim \baL^{-1}$. This root is in general well beyond the range of validity of the approximation $\W\lesssim v_w$ and must be discarded (in any case, it can be seen that this solution has a negative real part). Let us just assume, for simplicity, that $\baL$ is small enough that we can neglect the cubic term. Then, we have a quadratic equation with real coefficients, and it is trivial to see the structure of the solutions. We have either two complex conjugate roots or two real roots. Since the coefficients of the quadratic and linear terms have the same sign, in the case of complex roots the real part is negative. In the case of real roots, one of them is negative. The other has a smaller absolute value and may be positive or negative, depending on the sign of the $\W$-independent term. Thus, we have, at most, only one solution with $\mathrm{Re}(\Omega)>0$. This solution has $\mathrm{Im}(\Omega)=0$. To study this solution, it is convenient to analyze the behavior of $k$ as a function of $\Omega$. We can thus go back to Eq. (\ref{kdnrcubic}) and consider $\Omega$ real. For $\W>0$ the lhs of Eq (\ref{kdnrcubic}) is positive, while most of the terms in the rhs are negative. Indeed, the coefficients of the $\W$-dependent terms are negative. The $\W$-independent term is negative for $\beta>1$. In such a case, there is no solution with $\W> 0$ (i.e., we would have unphysical values $k<0$), and the deflagration is stable under perturbations of any wavelength. For $\beta<1$, the $\W$-independent term in the rhs is positive and we have unstable solutions for $\W$ below a certain value $\W_{0}$. At $\W=\W_0$ we have $k=0$ and, as $\W$ decreases from $\W_{0}$, the value of $k$ increases. For vanishing $\W$ we obtain the maximum wavenumber $k_c$ for which the perturbation is exponentially unstable. \subsubsection{Analytic approximations} \label{analyt} In the limit of very small $\baL$, the critical value $\beta =1$ is attained for $v_w=v_c$. This is in agreement with Ref. \cite{hkllm} (in this limit, and for $\W\to 0$, we have $\beta_+=\beta_-=\beta$). If $\baL$ is not negligible, the critical velocity is somewhat higher, \beg v_{\mathrm{crit}}=v_c/\sqrt{1-\baL}\approx v_c(1+\baL/2). \label{vcL} \en For $v_w$ below $v_{\mathrm{crit}}$, we have $\Omega>0$ in the range $0<k<k_c$. The value of $k_c$ is obtained by setting $\W=0$ in Eq. (\ref{kdnrcubic}), \beg \label{kcdlinear} k_cd=\frac{\baL}{2}\frac{1-\beta}{1+(1+\fr{\baL}{2})\fr{\baL}{2\beta}}. \en This equation shows explicitly the fact that all perturbations are stable for $\beta>1$. As already mentioned, to obtain the value of $\Omega$ for a given $k$, we should in principle invert the cubic equation (\ref{kdnrcubic}). In Ref. \cite{hkllm} the corresponding equation is quadratic because the dependence of the parameter $\beta_+$ on $\Omega$ is neglected. Furthermore, it is argued that the smallness of the velocity (and the fact that $\W\sim v_w$) insures that the term quadratic in $\Omega$ is small and can also be neglected, obtaining a linear equation. Notice, however, that $\W\sim v_w$ is not a good argument to neglect any of the terms in the equation. Nevertheless, since the independent term in the rhs of Eq. (\ref{kdnrcubic}) is always smaller than $\baL/2$, we have $\W_{0}/v_w< \baL/2$. Since we always have $\baL<1$ (and often $\ll 1$), this is an important constraint. As a consequence, for the unstable range $0<\Omega<\Omega_{0}$, the value of $\W/v_w$ will be, in most cases, small enough to safely neglect the quadratic and cubic terms in Eq. (\ref{kdnrcubic}). Keeping only the linear terms we obtain \beg \label{omegalinear} \fr{\W}{v_w }=\frac{\baL}{2}\frac{(1-\beta)(1-k/k_c)}{1+(1-\baL) \fr{\baL}{2}\beta+kd}, \en which agrees with the result of Ref. \cite{hkllm} for $\baL\ll 1$. Thus, we have \beg \label{omegamaxlinear} \fr{\W_{0}}{v_w }=\frac{\baL}{2}\frac{1-\beta}{1+(1-\baL)\fr{\baL}{2}\beta} \en We see that both $\W/v_w$ and $kd$ are at most of order $\baL$. We plot the function $\W$ in Fig. \ref{figlinear}. Notice that Eq. (\ref{omegalinear}) is valid only for $\W>0$. Although we are not interested in general in the case $\mathrm{Re}(\Omega)<0$, which is exponentially stable, it is important to consider the limit $\Omega\to 0^-$. As we have seen, for $\mathrm{Re}(\Omega)<0$ the special mode must be considered in front of the wall. This results in a change in the third column of the matrix in Eq. (\ref{matrnr}), which becomes $\mathrm{col}(-\W/v_+,-v_-,-2,-1/2)$. This will give a jump in $k$ as a function of $\Omega$ (see Fig. \ref{figlinear}). Proceeding as before we obtain, for small (and negative) $\W/v_w$, \beg \label{omegalinearneg} \fr{\W}{v_w}= \frac{\baL}{2}\frac{(1-\beta+\frac{\baL}{2})(1-k/k_c')}{1- (1+\baL)\fr{\baL^2}{2}\beta-(1+2\baL)kd} \;\;\; (\Omega<0), \en where the minimum wavenumber for which the perturbation is exponentially stable is given by \beg \label{kcdlinearneg} k_c'd=\frac{\baL}{2}\frac{1-\beta+\baL}{1-(1+\fr{3\baL}{2})\fr{\baL}{2\beta}}. \en \begin{figure}[bth] \centering \epsfysize=5.5cm \leavevmode \epsfbox{linear.eps} \caption{$\W$ vs $kd$ in the small $v_w$, small $L$ and small $\W/v_w$ approximation, for $\baL=0.01$, $T_+/T_c=0.995$ ($v_c\simeq 0.07$), and $v_w=0.05$ (left), $0.007$ (center) and $0.003$ (right).} \label{figlinear} \end{figure} Comparing Eq. (\ref{omegalinearneg}) with Eq. (\ref{omegalinear}), we observe some differences of order $\baL \sim\Delta v/v$ in the expressions for $\Omega>0$ and $\Omega<0$ (due to the changes $v_\pm \leftrightarrow v_\mp$ in the third column of the matrix). On the other hand, from Eq. (\ref{kcdlinearneg}) we see that there is also an important sign difference in the denominator of $k_c'$ with respect to that of $k_c$. For $\beta\sim 1$ (i.e., $v_w\sim v_c$), $k_c'$ will be slightly higher than $k_c$. As a consequence, there will be a small gap in the plot of $\Omega$ vs $k$, as can be seen in the left panel of Fig. \ref{figlinear}. This gap grows significantly as $v_w$ decreases from $v_c$ (center panel of Fig. \ref{figlinear}), since $k_c'$ has a pole at $\beta\approx \sqrt{\baL/2}$. Hence, for a wall velocity $v_c'\approx (\baL/2)v_c$ the interval $k_c<k<k_c'$ becomes infinite. Below this velocity, $k_c'$ takes negative values (right panel of Fig. \ref{figlinear}) and the solution with $\Omega<0$ becomes unphysical. There are, in general, more solutions with $\mathrm{Re}(\Omega) <0$, for higher values of $|\Omega|$. We are not interested in them, though, since they correspond to stable perturbations. In this case, we have linear stability for $k>k_c$. In the region of $\Omega>0$ our results agree with those of Ref. \cite{hkllm}. In this region, $\W$ decreases from $\W_{0}$ to $0$ as $k$ increases from $0$ to $k_c$. Thus, the instability range of $k$ is limited by $k_c\lesssim \baL/2$, whereas $\W$ is bounded by $\W_{0}\lesssim v_w\baL/2$. These values are also proportional to $1-v_w^2/v_{\mathrm{crit}}^2$. Therefore, for higher velocities they are even smaller, and stability is recovered at $v_w=v_{\mathrm{crit}}$. According to Eqs. (\ref{kcdlinear}) and (\ref{omegamaxlinear}), for small velocity we have $k_cd\sim v_w^2$ and $\W\sim v_w$. Hence, stability is also recovered as $v_w$ vanishes. This is shown in Fig. \ref{figkcommaxnr} (solid lines). For a given velocity, the exponentially unstable wavenumbers are those below the curve of $k_c$, whereas the possible values of $\W$ lie below the curve of $\W_{0}$. Beyond the critical velocity both $k_c$ and $\W_{0}$ become negative. \begin{figure}[bth] \centering \epsfysize=4.5cm \leavevmode \epsfbox{kcommaxnr.eps} \caption{The values of $k_c d$ (solid line), $k_c'd$ (dashed line), and $\W_{0}$ as functions of $v_w$, for the same set of parameters of Fig. \ref{figlinear}.} \label{figkcommaxnr} \end{figure} For $k\geq k_c$, on the other hand, we have significant differences with the results of Ref. \cite{hkllm}. The gap between $k_c$ and $k_c'$ was not observed in that analysis. The reason may be the following. As we have seen, for $\Omega<0$ the only changes in the matrix in Eq. (\ref{matrnr}) are in the third column. For $\W=0$, the discontinuity in this column is given by the change $\mathrm{col}(0,v_+,2,-1/2)\to \mathrm{col}(0,-v_-,-2,-1/2)$. Since changing the sign of a column does not alter Eq. (\ref{matrnr}), the above change is equivalent to $\mathrm{col}(0,v_+,2,-1/2)\to\mathrm{col}(0,v_-,2,1/2)$. In the case of small $\baL$, the change $v_+\to v_-$ is not relevant. Therefore, the discontinuity is dominated by the change of sign $-1/2\to 1/2$ in the element 43. However, due to the simplified treatment of Ref. \cite{hkllm}, in that work this element is $0$ instead of $-1/2$. In principle, at $k=k_c$, $\Omega$ jumps to a value $\Omega<0$ (not shown in Fig. \ref{figlinear}, as it lies beyond the linear approximation). It is not clear, though, from the linear stability analysis, whether there are unstable perturbations or not in the range $k_c<k<k_c'$. For $k<k_c$ the perturbations grow exponentially, whereas for $k>k_c'$ they decay exponentially. In the range between $k_c$ and $k_c'$, one may expect marginal stability with $\Omega=0$. Notice that, as $\Omega\to 0$, the special solution, which is proportional to $\exp(\Omega z/|v|)$, becomes non-normalizable in either side of the wall, and cannot be included at all in Eqs. (\ref{Weak2}-\ref{Weak4}). As a consequence, we will have four equations for the three unknowns $A,B,D$, and the only solution will be the trivial one, $A=B=D=0$. Thus, the approximation of keeping to linear order in perturbations breaks down. It is out of the scope of the present paper to go beyond the linear stability analysis. In any case, in the range of wavenumbers between $k_c$ and $k_c'$, the perturbations will not grow exponentially. For our purposes, it will be enough to assume that, in this range, the possible instabilities would grow more slowly than for $k<k_c$. From now on, we shall concentrate on the case of $\Omega>0$. \subsubsection{Reheating effects} The stability of the perturbations depends on the wall velocity and on the temperature $T_+$. Notice, however, that the boundary conditions for the fluid fix the temperature beyond the shock front (the nucleation temperature $T_N$), while $T_+$ depends on the amount of reheating (see Fig. \ref{figdefla}). For a given nucleation temperature $T_N$, the temperature $T_+$ will depend on the wall velocity. As a consequence, $T_+$, as well as $v_w$, depend on $T_N$ and the friction. For the stability analysis it is useful to eliminate the friction and use the wall velocity as a free parameter. However, it is not reasonable to regard $T_+$ and $v_w$ as free independent parameters. In particular, some combinations of $T_+$ and $v_w$ will be unphysical. For small supercooling (i.e., $T_N$ close to $T_c$), the reheating is given by \cite{ms09} \beg \frac{T_+}{T_c}=\frac{T_N}{T_c}+\frac{L}{\sqrt{3}w_+}v_w. \label{rehnr} \en In the case $L/w_+\ll 1$ this gives $T_+\approx T_N$. However, a small temperature variation may cause important effects on the wall dynamics. Using Eq. (\ref{rehnr}) we can write the wall velocity (\ref{vwsmallsup}) in terms of $T_N$, \beg v_w=\fr{w_+}{w_-} \frac{L(1-T_N/T_c)}{\eta_{\mathrm{eff}}}, \label{vwaproxtn} \en where we have defined an effective friction coefficient which takes into account the reheating in front of the wall \cite{ms09}. We have \beg\label{etaeff} \frac{\eta_{\mathrm{eff}}}{L}= \frac{\eta}{L}+\frac{L}{\sqrt{3}w_-}. \en Notice that the effects of reheating will depend on the two ratios $\eta/L$ and $L/w_-\sim\baL$, whose values are quite unrelated. Thus, the effective friction coefficient $\eta_{\mathrm{eff}}$ may be considerably larger than $\eta$ (even for small $\baL$). In particular, for vanishing $\eta$ we will still have a finite effective friction. This hydrodynamic obstruction to the wall motion was discussed more recently in Ref. \cite{kn11}. Notice that, for $T_+$ fixed, the velocity would only be bounded by relativity for $\eta\to 0$ [see Eq.(\ref{vwsmallsup})]. In contrast, for $T_N$ fixed, according to Eqs. (\ref{vwaproxtn}-\ref{etaeff}), the velocity may be bounded by a relatively low value. As a consequence, some velocities will be unreachable, as they would require a negative $\eta$. Similarly, the velocity $v_c$ can be written in terms of $T_N$, \beg v_c^2=\frac{T_c-T_N}{T_c}\frac{\eta}{\eta_{\mathrm{eff}}}. \label{vcn} \en As we have mentioned, in Ref. \cite{hkllm}, the approximation $T_+= T_N$ was used. Fixing $T_N$ and $v_w$ instead of $T_+$ and $v_w$, the stability parameter $\beta\approx v_w^2/v_c^2$ is enhanced, with respect to that approximation, by a factor of $\eta_{\mathrm{eff}}/\eta$. The ratio \beg \frac{\eta_{\mathrm{eff}}}{\eta}=1+\fr{1}{\sqrt{3}}\frac{L}{w_-}\fr{L}{\eta} \en can make a difference if $L/\eta$ is large. Thus, for small velocities (i.e., large $\eta$) the approximation $T_+=T_N$ will not be too bad. Notice, on the other hand, that this is a stabilizing effect. The factor $\eta/\eta_{\mathrm{eff}}$ in Eq. (\ref{vcn}) opposes to the increase with $\baL$ in Eq. (\ref{vcL}). For high velocities the approximation $T_+=T_N$ will fail as well as the approximation $T_-=T_+$. \subsection{Arbitrary velocities} \label{arbv} The previous analytic treatment may be extended beyond the limits of the above approximations (e.g., by considering higher orders in $v_w$, $\W$, or $\baL$). However, the equations are rather lengthy to write down here. We have also explored the solutions of Eq. (\ref{matriz}) in all the regions of parameter space. We found that most of the qualitative features of $\W(k)$ hold in the whole range $0<v_w<c_{s}$. Thus, the only solution with $\mathrm{Re}(\W)>0$ is real and is always bounded by the value $\W_{0}$ corresponding to $k=0$. For $k>0$ the value of $\W$ decreases. In general, $\W$ vanishes at a finite value $k_c$, and we have $\W_{0}\lesssim v_w\baL$ and $k_cd\sim\baL$ (there are some exceptions, though; see the next section). Beyond $k_c$, there may be a range $k_c<k<k_c'$ of marginal stability. For $k>k_c'$ the perturbations are exponentially stable. The general behavior of $k_c'$ is qualitatively similar to that observed analytically. We shall be interested mostly in the case of exponentially unstable perturbations $k<k_c$.\footnote{A characteristic feature is, thus, that the instability range appears continuously below a critical velocity. Furthermore, $\W$ grows continuously below $k=k_c$, and is in general small. In contrast, in the case of detonations, instabilities generally arise, below a critical velocity, with large values of $\W$ for all wavenumbers \cite{stabdeto}.} Setting $\W=0$ in Eq. (\ref{kd}) we obtain the critical wavenumber, \beg \label{kcd} k_c d=\frac{\Delta v \, \det_{14}^0}{({F_{\mathrm{dr}}}/{w_+}) \det_{34}^0/({v_+\ga_+^2})-\langle\ga v\rangle\det_{44}^0} . \en Setting $k=0$ we obtain, to linear order in $\W$, \beg \label{ommax} \W_{0} \simeq\frac{\Delta v \, \det_{14}^0}{\langle\ga \rangle\det_{44}^0-\Delta v\left[\det_{24}^0(1-v_+v_-)+\det_{14}^1\right]}. \en We have used the notations $\det_{ij}=\det_{ij}^0+\det_{ij}^1\W+\mathcal{O}(\W^2)$. We write down, as an example, the determinant $\det_{14}^0$ (i.e., $\det_{14}$ evaluated at $\W=0$), \begin{equation} \det{}_{14}^0=\Delta v \left[\gamma_{s-}^2(\gamma_--v_-b_-)- \fr{\gamma_-}{2}\right]+\fr{-v_+}{\gamma_-^2}\left\langle\gamma_{s}^2(\gamma-vb)\right\rangle , \label{det14} \end{equation} where $b_\pm$ are the stability parameters defined in Eq. (\ref{betapm2}). This determinant dominates the behaviors of $k_c$ and $\W_{0}$. In particular, it can be seen that the denominators in Eqs. (\ref{kcd}) and (\ref{ommax}) are always positive. Thus, the signs of $k_c$ and $\W_{0}$ depend essentially on the factors $\gamma_{\pm}-v_{\pm}b_{\pm}$ in Eq. (\ref{det14}) (notice that $\Delta v$ and $-v_+$ are positive). Hence, the stability is dominated by the quantities $1- \beta_{\pm}$, with \beg \label{betapmfinal} \beta_\pm =v_\pm b_\pm/\gamma_\pm . \en These definitions of $\beta_\pm$ are essentially the same as in the previous subsection, but evaluated at $\W=0$ [cf. Eqs. (\ref{Rnr}),(\ref{betanr})]. The denominators in Eqs. (\ref{kcd}) and (\ref{ommax}) become important for $v_w$ close to $c_{s-}$. Indeed, in the limit $v_-\to c_{s-}$ we have $\gamma_{s-}\to \infty$, and $\det_{14}^0$ diverges. This divergence is canceled by factors of $\gamma_{s-}$ appearing in the denominators. In any case, $\W_{0}$ and $k_c$ are still dominated by the quantities $1-\beta_\pm$, which appear in all the determinants. \subsubsection{High velocities and the Jouguet point} \label{highv} In the non-relativistic, small $\baL$ case, we have $v_\pm\simeq v_w$, $b_\pm \simeq v_w/v_c^2$ [cf. Eq. (\ref{bnr})]. As a consequence, $k_c$ and $\W_{0}$ are proportional to $1-\beta\simeq 1-v_w^2/v_c^2$. This simple expression is a consequence of the fact that the driving force is proportional to $v_c^2=1-T_+/T_c$ [cf. Eq. (\ref{fdrvcnr})]. In the general case, it is always possible to define a velocity $v_c$ which is proportional to the driving force (hence, $v_c$, as well as $F_{\mathrm{dr}}$, will vanish for $T_-=T_+=T_c$). Thus, according to Eq. (\ref{betapm2}), the quantities $b_{\pm}$ will be of the form $\langle\gamma v\rangle /v_c^2$, and we have \beg \beta_\pm\approx\fr{\gamma_\pm^{-1}\langle\gamma v\rangle v_\pm }{v_c^2} . \en Notice that we have $\beta_\pm>0$. In the case $v_+\simeq v_-$, we obtain $\beta_\pm\approx v_w^2/v_c^2$, and the behavior is similar to the non-relativistic case (namely, $\W_{0}$ will become negative for a velocity $v_{\mathrm{crit}}\approx v_c$). However, for large $v_w$ we may have a relatively large difference between $v_+$ and $v_-$. For a deflagration we have $|v_+|<|v_-|=v_w$ and, consequently, $\beta_+<\beta_-$. Therefore, the factor $1-\beta_-$ vanishes for a certain velocity $v_w\gtrsim v_c$, but $1-\beta_+$ remains positive until $v_w$ is increased further. As a consequence, the critical velocity $v_{\mathrm{crit}}$ (at which $\W_0$ and $k_c$ vanish) will be higher than $v_c$. For large $\Delta v$, we may have $v_{\mathrm{crit}}$ close to $c_{s-}$ for relatively low values of $v_c$. Moreover, $|v_+|$ is bounded by a subsonic value $v_J^{\mathrm{def}}$. If $v_c$ is higher than this value, then $1-\beta_+$ may be positive in the whole range $0<v_w<c_{s-}$. Then, it may happen that $\W_{0}$ never becomes negative, i.e., that there is no critical velocity at all. In such a case (which will depend on the amount of supercooling), the deflagration will be unstable for any subsonic velocity. Moreover, as we shall see in the next section, the values of $\W_0$ and $k_c$ may become large as $v_w$ approaches the speed of sound. This result is in clear contradiction with Ref. \cite{hkllm}, where it is claimed that it is possible to show that, in the limit $v_w\to c_s$, the equation for $\Omega$ has no positive roots, for any value of $k$. This discrepancy is, probably, due to the approximations $v_+=v_-=v_w$, $T_-=T_+$ used in \cite{hkllm} for the interface equation. Physically, the stability found for the weak deflagrations in this limit is explained in Ref. \cite{hkllm} by the fact that the result matches with the stability of detonations. However, weak deflagrations never match detonations, as the latter have higher, supersonic velocities $v_w\geq v_J^{\det}$. Between the speed of sound and the Jouguet detonation velocity $v_J^{\det}$, we may have, in principle, either strong deflagrations or Jouguet deflagrations. As we discussed in Sec. \ref{stationary}, both match the weak deflagration at $v_w=c_{s-}$ (i.e., the hydrodynamic solution bifurcates at the Jouguet point). As we have seen, the strong deflagration is unstable, whereas the supersonic Jouguet deflagration is presumably stable in general. Regardless of the behavior for $v_w\to c_{s-}$, it is easy to show that \emph{there cannot be a solution with $\Omega<0$ for $v_w=c_{s-}$}. Indeed, since $v_-=-c_{s-}$, Fig. \ref{fighip} (central panel) shows that, for $\Omega<0$, all the modes have $q<0$. Thus, we have no mode behind the wall. In front of the wall, we have $v_+>-c_{s+}$ (right panel), and we see that there are two modes ($q_1$ and $q_2$) with $q>0$. Applying the linear perturbation analysis, we will have only three unknowns (namely the amplitudes of these two modes and that of the surface deformation) for our four equations (\ref{EM4.1}-\ref{EM4.3},\ref{EM17}). This means that the analysis of linear perturbations breaks down\footnote{For supersonic Jouguet deflagrations this argument does not apply, since we are considering perturbations from a constant velocity, while this solution has a rarefaction wave immediately after the wall.} for $\Omega<0$. For $\Omega>0$, in contrast, the calculation is similar to the subsonic case (cf. the center and right panels of Fig. \ref{fighip}), only we must use Eq. (\ref{Fou6}) for $q_2$ instead of Eq. (\ref{Fou11}). We have checked that the result of such calculation matches the result of the subsonic calculation in the limit $v_w\to c_s$. \section{Numerical results} \label{numres} \subsection{The Bag equation of state} To proceed to the calculation of $\Omega(k)$, we need to consider a concrete equation of state. The simplest phenomenological model for a phase transition is the bag EOS, which consists of radiation and vacuum energy densities (see, e.g., \cite{s82}). The pressure in each phase can be written in the form \beg p_+(T)=\fr{a}{3}T^4-\fr{L}{4},\;\; p_-(T) =\left(\fr{a}{3}-\fr{L}{4T_c^4}\right)T^4. \en The entropy and enthalpy densities can be obtained from $s=dp/dT,w=Ts$. This model depends on three parameters, namely, the critical temperature $T_c$, the latent heat $L$, and the coefficient $a$. The latter is related to the number of effective massless degrees of freedom in the $+$ phase. The simplicity of the model often allows to obtain analytic results. The speed of sound is the same in both phases $c_{s\pm}=1/\sqrt{3}\equiv c_s$. The solution for the wall velocity can be obtained from Eq. (\ref{solfric}), $\eta\lan\gamma v\ran=-F_{\mathrm{dr}}$, using the matching conditions (\ref{EM3.1}-\ref{EM3.2}) and the boundary conditions. Since the pressure in both phases is a function of $T^2$, it is convenient to use Eq. (\ref{fdr2}) for the driving force. We obtain \beg F_{\mathrm{dr}}=\fr{L}{4}\left(1-\fr{T_-^2T_+^2}{T_c^4}\right). \label{fdrbag} \en The matching conditions give the relations \beg \fr{T_-^2}{T_+^2}=\sqrt{\fr{v_+ \gamma_+^2}{v_- \gamma_-^2 (1 - \baL)}}, \en \begin{equation} \label{steinhardt} v_{+}=\frac{1}{1+\alpha_+ }\left[ \frac{1}{6v_{-}}+\frac{v_{-}}{2}\pm \sqrt{\left( \frac{1}{6v_{-}}% +\frac{v_{-}}{2}\right) ^{2}+\alpha_+ ^{2}+\frac{2}{3}\alpha_+ -\frac{1}{3}}\right], \end{equation}% where \beg \baL\equiv \fr{L}{4aT_c^4/3}=\fr{L}{w_+(T_c)}, \en and \beg \alpha_+\equiv \fr{L}{4aT_+^4}=\fr{\baL}{3}\fr{T_c^4}{T_+^4}. \en The $+$ sign in Eq. (\ref{steinhardt}) corresponds to detonations, and the $-$ sign to deflagrations. For weak deflagrations, we have $v_-=-v_w$, and the reheating temperature $T_+$ is related to the nucleation temperature $T_N$ by \begin{equation} \frac{\sqrt{3}\left( T _{+}^4-T_N^4\right) }{\sqrt{\left( 3T_+^4+T_N^4\right) \left( 3T_N^4 + T_+^4\right) }} =\frac{v_{+}-v_{-}}{1-v_{+}v_{-}} . \label{tmatn} \end{equation} We define the velocity $v_c$ by \beg v_c^2 \equiv \fr{1}{4}\left(1-\fr{T_-^2T_+^2}{T_c^4}\right) \label{vcbag} \en so that we have $ F_{\mathrm{dr}}=L v_c^2 $. The velocity $v_c$ is symmetric in $T_+$ and $T_-$. For small supercooling we have $v_c^2\simeq \fr{1}{2}(1-T_+T_-/T_c^2)\simeq 1-\sqrt{T_+T_-}/T_c$. For small latent heat we have $T_-\simeq T_+$ and we recover the definition $v_c^2= 1-T_+/T_c$. From Eqs. (\ref{betapm2}) and (\ref{fdrbag}) we see that the coefficients $b_\pm$ are equal, \beg \label{bbag} b_\pm=\fr{\langle\gamma v\rangle}{v_c^2}\fr{T_+^2T_-^2}{T_c^4}, \en and the quantities $\beta_\pm$ defined in (\ref{betapmfinal}) are given by \beg \label{betabag} \beta_\pm=\fr{\gamma_\pm^{-1}\langle\gamma v\rangle v_\pm}{v_c^2}\fr{T_+^2T_-^2}{T_c^4}. \en For small supercooling and small latent heat, we have $\beta_\pm\simeq v_w^2/v_c^2$, and we have a critical velocity $v_{\mathrm{crit}}=v_c$. In the general case, the temperature ratios in Eq. (\ref{betabag}) enhance the value of $v_{\mathrm{crit}}$ with respect to $v_c$. Furthermore, the fact that $|v_+|<|v_-|$ implies that $\beta_+<\beta_-$, as already discussed. \subsection{Stability of deflagrations} In Fig. \ref{figapps} we plot the set of real solutions for $\W$ as a function of $kd$ (the right panel zooms at small $kd$). For $\mathrm{Re}(\W)>0$, we have found no other solutions (neither real nor complex). We considered values of the parameters similar to those considered in Ref. \cite{hkllm}, namely, a very small value of the latent heat ($\baL=0.01$) and a value of $T_+$ very close to $T_c$, which gives a small\footnote{See the discussion below Eq. (\ref{betanrfinal}).} critical velocity, $v_{\mathrm{crit}}\simeq v_c\simeq 0.07$. We have chosen a wall velocity below the critical one, $v_w=0.05$, so that there is a range of unstable wavenumbers. In the right panel we have also plotted the approximations (\ref{omegalinear}) and (\ref{omegalinearneg}). Notice that these do not match the exact solution even for vanishing $\W$. This is because the approximations are linear not only in $\W$, but also in the parameters $v_w,v_c$ and $\baL$. \begin{figure}[bth] \centering \epsfysize=5cm \leavevmode \epsfbox{apps.eps} \caption{$\W$ vs $kd$ for $\baL=0.01$, $T_+/T_c=0.995$, and $v_w=0.05$. The right panel shows also the linear approximations (dotted line).} \label{figapps} \end{figure} Changing the values of the parameters, the behavior is qualitatively similar (as we already discussed analytically). Essentially, the effect will be a variation of the points where the curves cut the axes, i.e., of the parameters $\W_0$, $k_c$, and $k_c'$ (see Fig. \ref{figlinear}). In Fig. \ref{figkcommax}, the two parameters which characterize the instability (namely, $k_c$ and $\W_{0}$) are plotted as functions of the wall velocity. The lower curve corresponds to the parameters of Fig. \ref{figapps} and is well approximated by the non-relativistic approximation shown in Fig. \ref{figkcommaxnr}. The other curves in Fig. \ref{figapps} correspond to higher values of the latent heat. As already seen with analytic approximations, the critical velocity increases with the latent heat. Notice, however, that this effect is quite small. Although $k_c$ and $\Omega_{0}$ are proportional to $\baL$, the critical velocity hardly varies with $\baL$. \begin{figure}[bth] \centering \epsfysize=5cm \leavevmode \epsfbox{kcommax.eps} \caption{The maximum value of $\Omega/k$ (corresponding to $k=0$) and the maximum unstable wavenumber $k_c$ (corresponding to $\W\to 0^+$), as functions of the wall velocity, for $T_+/T_c=0.995$ and, from bottom to top, $\baL=0.01$, $0.05$, $0.1$ and $0.3$.} \label{figkcommax} \end{figure} The dependence on the amount of supercooling is more important (see Fig. \ref{figommaxkcamas}). For small supercooling we have $v_{\mathrm{crit}}\simeq v_c \simeq\sqrt{1-T_+/T_c}$. As we increase the amount of supercooling, we observe that $v_{\mathrm{crit}}$ grows more quickly than $v_c$, as predicted in the previous section. Thus, e.g., for $T_+/T_c=0.92$, we have $v_c\simeq 0.28$ while $v_{\mathrm{crit}}\simeq 0.38$. We see that this behavior becomes critical for a value of $T_+/T_c\simeq 0.9$, which corresponds to $v_c\simeq 0.31$, while the critical velocity reaches a value $v_{\mathrm{crit}}\simeq 0.5$. Here, a second critical velocity appears, beyond which $\W_{0}$ and $k_c$ become positive again. Increasing slightly the amount of supercooling, the critical velocity ceases to exist and the deflagration is unstable in all the range $0<v_w<c_s$. \begin{figure}[bth] \centering \epsfysize=5cm \leavevmode \epsfbox{ommaxkcamas.eps} \caption{The values of $\W_{0}$ and $k_cd$ as functions of the wall velocity, for $\baL=0.1$ and, from bottom to top, $T_+/T_c=0.98$, $0.96$, $0.92$, $0.9$, $0.895$ and $0.89$.} \label{figommaxkcamas} \end{figure} Notice that, as we increase the supercooling, the non-relativistic regime of $v_w$ does not suffer qualitative modifications, while the relativistic regime changes considerably. This happens because the difference $\Delta v=v_+-v_-$ grows with $v_w$. As discussed in Sec. \ref{arbv}, a high value of $\Delta v$ prevents $\W_{0}$ and $k_c$ to become negative. This effect is so important that, beyond a certain amount of supercooling, $v_{\mathrm{crit}}$ disappears before reaching $c_s$. Moreover, the value of $\W_{0}$ at $v_w=c_s$ begins to grow very quickly, and the value of $k_c$ diverges. \subsection{Reheating effects} As already discussed, it is important to notice that $T_+$ is not the nucleation temperature, and should not be considered as a free parameter. The reheating in front of the wall increases with the wall velocity. Hence, if we fix $T_N$ and increase $v_w$, the temperature $T_+$ will get closer to $T_c$, reducing the value of $v_c$. To see the importance of this effect, let us consider the wall velocity as a function of the friction coefficient $\eta$. In the left panel of Fig. \ref{figvelocity} we show the relation between velocity and friction for fixed $T_+$, corresponding to some of the curves of Fig. \ref{figommaxkcamas}. The dots indicate the critical velocity. Thus, below the dots the deflagration is unstable under long wavelength perturbations ($\lambda>1/k_c$). For the upper curve, the critical velocity does not exist and deflagrations of any velocity have instabilities. Notice that (fixing $T_+$) deflagrations do not exist for small enough $\eta$. Besides, for some values of the friction, there are two possible wall velocities. \begin{figure}[bth] \centering \epsfysize=7cm \leavevmode \epsfbox{velocity.eps} \caption{The wall velocity as a function of the friction, for $\baL=0.1$. In the left panel $T_+$ is fixed. In the right panel, $T_N$ is fixed. The values of $T_+/T_c$ and $T_N/T_c$ are, from bottom to top, $0.98$, $0.96$, $0.92$, $0.9$ and $0.89$.} \label{figvelocity} \end{figure} In the right panel of Fig. \ref{figvelocity}, we fix instead the value of $T_N$ (to the same values given previously to $T_+$). For small $v_w$ (large $\eta$), the results are similar, indicating that $T_+\simeq T_N$. However, for higher $v_w$ (smaller $\eta$) it becomes apparent that the velocity is smaller than in the left panel. This is because $T_+$ is closer to $T_c$ and hydrodynamics acts as an effective friction. In particular, for small supercooling (lower curves), we see that the deflagration is always subsonic, even for $\eta=0$. This means that, depending on the parameters, not any velocity will be physically reachable. This fact may be missed when we consider $v_w$ instead of $\eta$ as a free parameter. In Fig. \ref{figommaxkcan} we plot again the values of $\W_{0}$ and $k_c$, this time fixing $T_N$ and taking into account the reheating. We considered some of the previous values of $T_N$, as well as higher amounts of supercooling. We see that the behavior is softened with respect to Fig. \ref{figommaxkcamas}. The critical velocity now grows more slowly, and reaches $v_{\mathrm{crit}}=c_s$ at a temperature $T_N\simeq 0.775$. As the amount of supercooling is increased further, wall velocities close to the speed of sound become more and more unstable. \begin{figure}[bth] \centering \epsfysize=5cm \leavevmode \epsfbox{ommaxkcan.eps} \caption{The values of $\W_{0}$ and $k_cd$ as functions of the wall velocity, for $\baL=0.1$ and different values of the nucleation temperature. From bottom to top, we have $T_N/T_c=0.98$, $0.94$, $0.9$, $0.84$, $0.8$, $0.775$, $0.77$, $0.765$ and $0.76$.} \label{figommaxkcan} \end{figure} Thus, we may have two different situations, depending on the amount of supercooling. If the latter is small enough, we have $v_{\mathrm{crit}}<c_s$ and the deflagration is stable in the range $v_{\mathrm{crit}}<v_w<c_s$. Since the weak deflagration matches the supersonic Jouguet deflagration at $v_w=c_s$, this may be an indication of the stability of the latter. On the other hand, after the critical velocity reaches the value $c_s$, the situation is inverted; velocities close to $c_s$ are the most unstable ones. This instability may be an indication of the instability of the supersonic Jouguet deflagration. This result suggests that, under the conditions which give weak deflagration velocities close to $c_s$ (a high amount of supercooling and a low friction), another solution is hydrodynamically favored, namely, the weak detonation. \section{Effects of the instability} \label{conseq} In this section we shall consider the effects of the instability on the dynamics of a cosmological phase transition. \subsection{Bubble growth and surface corrugation} As mentioned in Sec. \ref{scales}, bubbles nucleate with an initial radius which is of the order of the scale $d$. Their walls accelerate during a time which is also of order $d$, after which they reach a terminal velocity. The scale $d$ is in general much smaller than the final bubble radius or the duration of the phase transition (the latter two are related by $R_f\sim v_w\Delta t$). Indeed, although both $d$ and $\Delta t$ depend on the non-trivial dynamics of the phase transition, the former is determined by forces which are not related to the expansion rate of the universe, $H$, whereas the latter will be a fraction of the age of the Universe, $t\sim H^{-1}$. Roughly, we have $d\sim T^{-1}$ and $\Delta t\sim M_P/T^2$, where $M_P$ is the Planck mass. Hence, we have $d/\Delta t\sim T/M_P$, which is, for most phase transitions, many orders of magnitude less than 1. Therefore, the terminal velocity is reached almost immediately. During most of its growth, the bubble will be in a stationary state, unless the growth becomes unstable under small perturbations\footnote{In Ref. \cite{hkllm} the results of the stability analysis were applied to the acceleration stage (although they were derived for the stationary motion). The conclusion was that, since the terminal velocity is reached when the bubble size is $\sim d<\lambda_c$, the growth is not destabilized during this stage. We shall assume that the wall reaches the stationary state before it can become unstable.}. Let us assume that we have instabilities, and consider their growth. We shall concentrate on exponentially unstable perturbations (i.e., $\Omega>0$). We remark, though, that there is a velocity $v_c'$ around which perturbations of any wavelength are marginally unstable (see the discussion around Figs. \ref{figlinear} and \ref{figkcommaxnr}). In most cases, the behavior of the instability can be described (at least qualitatively) by the analytic approximations derived Sec. \ref{stab}, except in the limit in which both $v_{\mathrm{crit}}$ and $v_w$ are very close to $c_s$, where the behavior departs significantly from these approximations. For $v_w<v_{\mathrm{crit}}$, the essential features of the instability are already present in Link's result, namely, that $\Omega$ is of order $\baL$ and proportional to the wall velocity, as well as the dependence $\Omega(k)\propto k(k-k_c)$ [see Eq. (\ref{omegalink})]. Thus, for the purposes of the present discussion, we shall use the approximations (\ref{kcdlinear}-\ref{omegamaxlinear}), to lowest order in $\baL$. In particular, we have \beg k_cd\simeq \fr{\beta{\baL}/{2}}{\beta+{\baL}/{2}}(1-\beta). \en In most cases the parameter $\baL$ will be small. Hence, for $\beta\sim 1$ we have \beg \label{kcbeta1} k_cd\simeq \fr{\baL}{2} (1-\beta) \quad \quad (\beta\sim 1). \en However, in some cases we may have a very small velocity (e.g., due to a significant reheating during bubble expansion). In such a case we have \beg \label{kcbetall1} k_cd\simeq \beta \quad \quad (\beta\ll 1). \en In any case, we may neglect $kd$ in the denominator of Eq. (\ref{omegalinear}) and write \beg {\Omega}/{k} \simeq \W_{0}(1-k/k_c),\label{omaprox} \en with \beg \W_0\simeq \baL v_w(1-\beta)/2. \en Thus, perturbations above the critical wavelength $\lambda_c=1/k_c$ are unstable. Notice, though, that stability is recovered for $\lambda\to\infty$. Indeed, the finite value of $\W_0$ implies that $\Omega$ vanishes at $k=0$ (see Fig. \ref{figomega}, left panel). The stability of the zero mode may be understood as follows. This perturbation corresponds to acceleration of the wall without corrugation. However, we know that the uncorrugated wall has already undergone an acceleration stage and has reached a terminal velocity, ending in this stationary state. In a way, the stability for $k=0$ just confirms the existence of such a stationary state. On the other hand, if we allow the wall to be deformed, instabilities arise. In this case, the corrugation introduces a length scale, and the relevant quantity will not be the value of $\Omega$ but the dimensionless combination $\Omega/k$. The latter is a velocity and in principle should be contrasted with $v_w$. Thus, an important parameter will be $\W_0/v_w\sim k_c d\lesssim \baL/2$ (see Fig. \ref{figomega}, right panel). \begin{figure}[bth] \centering \epsfysize=5cm \leavevmode \epsfbox{omega.eps} \caption{The values of $\Omega$ and of the dimensionless quantity $\Omega/(kv_w)$ as functions of $k$, for the same parameters of Fig. \ref{figlinear}. } \label{figomega} \end{figure} As can be seen in the left panel of Fig. \ref{figomega}, there is a mode with maximum growth rate. For the approximation (\ref{omaprox}), the wavenumber of this mode is $k=k_c/2$, and the growth rate is $\Omega_{\max} =\W_0 k_c/4$. We may obtain a stability criterion by considering this mode, which has the shortest growth time $\tau \sim \Omega_{\max}^{-1}$ \cite{link,kaj}. Notice that bubbles have a finite size $R_b$ and, thus, cannot admit corrugations of arbitrary scales. At the beginning, bubbles are very small, $R_i\sim d$. From the equations above, we see that the critical wavelength is higher than that, $\lambda_c> 2d/\baL$. Hence, physical perturbations (i.e., those with $\lambda<R_b$) will be stable until bubbles reach a size $R_b>\lambda_c$. Besides, the time available for an instability to grow is bounded by the duration of the phase transition, $\Delta t\approx R_f/v_w$. The mode with the shortest growth time $\Omega_{\max}^{-1}$ will be able to develop if $R_f$ is larger than the corresponding wavelength $2/k_c$, and if $\Delta t> \Omega_{\max}^{-1}$. The two conditions are thus $R_f>2/k_c$ and $\Delta t>4/(\W_0k_c)$. The latter condition (which implies the former) gives \beg {R_f} > \fr{4/k_c}{\W_0/v_w}. \label{condlink} \en This instability criterion can be improved, if we notice that modes with longer growth times may have, on the other hand, more time available to develop. Indeed, a perturbation with wavenumber $k$ can only be formed after the bubble reaches a size $\lambda=1/k$. The time elapsed since $\lambda$ ``enters'' the bubble size until the bubble reaches a larger size $R_b$ is given by $\Delta t_k=({R}_b-1/k)/v_w$. On the other hand, the perturbation (if unstable) grows in a time $\tau\sim\Omega^{-1}$. Therefore, the mode $k$ will become dynamically important when $R_b$ is such that $\Omega\Delta t_k\gtrsim 1$. Using the approximation (\ref{omaprox}) we obtain \beg \label{stabcond} \Omega\Delta t_k = \fr{\W_{0}}{v_w}\left(1-\fr{\lambda_c}{\lambda}\right) \left(\fr{{R}_b}{\lambda}-1\right). \en This equation takes into account the fact that perturbations are linearly unstable only in the range $\lambda_c<\lambda<R_b$. The first two factors in Eq. (\ref{stabcond}) are smaller than $1$. Moreover, we have in general ${\W_{0}}/{v_w}\ll 1$. However, the last factor may be large, depending on the final bubble size. For a given $R_b$, the dynamically most relevant perturbation is now given by the maximum of $\Omega\Delta t_k$, \beg \label{omdynmax} (\Omega\Delta t_k)_{\max}=\fr{\W_{0}}{v_w} \fr{({R}_bk_c-1)^2}{4k_c{R}_b}, \en which is attained for a wavenumber \beg {k}=\fr{1}{2}\left(\fr{1}{{R}_b}+k_c\right). \label{kdynmax} \en This perturbation will be important if $(\Omega\Delta t_k)_{\max}\gtrsim 1$. If we apply this criterion to the final bubble size, for which we have $R_f\gg 1/k_c$, we obtain Eq. (\ref{condlink}). On the other hand, when the bubble size is still comparable to the critical wavelength, the criterion says that no instability is important. The instabilities become important once the bubble reaches the size \beg \label{rbinst0} {R_b^{\mathrm{inst}}}= \fr{1/k_c}{\W_0/v_w} \left[{1+\sqrt{1+\W_0/v_w}}\right]^2 . \en Since in general $\W_0/v_w$ is small, we have ${R_b^{\mathrm{inst}}}\approx (4/k_c)/(\W_0/v_w)$, as in Eq. (\ref{condlink}). The parameters $\W_0$ and $k_c$ are not independent. For $\beta\sim 1$ we can use the approximation (\ref{kcbeta1}), which gives $k_cd\approx \W_{0}/v_w\approx (1-\beta)\baL/2$. Thus, we have \beg \label{rbinst} {R_b^{\mathrm{inst}}}\simeq \left[\fr{4/\baL}{1-({v_w}/{v_{\mathrm{crit}}})^2}\right]^2 d \quad \quad (\beta\sim 1). \en For $v_w\to v_{\mathrm{crit}}$, Eq. (\ref{rbinst}) diverges, meaning that the instabilities need infinite time to develop. On the other hand, for $v_w\to 0$ we must use the approximation (\ref{kcbetall1}). Although $\W_0/v_w$ does not vanish in this limit, $k_c$ does, and we have \beg \label{rbinstbetall1} {R_b^{\mathrm{inst}}}\simeq \fr{8}{\baL}\fr{v_{\mathrm{crit}}^2d}{v_w^2}\simeq \fr{8}{\baL} \fr{\sigma/L}{v_w^2} \quad \quad (\beta\ll 1), \en where we have used the relations $d=\sigma/F_{\mathrm{dr}}$ and $F_{\mathrm{dr}}\simeq Lv_c^2$ to make explicit the fact that $v_{\mathrm{crit}}^2d$ does not depend on the temperature $T_N$. In most cases, we will have $\beta\sim 1$. Therefore, Eq. (\ref{rbinst}) can be used to determine the bubble size at which the instabilities become important. Roughly, we have $R_b^{\mathrm{inst}}/d\sim (4/\baL)^2$ (unless there is a fine tuning so that $v_w\simeq v_{\mathrm{crit}}$). Thus, $R_b^{\mathrm{inst}}$ will be in general quite higher than the initial bubble size $d$. On the other hand, as we have mentioned, the \emph{final} bubble size $R_f$ will be a fraction of the Hubble radius $H^{-1}$, which is many orders of magnitude larger than $d$ and, in general, much larger than $R_b^{\mathrm{inst}}$ as well. Notice also that, as can be seen from Eq. (\ref{rbinst0}) we have, in general, $R_b^{\mathrm{inst}}\gg 1/k_c$. As a consequence, the dynamically most relevant wavenumber will be, according to Eq. (\ref{kdynmax}), close to $k\simeq k_c/2$. As pointed out in Ref. \cite{kaj}, the stability may be recovered due to reheating. Indeed, once the shock fronts (which precede the phase transition fronts at a speed $v_{\mathrm{sh}}\simeq c_s$) meet, they may reheat the space back to a temperature $T_r$ very close to $T_c$. In such a case, a phase equilibrium stage begins, during which the regions of stable phase can grow only at the rate at which the adiabatic expansion takes the latent heat away \cite{w84,h95,ms08}. From the viewpoint of the instabilities, the effect would be, roughly, to change the boundary condition for the temperature from $T_N$ to $T_r$. As a consequence, both velocities $v_c$ and $v_w$ will decrease significantly. It is not clear which will be the value of $\beta$. Nevertheless, since $d\simeq \sigma/(Lv_c^2)$, we see that either of Eqs. (\ref{rbinst}) and (\ref{rbinstbetall1}) will give a large value of $R_b^{\mathrm{inst}}(T_r)$. To appreciate the importance of the change of the scale $d$ after reheating, we notice that the new wall velocity is proportional to the expansion rate, $v_w\sim R_b H/\baL$ \cite{h95,m04}. Thus, Eq. (\ref{rbinstbetall1}) gives an enhancement $\sim (H^{-1}/R_b)^2$, which ensures the stability of the deflagration. Notice that the time available for the instabilities to grow is the same as before ($\Delta t\sim R_f$), since the spacing between bubble centers is given by $R_f$ and the reheating will occur after a time of order $\Delta t=R_f/v_\mathrm{sh}$. The only difference is that, if a phase equilibrium stage is reached, the \emph{total} duration of the phase transition will be longer than $\Delta t$. In summary, the time and length scales are the following. After a bubble nucleates with size $\sim d$, it reaches the stationary motion in a time $\sim d$. The instabilities (provided that $v_w<v_{\mathrm{crit}}$) become dynamically relevant much later, after a time $\sim (4/\baL)^2 d$. In general, though, there will be ample time for this to happen during bubble expansion, since the final bubble size is still much larger ($R_f\sim$ a fraction of $H^{-1}$). After a time $\sim R_f$, the phase transition may end, or it may enter a phase-equilibrium stage during which stable growth is recovered. The dynamically relevant unstable modes are those of wavelengths $\sim 4d/\baL$. This is generally much shorter than the average bubble size $R_f$ by the end of the phase transition. \subsection{Deflagrations vs detonations} For strong supercooling or small friction, the stationary solution is in general a detonation with high velocity. On the contrary, for $T_N$ very close to $T_c$ or large $\eta$ we generally have weak deflagrations with small velocities. Between these two extremes, we may have coexistence of subsonic and supersonic solutions (see e.g. \cite{ms09,ms12}). In such a case, the question arises of which one will be realized during the phase transition, and of whether this can be elucidated by the analysis of instabilities. These issues were discussed previously from a different approach, namely, in the context of a numerical investigation of Eqs. (\ref{EM1.1}-\ref{EM1.2}) in a grid \cite{ikkl94}. Instead of using an approximation for the interface, the field configuration $\phi(\mathbf{x},t)$ was considered together with the fluid profile. Thus, the dynamical evolution of a phase transition front was studied, from the initial acceleration period to the collision between two bubbles. Regarding the coexistence of deflagrations and detonations, it was found in \cite{ikkl94} that, in the cases in which either of the stationary solutions is possible, it is the detonation the one which is realized during bubble expansion. Nevertheless, this seems to be due to the dynamical evolution rather than due to an instability of the deflagration, since the detonation configuration is reached without going through a deflagration configuration. According to our results, the instability of the deflagration is unrelated to the existence of detonation solutions. Although the instabilities become important for large amounts of supercooling and low friction, detonations already exist for more moderate values of these parameters, where deflagrations are still stable. Regarding possible instabilities, the results of Ref. \cite{ikkl94} were the following. The wall configuration was found to be unstable only for strong deflagrations and for detonation solutions close to the Jouguet point. In contrast, if given as an initial condition, the weak deflagration remains as such, indicating stability. This seeming contradiction with the instability of deflagrations found in the present paper and in Ref. \cite{hkllm} has a simple explanation. Due to calculation convenience, the amounts of time considered in \cite{ikkl94} (as well as the distance between bubbles) were much less than those in an actual cosmological phase transition. This is a general problem of lattice calculations. A similar calculation was carried out recently \cite{hhrw13}. In this case, the available time did not even allow the walls to reach the stationary state before they collided. As we have seen, the time needed for the instability to become dynamically relevant is much longer than the time it takes to reach the stationary state. Nevertheless, the duration of the phase transition is still much longer. \section{A physical model} \label{model} The strength of the phase transition \cite{quiros} depends on the separation between the two minima of the free energy, $\Delta \phi=\phi_+-\phi_-$, and is usually characterized by the value of $\Delta \phi/T_c$ (for instance, in the limit $\Delta\phi\to 0$ one gets a second order phase transition). However, the phase transition dynamics does not depend on this single parameter alone. The wall velocity depends essentially on three parameters. These are the amount of supercooling (which determines the pressure difference), the latent heat (which reheats the plasma slowing down the wall), and the friction coefficient. As can be seen, e.g., in Eq. (\ref{rbinst}), these parameters are relevant for the dynamics of the instabilities as well. The ratio $T_N/T_c$ determines the value of $v_c$ and, hence, of $\beta$. The ratio $L/w_+$ gives the parameter $\baL$. Finally, the friction coefficient determines the wall velocity $v_w$. Unfortunately, these parameters do not have a simple relation in general\footnote{While the latent heat can be directly computed from the free energy density $\mathcal{F}(\phi,T)$, the calculation of $T_N$ involves, first, calculating the nucleation rate (using thermal instantons \cite{linde}) and, then, considering the dynamics of the phase transition to compute the number of bubbles nucleated in a causal volume \cite{gw81}.}. For a given model, both the released energy and the amount of supercooling increase with the strength of the phase transition. Indeed, a higher value of $\Delta \phi$ implies a higher discontinuity of the energy density (i.e., a higher $L$) as well as a wider and higher barrier between minima. The latter causes the system to stay longer in the metastable minimum, i.e., a lower temperature will be reached before bubble nucleation effectively begins. A strong supercooling causes a large pressure difference between phases and, thus, favors a high wall velocity. In contrast, a large release of latent heat causes reheating and slows the wall down. Besides, the wall velocity depends on the friction coefficient. This parameter is quite difficult to calculate, model-dependent, and is the main source of uncertainty for the wall velocity\footnote{In general, it depends on the couplings of the particles with the field $\phi$ (the stronger the coupling, the higher the friction). However, it also depends on the particles interactions which determine the diffusion of particle densities near the wall \cite{micro}.}. Therefore, it is not easy to ascertain, without a detailed calculation, whether the wall velocity will be above $v_{\mathrm{crit}}$ or not. We shall address elsewhere such a detailed study of specific models. Here we wish to discuss in general the possibility that the deflagration becomes unstable in physical models, as well as some possible cosmological implications of the instability. For that aim, we shall consider the case of the electroweak phase transition, which may be quite different for different extensions of the Standard Model. \subsection{The electroweak phase transition} In Ref. \cite{hkllm}, the electroweak phase transition was considered for the minimal Standard Model (SM), with a Higgs mass $m_H=40 GeV$ in order to obtain a first-order phase transition which is strong enough to fulfil the requirement of electroweak baryogenesis ($\Delta\phi/T_c\gtrsim 1$). Still, the phase transition for such a model is relatively weak and has a small amount of supercooling. The critical velocity was found to be bounded by $0.07$, whereas microscopic calculations gave $v_w\gtrsim 0.1$ \cite{dlhll}. It was thus concluded that the propagation of the phase transition front as a deflagration is stable. For the actual value of the Higgs mass, the SM electroweak phase transition is just a smooth crossover. Nevertheless, many extensions of the SM have been considered in the literature. In particular, the Minimal Supersymmetric Standard Model (MSSM) has been extensively investigated in relation with electroweak baryogenesis (see, e.g., \cite{mssm}). Moreover, it is well known that extra scalar singlets may cause an extremely strong phase transition (see, e.g., \cite{scalars}). Thus, depending on the model (and on the model parameters) the electroweak bubble may grow either as a deflagration or as a detonation (see, e.g., \cite{ms10}), or it may even run away \cite{bm09}. The former possibility favors baryogenesis, whereas the latter two favor GW generation. In Fig. \ref{figmodels} we show the values of $L$ and $T_N$ for some extensions of the Standard Model. \begin{figure}[bth] \centering \epsfysize=7cm \leavevmode \epsfbox{modelos.eps} \caption{\small The space of the parameters $\baL$ and $T_N/T_c$ for the electroweak phase transition, for some extensions of the Standard Model (taken from Ref. \cite{lms12}). Black dots correspond to extra scalar singlets, with $g_s=12$ degrees of freedom. The strength of the phase transition increases with the coupling of the scalars to the Higgs, $h_s$. Thus, a higher $h_s$ implies a larger latent heat as well as a smaller $T_N$. The values of $h_s$ range from $0.7$ to $1.2$ and are equally spaced. Blue dots correspond to the same extension, with less singlets, $g_s=2$. In this case we have $1.4\leq h_s\leq 1.9$. Red dots corresponds to an extension with heavy fermions and bosons \cite{cmqw05}. The coupling of the fermions to the Higgs is in the range $2.2\leq h_f\leq 2.8$ (the strength of the phase transition increases with $h_f$). Green dots corrrespond to the MSSM in the light stop scenario, for stop masses $m_{\mathrm{stop}}=132,136$, and $140$GeV (from right to left).} \label{figmodels} \end{figure} We only considered phase transitions with $\Delta \phi/T>1$. Weaker phase transitions give less supercooling and smaller latent heat (in the limit of a second order phase transition we have $T_N=T_c$ and $L=0$). Thus, the models considered in Fig. \ref{figmodels} give values of the latent heat in the range $\baL\sim 10^{-2}-10^{-1}$. Higher values of $\baL$ may be possible in other models (the physical bound is $\baL<1$). On the other hand, it may be inferred from Fig. \ref{figmodels} that a very strong supercooling is hard to achieve in a physical model. Indeed, the two lower dots (those around $T_N/T_c=0.7$) were obtained for phase transitions which are already extremely strong (a slightly stronger phase transition would remain stuck in the false vacuum, causing an inflationary era). The numerical examples considered in Sec. \ref{numres} roughly spanned this region of parameter space. As we have seen, the possible instabilities need time to grow. The initial bubble radius at nucleation, which is related to the size scale $d$, is in general $R_i\sim 10 /T$. On the other hand, the instabilities become dynamically important for bubble sizes larger than $R_b^{\mathrm{inst}}\sim (4/\baL)^2d$, which is thus in the range $\sim 10^4/T-10^6/T$. The average bubble size by the end of the phase transition is a fraction of the Hubble size $H^{-1}$. The exact value depends on the whole dynamics of the phase transition, and is not easy to estimate without a complete numerical calculation. Numerical results (see, e.g., \cite{ma05,ms08}) give values which range from $R_b\sim H^{-1}$ for very strong phase transitions to $R_b\sim 10^{-5} H^{-1}$ or smaller for weak phase transitions. At the electroweak scale we have $H^{-1}\sim M_P/T^2$, with $T\sim 100\mathrm{GeV}$ and $M_P\sim 10^{19}\mathrm{GeV}$. Hence, the final bubble size for the electroweak phase transition will be in the range $R_f\sim 10^{12}/T-10^{17}/T$. Thus, this example confirms the general hierarchy $R_i\ll R_b^{\mathrm{inst}}\ll R_f$ obtained in the previous section. The most important perturbations will be those with $\lambda\sim 4d/\baL \sim 10$-$10^2 d\gtrsim R_i$. \subsection{Baryogenesis and gravitational waves} This phase transition may have several cosmological consequences, most of them depending on the dynamics of moving walls. In principle, the hydrodynamic instability may affect any of the cosmological remnants. For instance, the generation of magnetic fields due to instabilities of the bubble walls was considered in Ref. \cite{soj97}. Here we wish to discuss the generation of gravitational waves (GW) and of the baryon asymmetry of the universe (BAU), which require quite different values of the wall velocity. A successful electroweak baryogenesis requires $\Delta\phi/T_c\gtrsim 1$, so that baryon number violating processes (sphalerons) are turned off in the broken-symmetry phase, in order to avoid the washout of the generated BAU. Regarding the wall velocity, on the one hand, it should not be too large, so that sphalerons have enough time to generate baryons in the symmetric phase (sourced by $CP$-violating interactions of the wall with the particles of the plasma). On the other hand, the wall velocity should not be too small either, in order to avoid that sphalerons in the symmetric phase have enough time to reach the equilibrium and wash out the generated BAU. All in all, a relatively small wall velocity ($v_w\sim 10^{-2}-10^{-1}$) is needed. As a consequence, baryogenesis is favored for relatively weak phase transitions, which may give such small wall velocities\footnote{The fact that the incoming flow velocity $|v_+|$ is smaller than $v_w$ may increase the upper bound \cite{n11}. Moreover, the possibility of electroweak baryogenesis with detonations has been recently discussed \cite{cn12}.}. Weak phase transitions will generally have little supercooling and, consequently, small values of $v_{\mathrm{crit}}$ as well. Therefore, the presence of hydrodynamic instabilities will depend on details of the specific model. To see the effect of these potential instabilities on electroweak baryogenesis, let us assume that $v_w$ is below the critical velocity $v_{\mathrm{crit}}$. As we have seen, for a weak phase transition, the instability will become dynamically important when the bubble reaches a size $R_b^{\mathrm{inst}}\gtrsim 10^6/T$. After that moment, the growth of the bubble may be of dendritic type \cite{fa90}. One expects that the motion of the wall will become too quick to successfully generate baryons \cite{kf92,a97}. We may thus assume that baryogenesis stops as soon as bubbles reach the size $R_b^{\mathrm{inst}}$. Since the final bubble size is $R_f\sim 10^{6}R_b^{\mathrm{inst}}$, we see that the resulting BAU will be strongly suppressed with respect to a stable wall\footnote{In this argument we have used the rough approximation $R_b^{\mathrm{inst}}/d\sim (4/\baL)^2$ for $v_w<v_{\mathrm{crit}}$. Taking into account the factor $1/(1-v_w^2/v^2_{\mathrm{crit}})$ will not change the conclusion, unless $v_w$ is very close to $v_{\mathrm{crit}}$.}. We see that an accurate determination of the wall velocity becomes crucial since electroweak baryogenesis may be completely spoiled if $v_w<v_{\mathrm{crit}}$. On the other hand, we have seen that, once shock fronts meet and reheat the plasma, the motion of phase transition interfaces as stable deflagrations may be reestablished. Depending on the friction and latent heat, the value of the wall velocity during this phase-equilibrium stage may or may not be appropriate for baryogenesis \cite{h95,ma05,m01}. In case it is, bubble walls will generate baryons during the last stages of the phase transition. It is important to notice that a significant fraction of space may be spanned by the walls during this stage. Generating gravitational waves of sizeable intensity generally requires quite higher velocities ($v_w> c_s$) in order to generate a strong disturbance of the plasma (through bubble collisions and turbulence). Hence, the instability of the deflagration is preferable, as it accelerates the wall motion. In fact, gravitational waves of sizeable amplitude seem to be possible only in models with large amounts of supercooling (e.g., the lower dots in Fig. \ref{figmodels}), which give detonations with high velocities \cite{lms12}. Such models may also allow deflagrations with velocities $v_w$ close to $c_s$ or higher. In general, these models will give $v_{\mathrm{crit}}$ also close to $c_s$. According to Fig. \ref{figommaxkcan}, in this case the deflagration may have instabilities on all wavelengths (notice the divergence of $k_c$ at $v_w= c_s$). This opens the possibility of a new mechanism of GW generation, which may compete with the collisions of detonations, even for weaker phase transitions. The evolution of the system beyond the linear regime is difficult to guess. Furthermore, it will be characterized by turbulent motions of the fluid, which make the treatment more involved. The results of a simple geometrical model (described by an equation which depends only on the local geometry of the interface) suggest that the growth may be of dendritic type \cite{fa90}. This means that ``fingers'' grow out of the wall and then split into new fingers. A spherically symmetric bubble cannot generate gravitational radiation. As a consequence, the usual mechanisms (bubble collisions and turbulence) rely on the collision of bubble walls, once bubbles have grown up to there final size. The corrugation instability, in contrast, deforms the walls and stirs the fluid as soon as the bubble reaches the size $R_b^{\mathrm{inst}}\sim (4/\baL)^2R_i$, when bubbles are still much smaller than the final mean size. Therefore, the GW spectrum will be quite different. The characteristic wavelength of the gravitational radiation is given by the stirring scale. For the usual mechanisms, this is roughly the bubble size scale $R_b$, which is determined by the mean average separation between nucleation points. In the case of unstable growth, the relevant scale (or scales) will be smaller. Initially, the source of turbulence will be the unstable corrugations of the wall (accompanied by perturbations of the fluid). Thus, the initial stirring length scale is that of the most relevant unstable mode, $\lambda_{\mathrm{inst}}\sim 2/k_c\sim 4R_i/\baL$. These perturbations then grow in size and amplitude. In the case of dendritic growth, a new length scale may arise, namely, the length of the fingers. In any case, after a certain time the turbulent fluid will ``see'' also the nominal radius of the bubbles $R_b\sim R_b^{\mathrm{inst}}$. This gives another stirring scale. The bubble spacing $R_f$ may also play a role in the turbulence spectrum. As we have seen, both $\lambda_{\mathrm{inst}}$ and $R_b^{\mathrm{inst}}$ are much smaller than $R_f$. For the usual mechanisms, the (redshifted) peak of the spectrum is around the miliHertz (corresponding to $R_f\sim 10^{-2} H^{-1}$). For the unstable growing, the GW spectrum may have several peaks, some of them at frequencies much higher than that. \section{Conclusions} \label{conclu} The possibility that an observable background of gravitational waves was produced at the electroweak phase transition has motivated in the last years a renewed interest in the hydrodynamics associated to the propagation of phase transition fronts. It is well known that, while electroweak baryogenesis requires weak deflagrations with rather small interface velocities, $v_w\lesssim 0.1$, GW generation is favored by detonations or runaway solutions with ultra-relativistic velocities. Thus, the various extensions of the SM give quite different results, depending on the values of three relevant parameters, namely, the amount of supercooling, the latent heat, and the friction. In particular, small supercooling, large friction, and large latent heat will give in general small wall velocities, favoring baryogenesis. The instability of deflagrations may alter completely this picture. In this work, we have studied the hydrodynamic stability of deflagrations. We have calculated the linear instability under corrugation of the wall as a function of the relevant parameters, we have analyzed the dynamical relevance of the instabilities, and we have discussed the implications for the electroweak phase transition and its cosmological consequences. The instability of deflagration phase-transition fronts was previously considered in Ref. \cite{hkllm}. The treatment of that work improved significantly upon preceding analysis, by taking into account the perturbations of the force which drives the wall motion. This is an important aspect, since the pressure difference between phases is very sensitive to temperature variations. Unfortunately, some simplifications used for the driving force constrain the application of those results. Our approach improved several aspects of the calculation of Ref. \cite{hkllm}. In the first place, we have derived the equation for the perturbations of the wall directly from the field equation (\ref{EM1.2}), taking into account \emph{independent} perturbations of the fluid variables on either side of the wall. This is the main difference with the treatment of Ref. \cite{hkllm}. Its quantitative effect increases with the wall velocity. We have also performed a more exhaustive search of instabilities. In particular, we have looked for complex solutions of the equation for the exponential growth rate $\Omega$. In the case of a classical burning gas \cite{landau}, the unstable modes have $\mathrm{Im}(\Omega)=0$. Thus, the disturbances are not propagated but are only amplified. This feature was also found (numerically) in the work of Link \cite{link}. We investigated analytically as well as numerically this possibility for the case of a phase transition front. The result is that, indeed, we have $\mathrm{Im}(\Omega)=0$ for $\mathrm{Re}(\Omega)>0$. For small velocities and small supercooling, our results are qualitatively similar to those of Refs. \cite{link} and \cite{hkllm}. However, we have found a range of marginally unstable wavenumbers, which was not noticed in previous works. Outside this interval we have exponential (either growing or decaying) behavior. This wavenumber gap arises as a discontinuity at $\Omega =0$, and is due to the fact that the special mode $q_1(\Omega)$ jumps from one side of the wall to the other as $\Omega$ changes sign. Studying the stability in this range would require to go beyond linear perturbations. Unfortunately, the numerical analysis of Ref. \cite{fa03} did not explore regions of parameters where our results would differ from those of Ref. \cite{hkllm}. Moreover, a numerical investigation of the parameter region where linear perturbation theory predicts instabilities is still lacking. The general behavior of the linear stability is essentially the following. Below a critical velocity $v_{\mathrm{crit}}$, perturbations on wavenumbers $k$ smaller than a value $k_c$ are exponentially unstable. In general, we have $k_c\lesssim \baL/d$, and $\Omega\lesssim v_w \baL k_c$. The critical velocity depends strongly on the amount of supercooling. For small supercooling, we have $v_{\mathrm{crit}}\simeq \sqrt{1-T_+/T_c}$, in agreement with Ref. \cite{hkllm}. However, as we increase the amount of supercooling $v_{\mathrm{crit}}$ quickly departs from this simple behavior. Even taking into account the reheating effect $T_+>T_N$, the critical velocity soon approaches the speed of sound, which means that any subsonic velocity becomes unstable. Furthermore, in this case, those velocities which are closer to the speed of sound have a larger range of unstable wavenumbers and higher growth rates. This result is in disagreement with Ref. \cite{hkllm}, according to which weak deflagrations are always stable in the limit $v_w\to c_s$. The discrepancy is due to our more realistic treatment of the equation for the interface. We have briefly discussed supersonic deflagrations. The case of supersonic Jouguet deflagrations turns out to be considerably more involved, and shall be addressed elsewhere. Regarding strong deflagrations, we have checked, for the case of planar relativistic phase-transition fronts, that these are trivially unstable, by showing explicitly that the whole family of strong deflagrations (sketched in Fig. \ref{figstrong}, left panel) is not evolutionary. We have also studied the dynamical importance of the instabilities. Thus, we have improved the discussions of Refs. \cite{link,kaj}, and we have established a hierarchy of time and length scales for the growth of bubbles and instabilities. We also discussed briefly a physical model, namely, the electroweak phase transition, and considered two of its possible outcomes, namely, the BAU of the universe and a stochastic background of gravitational waves. In general, for a cosmological phase transition, the instabilities have ample time to develop, provided that $v_w<v_{\mathrm{crit}}$. This may be a serious problem for electroweak baryogenesis and deserves further investigation for specific models. On the other hand, the deflagration instability favors the production of gravitational waves, by accelerating and deforming the walls almost from the beginning of bubble growth. However, to estimate the GW spectrum would require to go beyond the linear stability analysis. \section*{Acknowledgements} This work was supported by Universidad Nacional de Mar del Plata, Argentina, grant EXA 607/12.
\section{Introduction} In preparation for the Kepler mission \citet{Brown:2011dr} produced the Kepler Input Catalog (KIC), providing spectral energy distribution fit parameters for stars in the Kepler field of view. Since then the Kepler satellite has produced high-precision light curves for some 150000 stars in this field, many of which have proven to be eclipsing binaries. These are catalogued in the Kepler Eclipsing Binary Catalogue (KEBC) \citep{Prsa:2011dx,Slawson:2011fg,Matijevic:2012di}, and number well over 2000. With this catalogue as a guide, many interesting results have been found \citep[e.g.][]{Carter:2011kx,Rappaport:2012ic,Bloemen:2012je,Armstrong:2012ie,Lee:2013ee} not least those of the circumbinary planets \citep[e.g.][]{Doyle:2011ev,Welsh:2012kl}. The parameters presented by the KIC are used in target selection for both the primary Kepler purpose of planet hunting and other guest observer programs, as well as to provide estimated radii for candidate planets \citep{Batalha:2013fg}. They have been subject to latter testing, through for example population synthesis \citep{Farmer:2013wj}. Here we aim to produce a catalogue similar to the KIC for the eclipsing binary systems of the KEBC, taking into account our new knowledge of their binary nature with the information presented by the various photometric surveys of the Kepler field. In this way we can improve upon the KIC for these binary systems, through extended wavelength coverage (particularly inclusion of the U band), and consideration of both stars. Although the primary star often dominates the observed flux, not including the secondary (as in the KIC) can lead to biases. The structure of the paper is as follows. We lay out our input data in Section \ref{sectdata}, and explain the model used to fit the observed colour bands in Section \ref{sectmodel}. We test the model against simulated stellar parameters in Section \ref{secttesting}, and present our results catalogue and parameter distributions in Section \ref{sectresults}, while Section \ref{sectdiscuss} discusses the results and observed distributions. \section{Data} \label{sectdata} \subsection{KEBC} We make use of data from the Kepler Eclipsing Binary Catalogue (KEBC). Our targeted objects' KIC Identification Numbers are taken from the Catalogue, along with the primary and secondary eclipse depth ratio, which were used to produce an estimate of the temperature ratio $T_2/T_1$ of each binary as described in Sect. \ref{t2t1gen}. We took a version of the KEBC as presented online on 18-09-2013 to give KIC IDs and eclipse depth ratios, yielding 2610 systems. Thirteen of these systems contained multiple entries in the catalogue, we took the first entry only in each of these cases. Specific KIC IDs used are presented with our results in Section \ref{sectresults}. \subsection{HES} \label{sectHES} We use photometry from the Howell-Everett Survey (HES)\citep{Everett:2012kq}. The survey consists of the three optical filters Johnson U B and V, in the Vega system \citep{Morgan:1953hy}, and contains data on 2424 objects of the 2610 from the KEBC. Errors were taken as presented in the HES catalogue. \subsection{KIS} \label{sectKIS} In parallel to the Howell-Everett Survey, we use data from the Kepler INT Survey (KIS) \citep{Greiss:2012ii,Greiss:2012uq}, Data Release 2. This allows models to be fit to two independent sets of photometry separately, increasing reliability and allowing bad data to be more easily flagged. The KIS provides data in the RGO U, Sloan g, r and i bands, in the Vega system. Errors provided in the catalogue are photometric only, we add systematic errors to the photometric errors in quadrature. We use Table 3 of \citet{Greiss:2012ii}, which gives the systematic offset used in calibrating each band to the KIC, as an estimation of the systematic error for each band (a 0.05 mag systematic error was used for the U band). Some objects in the KIS are observed more than once, these are available as separate sets of data, duplicates or triplicates (no object had more than 3 sets of data), for the same object. There are 2439 of the KEBC systems present in the KIS, of which 764 are duplicates and 111 triplicates. These multiple dataset systems were treated as independent objects during the subsequent analysis. Data for individual bands were filtered using the KIS class flag. Only bands of data with class -1 (stellar) or -2 (probably stellar) were used. \subsection{2MASS} \label{sect2MASS} To each of the HES and KIS surveys we added data from 2MASS \citep{Skrutskie:2006hl}. This consisted of the Johnson J H and 2MASS specific Ks bands, in the Vega system. The combined total photometric uncertainties were used as presented in the 2MASS catalogue. Data were accepted if the SNR in that band was $\geq 5$ and the object was not flagged as blended or contaminated. 2590 objects were found in the 2MASS catalogue. \subsection{Combination} For each object the above survey data was combined to produce two partially independent datasets, HES + 2MASS, hereafter UBVJHK, and KIS + 2MASS, hereafter UgriJHK. Each dataset is used separately in what follows, allowing comparison between results derived from the HES and KIS surveys and hence increasing reliability. \section{Model} \label{sectmodel} \subsection{Setup} We use a Markov Chain Monte Carlo code utilising the Metropolis-Hastings algorithm, enacted using the python module PyMC \citep{Patil:Huard:Fonnesbeck:2010:JSSOBK:v35i04}. We assumed each system to be composed of two stars, and fit the combined contribution from these two stars to the observed colour data. This intrinsically assumes that the data were taken when the stars were out of eclipse. This is reasonable for many eclipsing binaries, for those in overcontact systems (where the two stars are permanently in contact) it is less so but hard to avoid. In these cases it is only the apparent radii of the eclipsed star which will change; this is in general fit poorly anyway (see Section \ref{secttesting}), and the fit temperatures should be unaffected. \subsubsection{Model Atmospheres} We use the Castelli-Kurucz 04 Model Atmospheres \citep{Castelli:2004ti}. These cover a grid of [$3500K<\textrm{T}<50000K$], [$-2.5<[M/H]<0.5$], and [$0.0<\textrm{Log g}<5.0$], of which for computing efficiency purposes we used temperatures up to 13000K (some systems were run with higher temperature limits, see Sect \ref{sectresfitting}). The grid spacing was 250K in Temperature (1000K for atmospheres above 13000K), 0.5 in $[M/H]$ (with an additional point at $[M/H]=0.2$) and 0.5 in Log g. Model values were interpolated linearly between the two closest grid points of each parameter. Although the CK atmospheres depend on surface gravity and metallicity as well as temperature, we found that they were retrieved extremely poorly (See Sect \ref{secttesting}). As such we did not include them in our model, using CK atmospheres with the KIC surface gravity and zero metallicity for each system. To make use of the CK atmospheres, they must be integrated over a response function for the relevant filter to produce band-integrated flux densities. We used filter transmission curves as detailed in the respective papers of the HES and KIS. For the 2MASS data, relative spectral response functions from \citet{Cohen:2003gg} were used. These provide an absolute flux calibration using the calibrated spectrum of Vega, matching with the Vega system magnitudes of the HES and KIS. \subsubsection{Interstellar Extinction} We use the extinction relations of \citet{Cardelli:1989dp} with a constant $R_V$ of 3.1, resulting in two analytical relations relevant for optical wavelengths (U to i) and IR wavelengths (JHK). This allowed extinction in each band to be calculated as a function of that of the V-band. The specific conversion factor for each photometric band depends on the spectrum of the star under question (due to the distribution of stellar flux within the band), but we found that making the simplifying assumption of an extinction factor for each band calculated at the central wavelength of the relevant band had negligible effect for the stars considered here. We took V band extinctions from the $A_V$ values of the KIC, and used the mean value of the KIC EBs of 0.4 mag for systems where no KIC values were available. This applied to 244 systems of the 2610; these systems are flagged in the presented catalogue. While the KIC values for $A_V$ are by no means perfect (see \citet{Brown:2011dr} for a full discussion) we found that fitting them ourselves did not constrain them, and results in an additional free parameter in what is already a large parameter space. As such we use the KIC values both to constrict the parameter space and to allow easier comparison between our results and the $T_{\textrm{eff}}$ of the KIC. \subsubsection{Generation of $T_2/T_1$} \label{t2t1gen} As direct values for $T_2/T_1$ were not available at the time of submission, we estimate it from the ratio of secondary to primary eclipse depth (as $T_2/T_1 \simeq (\textrm{depth}_\textrm{sec}/\textrm{depth}_\textrm{pri})^{0.25}$). For circular binaries, this represents a good proxy for the temperature ratio. For increasingly eccentric orbits, due to the possibility of different surface areas being occulted in primary and secondary eclipse, the eclipse depth ratio becomes an increasingly less accurate estimator of $T_2/T_1$. We formed a distribution for $T_2/T_1$ by including 1) The known parameters (period, eccentricity, argument of periapse) of each binary, 2) measurement scatter in recovering the eclipse depths (gaussian errors of 0.025 and 0.05 for over contact and non-overcontact binaries respectively, from the test Figures 8 and 10 of \citet{Prsa:2011dx}) and 3) a correction for the effect of eccentricity, derived for each binary individually. For full details see Appendix \ref{appt2t1gen}. \subsubsection{Fit Parameters} Four fit parameters were used. These comprised the primary star temperature $T_1$, secondary star temperature $T_2$ (constrained through the temperature ratio as measured from the lightcurves), radius ratio of the stars $R_2/R_1$, and the primary radius to system distance ratio $R_1/D$. Note that stellar radii as used provide no allowance for non-sphericity of stars, and as such represent an `effective radius', particularly in the case of overcontact eclipsing binary systems. No constraining relations were used, each parameter was allowed to vary according to its prior (see Section \ref{sectpriors}). Observables were treated as having normally distributed errors, and comprised each available colour band. \subsubsection{Input Data and Priors} \label{sectpriors} The fits were performed to the combined UBVJHK dataset (6 colour bands), and separately to the UgriJHK dataset (7 colour bands). Each covered the wavelength range 0.36 to 2.16 \AA. Missing (not available from the relevant photometric survey catalogue) or bad as defined in Sections \ref{sectHES}, \ref{sectKIS}, and \ref{sect2MASS}) data was given an error of $10^5$ magnitudes to ensure it did not affect the fit. We treated the KEBC temperature ratio as an observable with distribution as described in Section \ref{t2t1gen}, and used this to constrain the temperature of the secondary star from that of the primary. We assumed priors on the model as detailed in Table \ref{Tablepriors}. Where no KIC $T_{\textrm{eff}}$ was available for an object, 5000K was used as the prior mean for $T_1$. We tested the effect of the prior on T1 by running 1000 of the binary systems with firstly a prior of 5000K with standard deviation 2000K, and secondly a prior of the KIC $T_{eff}$ with the same standard deviation, in each case with no extinction. The offset in the means of the $T_1$ distributions was 14K, so no significant systematic effect is caused by the prior. The standard deviation of the difference between the two cases, excluding non-converged systems, is \mytilde200K, well within the errors we quote in Section \ref{secterrors}. As such we conclude that the choice of this prior has no significant effect on the retrieved values. The `primary' star of each system was chosen using the KEBC temperature ratios - these values were taken for the purposes of fitting, even when they were greater than unity. In the final catalogue the primary star values have been set as the star with the dominant flux contribution, as calculated from the temperature and radius ratios of the model output. \begin{table*} \caption{Model Priors} \label{Tablepriors} \begin{tabular}{@{}lllllr@{}} \hline Parameter & Distribution & Parameters & & &\\ & & Mean& Standard Deviation & Lower Limit & Upper Limit\\ \hline $T_1$ & Normal & KIC $T_{\textrm{eff}}$ &2000K & 3500K & 13000K (see Sect \ref{sectresfitting}) \\ $T_2/T_1$ &See Sect \ref{t2t1gen} & & & & \\ $R_1/D$ & Normal & 0.003$R_\odot\textrm{pc}^{-1}$ &0.05$R_\odot\textrm{pc}^{-1}$ & $10^{-5}R_\odot\textrm{pc}^{-1}$ & 0.2$R_\odot\textrm{pc}^{-1}$\\ $R_2/R_1$ & Normal & 0.8 &0.3 & 0.01 & 3.0\\ \hline \end{tabular} \end{table*} \subsection{Testing} \label{secttesting} The model was tested on a simulated distribution of 1000 binary systems, for both the UBVJHK and UgriJHK datasets. These systems were generated with separate distributions for each physical stellar parameter, as no complete unbiased distribution could be found for these parameters in binary stars. We used the distributions as laid out in Table \ref{Testdistributions} (with all values constrained to be above zero), which were designed to cover the expected parameter space for the Kepler mission EBs. The distribution of Temperature Ratio $T_2/T_1$ was taken as a gaussian approximation to the distribution of the KEBC. Note that this form of test involves generating fake colours using the very model atmospheres and filter transmissions used to fit them. Also while it involves realistic parameter values, these do not combine to represent `real' stars. Hence this is purely a test of information content in the used colour bands. No significant difference was seen between each dataset, as expected from their similar colour bands and wavelength ranges. Simulated colour bands were generated via integrating over the CK04 model atmospheres as detailed above. The MCMC was run for 50000 iterations with a burn in period of 20000 iterations. No significant extinction was included in this test. The retrieved values of $T_1$, $T_2$, $R_1/D$ and $R_2/R_1$ as compared to their input values for each dataset are shown in Figures \ref{t1test}, \ref{t2test}, \ref{r1overdtest} and \ref{rrattest}. The agreement in all parameters except $R_2/R_1$ shows that 20000 is a sufficient number of iterations to allow convergence for the majority of systems. To remove as many as possible of those few remaining unconverged, a higher number of iterations is specified for the real data. The retrieval of surface gravity and metallicity was extremely poor. These parameters have very little impact on the observed colours within their error. As such these parameters were not included in the model. As shown above, the combination $R_1/D$ replaces $R_1$ and $D$ in the actual model run, as the latter two were not individually constrained. \begin{table} \caption{Distribution of Test Parameters} \label{Testdistributions} \begin{tabular}{@{}lllr@{}} \hline Parameter & Distribution & Parameters &\\ & & Mean,& Standard Deviation\\ & &Lower Limit, &Upper Limit\\ \hline $T_1$ & Uniform & 3500K &10000K \\ $T_2/T_1$ & Normal & 0.9123& 0.1668 \\ $R_1$ & Normal & 0.8$R_\odot$ &0.2$R_\odot$\\ $D$ & Uniform & 50 pc &1500 pc \\ $R_2/R_1$ & Normal & 1.0 &0.4\\ Log $g_1$ & Normal & 4.5 cgs&0.2 cgs\\ Log $g_2$ & Normal & 4.5 cgs& 0.2 cgs\\ $[M/H]$ & Uniform & -2.5& 0.5 \\ $A_V$ & Normal & 0.05 mag &0.02 mag \\ \hline \end{tabular} \end{table} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{T1testsmall.pdf}} \caption{Input and MCMC fit primary star temperatures T1 for 1000 simulated sets of stelar parameters. The dashed line shows a perfect match. Dots represent the UBVJHK dataset and crosses the UgriJHK. The small number of systems highly deviant from the dashed line have not converged.} \label{t1test} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{T2testsmall.pdf}} \caption{As Figure \ref{t1test} for T2} \label{t2test} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{R1overdtestsmall_log.pdf}} \caption{As Figure \ref{t1test} for $R_1/D$} \label{r1overdtest} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Rratiotestsmall.pdf}} \caption{As Figure \ref{t1test} for $R_2/R_1$} \label{rrattest} \end{figure} For two stars with well separated temperatures it should be possible to fit each stellar atmosphere and hence obtain information on both stellar temperatures and also the ratio of the radii of the stars (to each other and to the system distance). For the systems in the KEBC however, the two stellar temperatures in general proved to be too close to allow this, as the peak emission of each star is often located too close in wavelength to that of the other star. This led to the well-retrieved information being in general the primary star temperature $T_1$ (and through the temperature ratio the secondary temperature $T_2$) along with a combination of parameters which we term the binary solid angle, equal to $(R_1^2 + R_2^2)/D^2$. The relevance of the binary solid angle as opposed to $R_1/D$ and $R_2/R_1$ depends on the temperature ratio. For ratios significantly different from unity the individual components $R_1/D$ and $R_2/R_1$ become well retrieved. The difference between input and MCMC fit values for $R_2/R_1$ using the UBVJHK colours is shown as a function of $T_2/T_1$ in Figure \ref{rrattrattest} (with an additional 1000 systems with lower temperature ratios added to illustrate the correlation). Note the systematic offset of about -0.2 even at lower values of $T_2/T_1$. In what follows we publish both $R_1/D$ and $R_2/R_1$, as each are in some cases accurate, but users should note the above in choosing whether to use these values individually or combined into the binary solid angle mentioned. The relevance of $R2/R1$ should be determined from Figure \ref{rrattrattest} in line with the needs of the user. We note that when $T_2/T_1$ approaches unity that for main sequence stars $R_2/R_1$ should also be close to unity. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{rratiodeltavstratsmall_all.pdf}} \caption{The dependence of $R_2/R_1$ fit quality on $T_2/T_1$. The 1000 simulated sets of parameters used in Figure \ref{t1test} plus 1000 additional sets with lower temperature ratios are shown.} \label{rrattrattest} \end{figure} \section{Results} \label{sectresults} \subsection{Fitting Parameters} \label{sectresfitting} Each object was run through the MCMC for 100000 iterations, including a burn in period of 30000 iterations (significantly more than was used in the test of Section \ref{secttesting}). At 250 iteration intervals through the burn in phase, the model was tuned. Objects with KIC $T_{eff} <= 9000K$ were run using a high temperature limit of 13000K for efficiency. Other objects (including those with no KIC value for $T_{eff}$) were run with a limit of 50000K, and are flagged in the catalogue. In forming final values for each object, output parameters were checked for consistency between both datasets, and also between duplicate or triplicate UgriJHK datasets where available. We used $T_1$ as the parameter for checks (with $T_1$ defined as the temperature of the star dominating the system flux) , as this was the strongest recovered parameter while testing. First, all fits with 3 or less bands were excluded. Then, in the cases where more than one fit was available, each fit was checked against each of the others to see if it was within $3\sigma$ (using the maximum $\sigma$ of the two fits being checked). If the fits were thus consistent, they were both included in the weighted average (using their MCMC derived errors) to calculate the final set of values. This process was repeated for all possible fit combinations (the maximum number of fits for a systems is 4, one UBVJHK and 3 UgriJHK). In this way highly deviant fits (for example due to bad photometric data) can be excluded from the final values where possible. For systems where no good fits were present, final values were formed using the weighted average of all available fits, and errors by the standard deviation of those fit values. These systems are flagged in the catalogue to highlight the systematic difference between their fits and/or the lack of photometry available. \subsection{Errors} \label{secterrors} We use the MCMC derived errors to represent the gaussian noise associated with our model fitting. These have median values of 120K for $T_1$, 310K for $T_2$, \num{8.2e-5}$R_\odot/pc$ for $R_1/D$ and 0.17 for $R_2/R_1$. We also estimate the effect of extinction on the presented values, as the KIC extinctions are known to have particularly high error. We compare a run of the model on the UBVJHK dataset with the KIC extinction values (as in the final run) and with no extinction. The error on the KIC extinction values (\mytilde0.3) is of the order of the values themselves (mean \mytilde0.4), so the difference in model fits generated by removing extinction represents a reasonable estimate of the error they could cause. We found that there was a median difference between fits with and without extinction of 350K and 540K for $T_1$ and $T_2$, \num{7.3e-5}$R_\odot/pc$ for $R_1/D$, and for $R_2/R_1$ 0.15. As such, extinction has a significant effect. We treat these median extinction effects as a $1\sigma$ additional gaussian error on the presented values, and give the combined error in the catalogue. While this is an estimate, neither the KIC extinction values nor their errors are characterised enough for a more detailed approach to be meaningful. We note that this error estimation does not take account of systematic offsets caused by contamination (`third light'), bad data, or bad extinction values, and objects affected will have larger errors. While we removed bad or contaminated photometric data where it was labelled as such, and applied errors to the input temperature ratios to reduce the impact of bad values, these issues will remain for some objects. We then adopt $1\sigma$ errors formed from a combination of the effect of the extinction systematic and fitting noise values. Errors are presented individually for each system, but to give a guideline the median catalogue error is 370K in $T_1$, 620K in $T_2$, \num{2.5e-12} in $R_1/D$, and 0.23 in $R_2/R_1$. For high temperatures ($>9000K$) the errors on the temperature are larger, as seen in Figure \ref{t1test}) and discussed in Section \ref{shortcomings}. The errors on the radius parameters will vary with $T_2/T_1$, as described in Sections \ref{secttesting} and \ref{shortcomings}. The derived temperature errors are consistent with the temperatures of various systems in the KEBC which have been studied in detail, see Section \ref{sectdiscuss} for detail. \subsection{Catalogue} The format of our results is presented in Table \ref{catformat} (the full catalogue available online). For each object the results are given for each fitted dataset, along with `final' values. These are formed from the average of the available `good' fits - fits can be excluded as explained in Section \ref{sectresfitting}. Entries are flagged for various reasons; a summary of the used flags, in order in which they appear in the catalogue, is given in Table \ref{tabflags}. \begin{table} \caption{Catalogue Flags} \label{tabflags} \begin{tabular}{@{}ll@{}} \hline Column & Flag Description \\ \hline 1 & 1 if object run with 50000K temperature limit\\ 2 & 1 if no KIC $T_{eff}$, $A_V$ or Log g values available\\ 3 & 1 if candidate or proven three-body system\\ 4 & 1 if no KEBC eclipse information available\\ 5 & 1 if no matching fits available with 4+ colour bands\\ \hline \end{tabular} \end{table} \begin{table*} \centering \caption{Catalogue Format. UBV = UBVJHK, KIS=UgriJHK} \label{catformat} \begin{tabular}{@{}llll@{}} \hline Column Header & Description & Column Header & Description \\ \hline KIC ID & Kepler Input Catalogue Identifier & $T_2$ KIS$_1$, Error & -1 if None\\ $T_1$ Final, Error& Final Primary Temperature and error (K)& $R_1/D$ KIS$_1$, Error &-1 if None \\ $T_2$ Final, Error& Final Secondary Temperature and error(K)& $R_2/R_1$ KIS$_1$, Error& -1 if None \\ $R_1/D$ Final, Error & Primary Radius / System Distance and error& KIS$_1$ Bands Used & 1 if used, order -- -- -- UgriJHK \\ $R_2/R_1$ Final, Error & Component Radii Ratio and error& $T_1$ KIS$_2$, Error & -1 if None \\ $T_2/T_1$ Input & Input prior temperature ratio &$T_2$ KIS$_2$, Error & -1 if None \\ Flags & 4 digit flag of fit quality, see Table \ref{tabflags} & $R_1/D$ KIS$_2$, Error& -1 if None\\ Output Source & Final values source, UBV, KIS$_1$,KIS$_2$,KIS$_3$, 1 if used & $R_2/R_1$ KIS$_2$, Error& -1 if None \\ $T_1$ UBV, Error & -1 if None & KIS$_2$ Bands Used&1 if used, order -- -- -- UgriJHK \\ $T_2$ UBV, Error & -1 if None &$T_1$ KIS$_3$, Error& -1 if None \\ $R_1/D$ UBV, Error & -1 if None&$T_2$ KIS$_3$, Error & -1 if None \\ $R_2/R_1$ UBV, Error & -1 if None &$R_1/D$ KIS$_3$, Error& -1 if None\\ UBV Bands Used& 1 if used, order UBV-- -- -- -- JHK &$R_2/R_1$ KIS$_3$, Error& -1 if None \\ $T_1$ KIS$_1$, Error & -1 if None &KIS$_3$ Bands Used& 1 if used, order -- -- -- UgriJHK \\ \hline \end{tabular} \end{table*} \subsection{Distributions} We form distributions of our output parameters using the `final' values as detailed in the results catalogue. Systems for which no good final value could be formed were discarded (this left 2457 of the original 2610 systems, in at least one dataset), While the $R_1/D$ distribution is generally uninformative due to the unknown distance, the $T_1$ and $T_2$ distributions, and in combination the $T_2/T_1$ distribution, is worth noting. The $R_2/R_1$ distribution is in general poorly fit so is again uninformative. The results for $T_1$ and $T_2$ are shown in Figures \ref{t1dist} and \ref{t2dist}. In presenting the temperature ratio, we show the total distribution (Figure \ref{t1t2dist}), and also the distribution split by stellar spectral type. We show the results for `cool' ($T_1<5200K$, roughly K and M stars), solar type ($5200 \le T_1 < 7500K$, roughly F and G stars) and `hot' ($T_1 >= 7500K$, roughly A stars). These are presented normalised to their sample sizes in Figures \ref{gdist}, \ref{mdist}, and \ref{adist}, and are discussed in Section \ref{sectdiscuss}. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{T1dist.pdf}} \caption{The distribution of primary star temperature $T_1$, drawn from the `final' catalogue values. Systems with no consistent fits or lacking 4+ photometric bands are excluded. The cooler temperatures where the majority of our sample lies are shown in the inset.} \label{t1dist} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{T2dist.pdf}} \caption{As Figure \ref{t1dist} for secondary star temperature $T_2$.} \label{t2dist} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{T1T2dist.pdf}} \caption{The total distribution of temperature ratio $T_2/T_1$, drawn from the `final' catalogue values. Systems with $T_2/T_1$ are included as their inverse.} \label{t1t2dist} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{MCMC_trat_g52007500_nrm.pdf}} \caption{The normalised distribution of $T_2/T_1$ for solar type stars, total number 1908.} \label{gdist} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{MCMC_trat_m5200_nrm.pdf}} \caption{The normalised distribution of $T_2/T_1$ for stars cooler than 5200K, total number 303.} \label{mdist} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{MCMC_trat_a7500_nrm.pdf}} \caption{The normalised distribution of $T_2/T_1$ for stars hotter than 7500K, total number 246.} \label{adist} \end{figure} It is also worth comparing our results to the KIC itself. Assuming a single star as the KIC does will tend to focus on the primary star as the dominant source of flux. As such we compare the KIC $T_{eff}$ to our $T_1$ in Figure \ref{kiccompfig}. A general trend of an increase in our temperatures over the KIC's can be seen - this is expected, as in each of these systems an extra contribution from a usually cooler star has been included. The effect of our increased temperature limit over the KIC is also notable. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{KICcomp.pdf}} \caption{Comparison of KIC $T_{eff}$ and final catalogue $T_1$ values. The dashed line represents a 1:1 match. The difference arises from the inclusion of two stars in our model as compared to the single star of the KIC, and also our higher temperature limit.} \label{kiccompfig} \end{figure} \section{Discussion} \label{sectdiscuss} \subsection{Overview} We have applied a two stellar component MCMC model to the eclipsing binary stars of the KEBC, using a larger wavelength range than that previously utilised by the KIC. In particular we add the U band, in principle improving our results for hotter stars. We find that the useful results of our model are the effective temperatures of both stars, as well as the parameters $R_1/D$ and $R_2/R_1$. For values of $T_2/T_1$ close to unity, these parameters are not individually fit and the combination $(R_1^2+R_2^2)/D^2$ is instead. The errors on these parameters were found to be on average 370K in $T_1$, 620K in $T_2$, \num{2.5e-12} in $R_1/D$, and 0.23 in $R_2/R_1$. We anticipate that this catalogue will be of use in target selection of Kepler eclipsing binary stars, approximation of parameters and to generally inform anyone wishing to make use of these objects' extremely high-precision light curves. We note that these results use temperature ratios derived from the eclipse depth measurements of the KEBC, and as such they rely on that catalogue to that extent. While the primary star temperatures will only be slightly affected by a largely erroneous temperature ratio, the secondary star temperatures are more sensitive. \subsection{Shortcomings} \label{shortcomings} There are various issues involved in the production of this catalogue, the worst of which we summarise here. Some of these issues are in common with those of the KIC \citep{Brown:2011dr}. \subsubsection{High/Low T} The CK stellar atmospheres which we use have a lower limit of 3500K. As such, temperatures which lie close to this value ($T<3750K$) should not be taken as accurate. In terms of selecting cool stars however they can be used - while the temperature is not accurate, any object with a catalogue temperature around this level is unambiguously cool. At high temperatures ($T>9000K$) a large systematic effect, visible in Fig. \ref{t1test}, means that our temperatures have much higher errors, and are likely underestimated. Again, these are still hot stars, but the exact temperature values should not be trusted. \subsubsection{Binary Solid Angle vs $R_1/D$ and $R_2/R_1$} As described in Section \ref{secttesting}, the $R_1/D$ and $R_2/R_1$ values are generally poorly constrained. This is because for most of the systems in the KEBC, the temperatures of the two stars are close enough that little information is contained on the secondary star's relative contribution to the flux. Fig. \ref{rrattrattest} shows that only systems with $T_2/T_1\leq 0.6$ constrain $R_1/D$ and $R_2/R_1$, and in these cases there is still a systematic underestimate of \mytilde0.2 in $R_2/R_1$ which has not been adjusted for. In the other systems, which represent the majority of those here, the binary solid angle, equal to $(R_1^2 + R_2^2)/D^2$ and representing a combined measure of the level of flux incoming at the Earth is constrained instead, and can be recovered from the catalogue values as $(R_1/D)^2(1 + (R_2/R_1)^2)$. \subsubsection{Contamination} As has been mentioned previously, our model assumes only two stellar components, and does not take account of additional sources of flux. The proportion of KEBC objects with additional companions is \mytilde20\% \citep{Rappaport:2013de}, meaning they represent a significant part of our catalogue. We have removed data flagged as contaminated in the various photometric surveys used, and marked objects which are confirmed as or possible 3+ body systems \citep{Rappaport:2013de,Carter:2011kx,Derekas:2011jsa,Conroy:2013tw,Gies:2012ks}. These systems can still be of use - for example, KIC5897826 is a triple star and is tested successfully against our two star model in Table \ref{tabperformance}. In cases where more than two stellar components are present in a system, the derived physical properties will be affected to the extent that the extra companions contribute to the flux. \subsubsection{Extinction and Reddening} Our use of the KIC $A_V$ values allows us to reduce our already large parameter space while still using reasonable extinction values. Attempts to fit these values ourselves were very poorly constrained. Extinction is not generally well constrained - typical errors on the KIC E(B-V) values are of 0.1 mag, implying a 0.31 mag typical error on $A_V$ \citep{Brown:2011dr}, which is of the order of the $A_V$ values. The KIC paper itself notes the problems involved in its extinction parameters, including no account of small scale structure in the interstellar extinction. These errors are incorporated here, and will affect the systems in our catalogue. We have attempted to include the systematic effects of bad extinction in our errors (see Section \ref{secterrors}), but users should be aware that anomalously high extinctions will produce too high temperatures, and the reverse for low extinctions. \subsection{Distributions} While our individual stellar temperatures are not constrained by spectroscopy, and so have a larger potential for biases and non-physical effects to show themselves, they offer a particularly large sample size - 303 cool stars, 1908 solar type, and 246 hot stars. However, as the stars under consideration are drawn from an unconstrained sample of ages (and the expected stellar temperature varies with evolution), the temperature distributions are generally uninformative by themselves. The temperature ratio distributions split by spectral type show slight differences, however due to the lower number of samples in each bin of the cool and hot distributions there is not a strong case for any statistically significant difference. This implies that systematic effects in the fitting of different temperature stars were not strong, supporting the robustness of our results across the different temperature regimes. All of the presented distributions have potential biases present, in the effect of the particular combination of colour bands which we utilise, as a residual effect of the KEBC recovery of eclipse depths, or as a sampling effect in the eclipsing binaries which Kepler selects. \begin{table*} \caption{Performance on known systems} \label{tabperformance} \begin{tabular}{@{}llllllr@{}} \hline KIC ID & Description &T1 (Actual, K) & T1 (This Work, K) & T2 (Actual, K) & T2 (This Work, K) & Reference \\ \hline 12644769 & Kepler-16 & $4450\pm150$ & $4013\pm350$ & N/A & $3984\pm550$ & \citet{Doyle:2011ev}\\ 8572936 & Kepler-34 & $5913\pm130$ & $5937 \pm 380$ & $5867\pm130$ & $5943 \pm 610$ & \citet{Welsh:2012kl}\\ 9837578 & Kepler-35 & $5606\pm150$ & $5902 \pm 370$ & $5202\pm100$ & $4913 \pm 620$ & \citet{Welsh:2012kl}\\ 6762829 & Kepler-38 & $5623\pm50$ &$ 5834\pm 350$ & N/A & $ 3539\pm 540$ & \citet{Orosz:2012ip}\\ 10020423 & Kepler-47 & $5636\pm100$ & $ 5881\pm 350$ & $3357\pm100$ & $ 3985\pm 680$ & \citet{Orosz:2012ku}\\ 4862625 & PH1 & $6407\pm150$ & $ 6705\pm 390$ & $3561\pm150$ & $ 4626\pm 860$ & \citet{Schwamb:2012ts}\\ 10661783 & $\delta$-Scuti & $8000\pm160$ & $ 9197\pm 480$ & N/A & $ 7714\pm 810$ & \citet{Southworth:2011ks}\\ 5897826 & Triple System & $5875\pm100$ & $ 5798\pm 380$ & N/A & $ 5951\pm 750$ & \citet{Carter:2011kx}\\ 6889235 & A+WD binary & $14500\pm500$ & $ 10874\pm 470$& $9500\pm250$ & $ 9519\pm 1000$ & \citet{Bloemen:2012je}\\ 7975824 & sdB+WD binary & $34730\pm250$ & $ 32604\pm 2190$ & $15900\pm300$ & $ 10753\pm 1460$ & \citet{Bloemen:2010dc}\\ 9472174 & sdB+dm binary & $29564\pm106$ & $ 38868\pm 387$ & N/A & $ 16350\pm 2090$ & \citet{ostensen:2010dr}\\ \hline \end{tabular} \end{table*} \subsection{Performance on known objects} A limited number of these eclipsing binary systems, analysed in depth for other reasons, are available to compare with our results. While this sample is far from ideal (by definition these are systems which were selected as `interesting' for a variety of reasons, and hence unlikely to be typical) we would still hope to predict their temperatures reasonably well. We found 11 systems where detailed analysis had been done. Five of these represent interesting stellar objects (e.g. sdB+white dwarves, triple systems) whereas the other 6 were circumbinary planet hosts. In the stellar cases the primary star often dominates the flux, and in these circumstances we would expect to fit this star reasonably well. The spectroscopically derived individual temperatures are compared to ours in Table \ref{tabperformance}. We chose to contain our comparison to temperatures, as in the majority of cases no distance information is available. Encouragingly both the primary and secondary temperatures fit well, within the errors we derive. For the three hot star entries in the table, the fits are generally worse, in all cases lying outside $3\sigma$ for at least one star. They are however, unambiguously high, representing some of the highest temperatures in our catalogue. This highlights that while we can select `hot' stars reasonably well, precise temperatures at values greater than \mytilde9000K can have larger errors. This does not affect the selection of `hot' stars used in the above distributions. \section*{Acknowledgements} The authors would like to thank E. Stanway for use of computing resources to complete this work, and D. Steeghs for helpful conversations regarding the Kepler INT Survey. We also thank Andrej Pr\v{s}a for comments which improved the paper.
\section{Extrapolating the SM to Very High Scales and the Higgs Potential Instability} The main result of the first run of the LHC was the discovery of the Higgs boson, with mass $M_H\simeq 126$ GeV \cite{higgsdiscovery}, which further study has shown to be compatible with the properties expected for a Standard Model (SM) Higgs, although there is still room for some deviation in its properties \cite{higgscouplings}. Besides this great success, no trace of physics beyond the SM (BSM) has been found, and this typically translates into bounds on the mass scale of different BSM scenarios, supersymmetric or otherwise, of order the TeV \cite{LHCBSM}. If one is willing to hold on to the paradigm of naturalness, the hierarchy problem that afflicts the breaking of the electroweak (EW) symmetry would imply that BSM physics should be around the corner, probably on the reach of the LHC. In this talk I take a different attitude: I disregard naturalness as a requisite for the physics associated to the breaking of the EW symmetry and I explore the possibility that the scale of new physics, $\Lambda$, could be as large as the Planck scale, $M_{Pl}$. \begin{figure}[b] $$\includegraphics[width=0.47\textwidth]{SMRGE.pdf} \qquad \includegraphics[width=0.49\textwidth]{run1255.pdf}$$ \begin{center} \caption{\label{fig:run}\emph{Left: Evolution of SM couplings from the EW scale to $M_{Pl}$. Right: Zoom on the evolution of the Higgs quartic, $\lambda(\mu)$, for $M_h=125.7$ GeV, with uncertainties in the top mass, $\alpha_s$ and $M_h$ as indicated. (Plots taken from \cite{us}).}} \end{center} \end{figure} From that perspective, we have now in our hands a quantum field theory, the SM, that should then describe physics in the huge range from $M_W$ to $M_{Pl}$. All the model parameters have been determined experimentally, the last of them being the Higgs quartic coupling, fixed in this model by our knowledge of the Higgs mass. Fig.~\ref{fig:run}, left plot, shows the running of the most important SM couplings extrapolated to very high energy scales using renormalization group (RG) techniques. It shows the three $SU(3)_C\times SU(2)_L\times U(1)_Y$ gauge couplings getting closer in the ultraviolet (UV) but failing to unify precisely. It also shows how the top Yukawa coupling gets weaker in the UV (due to $\alpha_s$ effects, see below). The Higgs quartic coupling is also shown: it starts small at the EW scale, $\lambda(M_t)\sim 1/8$, because the Higgs turned out to be light, and gets even smaller at higher scales. The zoomed right plot in fig.~\ref{fig:run} shows that, in fact, $\lambda$ does something interesting: it gets negative at around $\mu\sim 10^{10}$~GeV. This steep behavior of $\lambda$ is due to the effect of one-loop top corrections, which represent the dominant contribution to the beta function of $\lambda$, which describes the evolution of $\lambda$ with scale. One has, at one loop: \begin{equation} \beta_\lambda =\frac{d\lambda}{d\log \mu}= \frac{1}{16\pi^2}\left\{{\color{red}\bf -6y_t^4 }+ 12y^2_t\lambda + \frac{3}{8}\left[2g^4 + (g^2 + {g'}^2)^2\right] -3\lambda (3g^2 + {g'}^2) + 24\lambda^2\right\}\ , \end{equation} where $\mu$ is the renormalization scale and $y_t$ is the sizable top Yukawa coupling, while $g$ and $g'$ are the $SU(2)_L$ and $U(1)_Y$ gauge couplings, respectively. Due to the dependence of $\beta_\lambda$ on the fourth power of $y_t$ (term in red) there is a strong dependence of the running of $\lambda$ on the top quark mass, as shown by the gray band in fig.~\ref{fig:run} (right), which corresponds to a $3\sigma$ variation of $M_t$ around is central value (as indicated). The bigger (smaller) $M_t$ is, the steeper (milder) the slope of the running $\lambda$. There is a smaller dependence of the running of $\lambda$ on the value of $\alpha_s$, which affects $\beta_\lambda$ indirectly through its effect on the running of $y_t$: \begin{equation} \beta_{y_t}=\frac{dy_t}{d\log\mu}=\frac{y_t}{16\pi^2}\left[ \frac{9}{2}y_t^2{\color{red}\bf -8g_s^2}-\frac{9}{4}g^2-\frac{17}{12}{g'}^2 \right]\ , \label{betayt} \end{equation} where $g_s$ is the $SU(3)_C$ gauge coupling. This smaller effect is illustrated in the same fig.~\ref{fig:run} (right) by the thinner $3\sigma$ pink band, with higher (lower) $\alpha_s$ corresponding to softer (steeper) running. Finally, the thinnest band, in blue, corresponds to $3\sigma$ variations in the Higgs mass, as indicated. One also sees that $\lambda$ flattens out after getting negative: in that range of scales, gauge couplings become comparable in size with $y_t$ (see Fig.~\ref{fig:run}, left) and there is a cancellation leading to $\beta_\lambda\simeq 0$. As we will discuss in the last section, at even higher scales gauge couplings dominate, turning $\beta_\lambda$ positive and eventually making $\lambda>0$ (although this might happen beyond $M_{Pl}$). The trouble with $\lambda$ becoming negative is that it will cause an instability in the Higgs potential. This is clear when one notices that: {\em 1)} the potential at very high values of the field is dominated by the quartic term, and {\em 2)} a good approximation to the full potential at some field value $h$ requires that couplings are evaluated at a renormalization scale $\mu\sim h$ (see discussion in section 3). Therefore, at very high values of the Higgs field, the potential is $V(h\gg M_t) \simeq (1/4)\lambda(\mu=h) h^4$, which for $\lambda(h)<0$ is much deeper than our vacuum at the EW scale. This instability phenomenon, caused by heavy fermions coupled to light scalars, has been known since a long time ago \cite{oldies} and has been investigated since then in the SM with increasing degree of precision \cite{stability}, especially recently \cite{us0,lindner,shap,us,Buttazzo}, after it became apparent that the Higgs mass would lie in a very special region concerning the stability of the potential. With the current precision in the Higgs mass determination and theoretical calculation of the stability bound (which will be reviewed in Sect.~3), one concludes that (given our theoretical assumptions about the absence of BSM physics) our vacuum would most likely be metastable. We should then worry about its lifetime against decay through quantum tunneling to a deeper minimum at very high field values. \section{Lifetime of the Metastable Electroweak Vacuum} The decay probability rate of the EW vacuum per unit time and unit volume can be calculated by semiclassical methods \cite{tunnel}: it is basically $\sim h_I^4 \exp(-S_4)$, where $h_I$ is the value of the field around the region of instability (the only relevant mass scale in the problem), and $S_4$ is the action of the 4-d Euclidean tunneling bounce solution, interpolating between the new phase at high field values and the EW phase. A simple analytical approximation for $S_4$ that captures the main effect\footnote{The tunneling rate has been calculated beyond tree level, including the effect of fluctuations around the bounce solution in \cite{rateloop}. Gravitational effects, which have a negligible impact on the rate, were included in \cite{rategrav}.} is in fact possible for a negative-quartic potential $V\simeq -|\lambda(h)|h^4/4$: it is $S_4\simeq -8\pi^2/(3|\lambda(h_I)|)$, showing the usual nonperturbative dependence on the coupling constant. The logarithmic dependence of $\lambda(h)$ on its argument breaks the scale invariance of the classical potential and the tunneling occurs preferentially towards the scale $h_I$ for which $\lambda(h)$ takes its minimum value (or, what is the same, for which $\beta_\lambda=0$). One gets for the decay rate the numerical estimate $dp/(dV \ dt)\sim h_I^4 exp[-2600/(|\lambda|/0.01)]$. This has to be multiplied by the 4-d space-time volume inside our past light-cone, which is basically the fourth power of the age of the Universe: $\sim \tau_U^4\sim (e^{140}/M_{Pl})^4$. It is then clear that the exponential suppression of the decay rate [for the observed value of $M_h$, which gives $\lambda(h_I)\sim -0.01$] wins over the large 4-volume factor: the decay probability is extremely small, $p\ll 1$ or, in other words, the lifetime of the metastable EW vacuum $\tau_{EW}$ is extremely long, much larger than the age of the Universe, see Fig.~\ref{fig:life}. \begin{figure}[b] $$\includegraphics[width=0.47\textwidth]{lifeprob.pdf}\qquad \includegraphics[width=0.47\textwidth]{lifetime.pdf} $$ \begin{center} \caption{\label{fig:life}\emph{Left: probability of EW vacuum decay by quantum tunneling as a function of $M_t$, for the measured values of $M_h$ and $\alpha_s$. Right: same, for the vacuum lifetime. The two branches correspond to different assumptions on the future of the universe evolution: matter dominated (labeled CDM) or cosmological constant dominated (labelled $\Lambda$CDM). (Plots taken from \cite{Buttazzo}).}} \end{center} \end{figure} \begin{figure} $$\includegraphics[width=0.5\textwidth]{run126.pdf}$$ \begin{center} \caption{\label{fig:dangerlambda}\emph{Running Higgs quartic coupling, for $M_h=126$ GeV, showing the dependence on $M_t$ and $\alpha_s$ and the parameter region (red hatched) corresponding to an unstable vacuum with lifetime shorter than the age of the Universe. (Plot taken from \cite{us}).}} \end{center} \end{figure} This conclusion could have been different if the Higgs mass were smaller, resulting in a stronger instability of the Higgs potential. This point is illustrated by fig.~\ref{fig:dangerlambda}, which shows the (hatched) region of negative $\lambda(\mu)$ that would give a decay probability of the EW vacuum of order 1 or higher ({\it i.e.} a vacuum lifetime $\tau_{EW}$ smaller than $\tau_U$). We will call this region of parameter space the instability region. Besides the danger of vacuum decay by quantum tunneling, the EW vacuum could have also decayed in the early Universe by thermal excitations over the barrier that separates it from the deeper region of the potential at high field values \cite{thermaldecay}. The condition that the EW vacuum does not decay during those high-temperature stages of the early Universe (taken into account that the potential itself is modified at finite temperature) can be used to set an upper bound on the reheating temperature $T_{RH}$ after inflation. However, it turns out that for the current values of the Higgs mass, the potential is safe in such thermal environment and the bound on $T_{RH}$ would require lower Higgs masses, $M_h\leq 122$ GeV, see \cite{us0}. Other cosmological implications can be found in \cite{cosmo}. \section{NNLO Stability Bound and Implications} From fig.~\ref{fig:dangerlambda} we also see that the possibility that the EW vacuum is absolutely stable up to $M_{Pl}$ would require values of $M_t$ and $\alpha_s$ in some $\sim 2-3\sigma$ tension with their central experimental values. Traditionally, this possibility has been phrased in terms of the so-called stability bound on $M_h$: that is, how heavy should $M_h$ be to ensure a stable potential up to $M_{Pl}$. The state-of-the-art determination of that stability bound \cite{us,Buttazzo} (at NNLO, see discussion below) gives \begin{equation} M_h[GeV] > 129.6 + 2.0\, [M_t(GeV)-173.35]-0.5\left[ \frac{\alpha_s(M_z)-0.1184}{0.0007}\right]\pm 0.3\ . \label{stabound} \end{equation} In this expression, the main error comes from the experimental uncertainty in the top mass. From a naive combination of the experimental measurements from Tevatron and ATLAS plus CMS at the LHC one gets $M_t=173.36 \pm 0.65_{exp} \pm 0.3_{th}$ GeV, see \cite{Buttazzo}. The total 1$\sigma$ error for $M_t$ in Eq.~(\ref{stabound}) has been rounded up to 1 GeV to allow for a somewhat larger theoretical error, see the discussion below. Next in importance comes the error associated with the uncertainty in $\alpha_s(M_z)=0.1184\pm 0.0007$ \cite{alphas}. Finally, the last error is theoretical and comes from an estimate of higher order corrections, beyond NNLO. Such small error has been achieved only quite recently, with refs.~\cite{shap,us,Buttazzo} being the main contributors towards this goal. In order to achieve this precision one has to calculate reliably the scalar potential in a wide range of field values, from the EW scale up to $M_{Pl}$. There are potentially large logs, $\log[h/M_t]$, that need to be resummed and this can be done using standard renormalization group techniques. We expect that the $n^{th}$-loop contribution to the effective potential, $V_n$, will have log-enhanced terms $\propto \alpha^n[\log(h/\mu)]^{n-k}$ [where $\alpha$ represents some perturbative expansion parameter, like $y_t^2/(16\pi^2)$], with a hierarchical ordering: the dominant leading-log order (LO) for $k=0$, next-to-leading-log order (NLO) for $k=1$ and so on, till the $k=n$ non-log terms. Resummation of these large logs is done by using the so-called RG-improved potential: a potential with couplings (and field) that are running with the renormalization scale, see {\it e.g.} \cite{RGV}, which is then chosen as $\mu\sim h$. In this way, a tree-level expression $V_0$ for the potential, with couplings running with their 1-loop beta functions resums the LO terms {\it to all loops}. A one-loop expression for the potential, with couplings running with their two-loop RG equations resums also the NLO terms {\it to all loops}, and so on. To match a given level of precision, the relations between running couplings and the observables that determine them has to be performed at the corresponding level of accuracy: tree-level matching for LO (as a loop-order error in the matching propagates to NLO corrections only), one-loop matching for NLO, etc. The ingredients for the NNLO calculation of the stability bound are then clear: use the RG-improved two-loop effective potential \cite{V2}, in which couplings are running with 3-loop beta functions \cite{beta3} and use 2-loop matching \cite{shap,us,Buttazzo} to relate $\lambda$ and $y_t$ to $M_h$ and $M_t$. \begin{figure} $$\includegraphics[width=0.69\textwidth,height=0.4\textwidth]{StabLOn.pdf} $$ $$ \includegraphics[width=0.69\textwidth,height=0.4\textwidth]{StabNLOn.pdf} $$ $$ \includegraphics[width=0.69\textwidth,height=0.4\textwidth]{StabNNLOn.pdf} $$ \begin{center} \caption{\label{fig:progress}\emph{Regions in the $(M_h,M_t)$ parameter space corresponding to absolute stability (green), metastability with lifetime $\tau_{EW}$ longer than $\tau_U$ (yellow), and instability, with $\tau_{EW}<\tau_U$, of the EW vacuum. The ellipses give the experimental values at 1, 2 and $3\sigma$. The different versions correspond to progressively more precise calculations, from LO to NNLO as indicated. }} \end{center} \end{figure} \begin{figure}[t] $$ \includegraphics[width=0.45\textwidth,height=0.45\textwidth]{deadoraliveG2012.pdf} \qquad \includegraphics[width=0.44\textwidth,height=0.44\textwidth]{SMht.pdf} $$ \begin{center} \caption{\label{fig:edgy}\emph{Regions in the $(M_h,M_t)$ parameter space corresponding to absolute stability (green), metastability with lifetime $\tau_{EW}$ longer than $\tau_U$ (yellow), and instability, with $\tau_{EW}<\tau_U$, calculated at NNLO. The ellipses give the experimental values at 1, 2 and 3 $\sigma$. The red-dashed lines in the zoomed-in version (left, from \cite{Buttazzo}) indicate the scale of instability, in GeV. The zoomed-out version (right, from \cite{us}) also shows the region corresponding to non-perturbative Higgs quartic below $M_{Pl}$. }} \end{center} \end{figure} In order to illustrate the need of such very precise calculation of this stability bound to determine the properties of the EW vacuum, given our precise knowledge of $M_h$ and $M_t$, fig.~\ref{fig:progress} shows the regions in $(M_h,M_t)$ parameter space corresponding to an EW vacuum that is stable (green area), metastable (with lifetime $\tau_{EW}$ longer than $\tau_U$, yellow area) or unstable (with $\tau_{EW}<\tau_U$, red region). The plots in Fig.~\ref{fig:progress} show the location of these regions resulting from a LO, NLO and NNLO calculation, from top to down. The experimental ellipses for $M_h$ and $M_t$ are also shown. This figure demonstrates that NNLO precision is crucial to answer questions about the stability of the EW vacuum. What about higher order (NNNLO) corrections? The NNLO plot shows also (dashed lines) the remaining error, obtained by combining in quadrature the (rather small) theoretical error expected from the non-inclusion of such higher order corrections and the uncertainty from $\alpha_s$: clearly a definitive answer to the stability question will require a better knowledge of the top mass rather than an even more refined theoretical calculation. In terms of the top mass, the stability bound reads \cite{Buttazzo}: \begin{equation} M_t < (171.36\pm 0.15 \pm 0.25_{\alpha_s} \pm 0.17_{M_h})\, {\mathrm GeV} = (171.36 \pm 0.46)\, {\mathrm GeV}\ , \end{equation} where, in the last expression, the theoretical error is combined in quadrature with the indicated experimental uncertainties from $\alpha_s$ and $M_h$. Concerning the impact of $M_t$ on the stability bound, there is some controversy in the literature regarding the relationship between the top mass measured at the Tevatron and LHC and the top pole-mass. Although the naive expectation would assign an error of order $\Lambda_{QCD}$ to the connection between these two numbers, a more drastic proposal has been advocated in \cite{mt}: to use instead the running top mass measured through the total production cross-section $\sigma(pp/p\bar p\rightarrow t\bar{t}+X)$ at Tevatron and the LHC, which allows for a theoretically cleaner determination of $M_t$. However, this leads to a value of the top mass compatible with the Tevatron and LHC values but with an error which is a factor of 4 worst: $M_t=173.3\pm 2.8$ GeV \cite{mt}. Of course, if one is willing to downgrade the error on $M_t$ in this way, there would still be room for absolute stability up to $M_{Pl}$ by moving into the lower range for $M_t$. Clearly, a better understanding of the theoretical errors in the top mass determination would be desirable. See \cite{mtreview} for a review of the issues involved, current status and future expectations (presumably down to $\delta M_t\sim 500-600$ MeV at the LHC) concerning this important measurement. In Fig.~\ref{fig:edgy}, the left plot shows again the different regions concerning stability of the EW vacuum calculated at NNLO, with further information on the scale of instability, in red dashed lines. The right plot shows the same NNLO stability regions [plus the region in which $\lambda(\mu)$ becomes non-perturbative below $M_{Pl}$] in a zoomed-out range for Higgs and top masses. This last version emphasizes the fact that we seem to be living in a very untypical region of parameter space, really close to the boundary for absolute stability in the narrow wedge for a long-lived EW vacuum. A complementary view of the same observation is offered by Fig.~\ref{fig:edgyPl}, which plots the different regions for the Higgs potential behaviour in the $\{\lambda(M_{Pl}),y_t(M_{Pl})\}$ plane (as these parameters should be more fundamental). There is a new phase without a vacuum at the EW scale for $\lambda(M_{Pl})<0$ and small values of $y_t(M_{Pl})$ and, inside the instability region, the dashed line delimits the range for which Planck-scale physics can play a significant role in determining the high-scale behavior of the potential. The SM location in the narrow metastability wedge is indicated by an arrow, showing once again how atypical our universe looks like. This intriguing fact has triggered many speculations concerning its possible significance \cite{shap,us, Buttazzo} including: high-scale Supersymmetry \cite{HighSUSY}, enforcing $\lambda(\Lambda)=0$ through $\tan\beta=1$; IR fixed points of some asymptotically safe gravity \cite{IRgrav}, among other ideas (even some that predate the Higgs discovery \cite{MPP}). Is $\lambda(M_{Pl})\simeq 0$ related to the fact that we also live very close to a second phase boundary, the one separating the EW broken and unbroken phases? This boundary is associated to the fact that the mass parameter in the Higgs potential, $m^2$, is extremely small in Planck units: $m^2/M_{Pl}^2\sim 0$. In this respect, it seems that the Higgs potential has a very particular form at the Planck scale, with both $\lambda$ and $m^2$ being very small. In addition, also $\beta_\lambda$ takes a special value $\simeq 0$ not far from $M_{Pl}$. Why do EW parameters seem to take such intriguing values at the Planck scale, the scale of gravitational physics, which is totally unrelated to the EW scale? No compelling theoretical explanation has been offered so far. \begin{figure}[t] $$ \includegraphics[width=0.49\textwidth,height=0.49\textwidth]{SMstabPlanckBig.pdf} $$ \begin{center} \caption{\label{fig:edgyPl}\emph{Different regions for the Higgs potential behaviour in the $\{\lambda(M_{Pl}),y_t(M_{Pl})\}$ plane. }} \end{center} \end{figure} \section{Vacuum Instability and Physics Beyond the Standard Model} Needless to say, the intriguing "near-criticality" discussed in the previous section could be just a mirage if new BSM physics appears below $M_{Pl}$ in such a way that the running of $\lambda(\mu)$ is modified significantly. Notice however, that the existence of the instability cannot be used by itself as a motivation for BSM, given the huge EW vacuum lifetime. Nevertheless, we do expect new physics BSM, {\it e.g.} to explain dark matter, neutrino masses or the matter-antimatter asymmetry and it is natural to ask how such physics could affect the near-criticality issue. We can distinguish three possibilities concerning the impact of new physics on the stability of the Higgs potential: {\em a)} it can make the stability worse; {\em b)} be irrelevant; or {\em c)} cure it. It is easy to find examples of the three options, {\it e.g.} in the framework of see-saw neutrinos, say of type I. In such scenario, neutrinos impact the running of $\lambda(\mu)$ through their Yukawa couplings, which scale like $y_\nu^2 \sim M_N m_\nu/v^2$, where $m_\nu$ is the mass of the lightest neutrinos, $M_N$ the mass of the heavy right handed ones and $v=246$ GeV is the Higgs vacuum expectation value. \begin{description} \item[a)] For sufficiently large $M_N$, the destabilizing effect of a large $y_\nu$ can make the instability much worse, even reducing the vacuum lifetime below $\tau_U$ [if $\lambda(\mu)$ is driven by this effect into the dangerous hatched region in Fig.~\ref{fig:dangerlambda}]. This would be in contradiction with our existence and can be used to set an upper bound on $M_N$, see \cite{casas,us0}. \item[b)] For values of $M_N$ significantly smaller than this upper bound, of order $M_N\simeq 10^{13-14}$ GeV for $m_\nu\simeq 0.06-1$ eV, the neutrino Yukawas would be too small to have a significant effect on the running of $\lambda$ and their presence would be irrelevant for the potential instability discussed in previous sections. \item[c)] Finally, a see-saw scenario that cures the instability is easy to build using a powerful stabilization mechanism through a heavy singlet field $S$ coupled to the Higgs as $\lambda_{HS} S^2 |H|^2$ and having a nonzero vacuum expectation value. When $S$ is decoupled, the low-energy $\lambda$ is reduced by a negative threshold effect. The apparent instability of the potential is a mirage, as $\lambda$ above the $S$ threshold is larger than a naive extrapolation in the pure SM indicates. This mechanism can be made fully compatible with a see-saw mechanism in which $M_N$ is generated by the singlet vacuum expectation value, taken to be smaller than the SM instability scale $\sim 10^{10}$ GeV, and satisfying the lower constraints on $M_N$ from leptogenesis \cite{singlet}. \end{description} Obviously, other stabilization mechanisms exist, and almost all extensions of the SM at the TeV scale will modify the behavior (or very existence) of the Higgs field at high energies. In any case, potential stability (at least in the weak sense of demanding $\tau_{EW}\gg\tau_{U}$) can be used to constrain BSM models that do not guarantee (unlike Supersymmetry) a good UV behavior of the Higgs potential. As an example, if there is in fact an instability of the potential below the Planck scale, the minimal scenario of Higgs inflation \cite{HI} (which already is known to suffer from a unitarity/naturalness problem \cite{HItrouble}) cannot be realized \cite{us0,alb}: the mechanism claimed to give a plateau at high field values requires that the potential grows like $\lambda h^4$ in the UV, with positive $\lambda$. One is then lead to non-minimal options that must cure, not only the unitarity problem but also the instability. This could in principle be achieved in the scenario proposed in \cite{GL}, through the singlet stabilization mechanism discussed above, although the range of parameters required is somewhat contrived \cite{singlet}. \section{Vacuum Instability in the Lattice} One expects the perturbative continuum approach to the calculation of stability bounds to be reliable, as the couplings remain in the perturbative range. Nevertheless, stability bounds have also been studied in the lattice in models with scalars and fermions, expected to suffer from this generic instability phenomenon. In the lattice, the stability bound appears as a lowest possible value for the scalar mass, value which is associated with the lowest possible bare scalar quartic coupling [or $\lambda(\Lambda)=0$, where $\Lambda$ is the cutoff scale]. \begin{figure}[b] $$\includegraphics[width=0.5\textwidth]{LatticeBound.pdf}$$ \begin{center} \caption{\label{fig:lattice}\emph{Lower stability bounds on $M_h$ from the lattice analysis of a Higgs-Yukawa model, taken from \cite{Karl}.}} \end{center} \end{figure} The state-of-the-art lattice analyses of stability bounds derived such bounds in a chirally invariant Higgs-Yukawa model \cite{Karl} (see also \cite{Zoltan}), which should capture the main key ingredients of the phenomenon in the SM: a scalar Higgs doublet and a fermion $(t,b)$ doublet coupled through Yukawa interactions. Lattice perturbation theory is used to obtain a lower bound $M_h(\Lambda)$ associated with $\lambda(\Lambda)=0$, bound which is later confronted with lattice simulations done for the simplest case with $N_f=1$ and $y_t=y_b$. After checking good agreement, the calculation in lattice perturbation theory is extrapolated to the more realistic case with $N_f=3$ and $y_t/y_b=0.024$. The bounds obtained are shown in Fig.~\ref{fig:lattice}, as a function of the cutoff scale. As expected, the bound is a bit higher (or the instability scale is significantly lower) than what one gets in the Standard Model for the same values of the Higgs mass, the reason being due to the non-inclusion of gauge couplings in the lattice analysis [as we saw, in the SM the top Yukawa coupling runs to smaller values in the UV due to $\alpha_s$, see Eq.~(\ref{betayt})]. From the continuum analysis we know that, if the effect of gauge couplings were included, the instability scale for $M_h\simeq 126$ GeV would appear at extremely large scales, inaccesible to lattice simulations. Nevertheless, besides confirming that the stability bound is indeed there, lattice simulations could also be useful to study some truly non-perturbative effects associated with the physics of a metastable vacuum. In particular, they could be used to study the tunneling process by which the metastable vacuum decays. One (apparent) obstacle for this, which has been previously discussed in the literature, is the need of accessing the field range with negative values of the Higgs quartic coupling, as this is the field range at which the potential is lower than the EW vacuum, and towards which the tunneling occurs. This is a problem for the lattice as properly defining the theory being put in the lattice requires $\lambda(\Lambda)\geq 0$. This difficulty, which lead some authors to even doubt the very existence of an instability in the potential \cite{KutiHolland}, can be easily circumvented, as we will see next. First, this is not a problem for the continuum calculations: one simply assumes that some BSM physics will appear at some scale heavier than the instability scale, eventually stabilizing the potential (and therefore creating a new vacuum at large field values, which is deeper than the EW one). The actual calculation of the decay rate of the EW vacuum is not sensitive to the details of this stabilization under some reasonable assumptions: that it occurs well above the instability scale (which is roughly the scale that controls the tunneling and can be orders of magnitude below the scale at which the minimum appears) and that the instability of the potential is not made worst by the heavy new physics before it gets stabilized at even higher energies (a condition violated by the analysis in \cite{UVtrouble}). \begin{figure}[t] $$\includegraphics[width=0.45\textwidth,height=0.3\textwidth]{examplerun.pdf}\qquad \includegraphics[width=0.45\textwidth]{exampleV.pdf} $$ \begin{center} \caption{\label{fig:dip}\emph{Example of parameter choice leading to a Higgs quartic coupling getting negative in a limited range of scales, with $\lambda(\Lambda)>0$ (left plot) and the corresponding Higgs potential (right plot), showing the deep non-standard minimum.}} \end{center} \end{figure} Moreover, in the pure SM there are some special choices of $M_t$, $M_h$ and $\alpha_s$ for which the unstable potential gets stabilized in the UV simply by the RG flow of couplings: with $y_t$ getting smaller and smaller in the UV (due to the effect of $\alpha_s$), eventually EW gauge couplings make $\beta_\lambda$ positive and $\lambda(\mu)$ turns positive after an interval in which it is negative. This is a well known possibility, illustrated by Fig.~\ref{fig:dip}, which shows the running $\lambda(\mu)$ and the corresponding Higgs potential (in a log-log plot) for one such parameter choice. Such behavior demonstrates in principle that it should be possible to study the potential instability in the lattice: one can start at some heavy cutoff with a well defined theory with $\lambda(\Lambda)>0$, in spite of which, the theory can develop an instability at some intermediate range of scales that can in principle be much smaller than the cutoff (if one wishes to get rid of potential cutoff artifacts). As we have mentioned already, relying on $\alpha_s$ as a way to diminish the destabilizing impact of the Yukawa coupling results in an instability scale that is many orders of magnitude above the electroweak scale, beyond the reach of lattice simulations. In order to be able to study in the lattice the decay of the metastable vacuum, some other (more efficient) stabilization mechanism should be used to ensure that the true minimum is closer to the EW scale. One possibility is to use the singlet mechanism \cite{singlet} discussed in the previous section, but in this case choosing the mass of the singlet somewhat above the instability scale, in such a way that stabilization of the potential occurs only after an interval of scales with $\lambda(\mu)<0$. Another possibility, which is currently under investigation (see talk by Attila Nagy \cite{attila} at this conference), is to stabilize the potential in the UV by higher order operators. \section{Conclusions} We finally have data to explore the physics of EW symmetry breaking. So far, we have learned that the breaking is associated with a Higgs scalar of mass $M_h\simeq 126$ GeV that looks very much compatible with SM expectations, although there is still room for deviations in the properties of this scalar particle. On the other hand, the promise of nearby BSM physics, based on naturalness arguments, has not been fulfilled yet. If one is willing to take this as indication of a fine-tuned Higgs sector, extrapolation of the SM to very high energies reveals a potentially dangerous instability in the Higgs potential. According to this, we would be living in a metastable EW vacuum, which however has a lifetime enormously large compared with the age of the Universe. The previous statement can be understood as resulting from the fact that we live in a very particular region of parameter space in the $(M_h,M_t)$ plane: very close to the boundary that separates the region of full stability of the potential from that of metastability. Only time will tell whether there is a deeper meaning in that intriguing fact or whether new physics awaits us in the next LHC run (as expected from naturalness of EW breaking) that would expose this coincidence as a pure accident. \section*{Acknowledgments} I would like to thank my collaborators on this topic over many years: M. Quir\'os, A. Casas, A. Riotto, G. Giudice, J. Ellis, A. Hoecker, A. Strumia, H.M. Lee, G. Isidori, J. Ellias-Mir\'o, G. Degrassi and S. di Vita. This work has been partly supported by Spanish Consolider Ingenio 2010 Programme CPAN (CSD2007-00042) and the Spanish Ministry MICNN under grants FPA2010-17747 and FPA2011-25948; and the Generalitat de Catalunya grant 2009SGR894.
\section{INTRODUCTION AND OBSERVATION CONSTRAINTS} As is well known, HST-1 is the innermost knot of the M87 jet, located $\sim$80 pc from the core (Biretta et al. 1999). The multi-wavelength light curves of HST-1 have been previously studied using radio data (Chang et al. 2010), optical and UV data (Perlman et al. 2003; Madrid 2009) and X-ray data (Harris et al. 2003, 2006, 2009). Chen et al. (2011) investigated the radio polarization and spectral variability of HST-1, and Perlman et al. (2011) researched the optical polarization and spectral variability of the M87 jet. Cheung et al. (2007) argued that HST-1 may be the site of the flaring TeV gamma-ray emission reported by the H.E.S.S. (Aharonian et al. 2006). A study of particular note was by Abramowski et al. (2012), who released 10 yr of multi-wavelength observations of M87 and the very high energy $\gamma$-ray flare of 2010. The quasi-simultaneous spectrum (from the radio to X-ray band; e.g., Marshall et al. 2002; Waters \& Zepf 2005; Perlman \& Wilson 2005; Harris et al. 2006; Cheung et al. 2007) and polarization observations (e.g., Perlman et al. 1999, 2011; Chen et al. 2011) of the M87 knots demonstrate the nature of synchrotron radiation. From the multi-wavelength light curves of HST-1 (Fig. 1), we could obtain some physical constrains on HST-1: the light curves show two big flares, with the main peaks of the light curves around the year 2005.30 and the second ones around the year 2007. Here, we discuss a pure (or single) process (Doppler effect). For synchrotron emission in the case of a moving sphere, with observed fluxes $I_{\nu,\mbox{\scriptsize obs}}=\delta^{3+\alpha}I_{\nu}$ (Dermer 1995; $\delta$ is the Doppler factor of HST-1 and $\alpha=(p-1)/2$, where $p$ is the spectral index of the particles), the change of the Doppler factor may explain the change in the light curve. However, Harris et al. (2006) suggested a modest beaming synchrotron model with a Doppler factor of three or four, while Wang \& Zhou (2009) obtained the Doppler factor of HST-1 to be $3.57\pm0.51$ through a synchrotron model fitting. The HST-1 complex could be model fitted with multiple components (e.g., Cheung et al. 2007; Giroletti et al. 2012); in this case, the speed of one component in HST-1 is obtained through a longer monitoring of the same one, but the velocity estimate usually has fairly large uncertainties. Giroletti et al. (2012) reported the apparent velocities of very long baseline interferometry (1.7 GHz Very Long Baseline Array, VLBA and 5 GHz European VLBI Network, EVN) components in HST-1 with high precision ($<$$2\%$, this is an unprecedented accuracy for determining the apparent velocities of components in HST-1) during the decay period of the HST-1 flare, which implies that components in HST-1 have uniform motion with high precision. Hence, the variation range of the Doppler factor may be very small and the change in the Doppler factor in HST-1 may not explain the order of the flux change. Excluding the aforementioned single mechanism, we believe that the flare of HST-1 may be correlated with some complex processes which include a changing magnetic field strength. In Section 2, we describe in detail our model for HST-1. In Section 3, we present and discuss the fitting results of this model to the main peak in the multi-wavelength light curves of HST-1. A summary is provided in Section 4. \section{The Model} Tavecchio \& Ghisellini (2008, TG08) suggested a two-zone scenario in subparsec-scale jets to explain the TeV emissions in M87: a slow hollow cylindrical layer (the velocity relative to the M87 core is $v_l$) surrounds a fast cylindrical zone (the velocity relative to the M87 core is $v_s$, $v_s\gg v_l$, and the velocity of the inner zone relative to the cylindrical layer is $v$). The high energy from GeV to TeV could be produced through inverse Compton by the two-zone, and the photons from radio to X-ray are mainly radiated by the fast inner zone. If this subparsec-scale structure is located within HST-1, HST-1 may be a TeV emission source. However, the model of TG08 could not explain the multi-wavelength light curves (or flare) from radio to TeV. We found that a modified scheme of the TG08 model could achieve this; here we only discuss the light curves of the inner blob from radio to X-ray. We believe that the slow layer may be an hourglass-shaped or Laval nozzle-shaped layer connected by two revolving exponential surfaces (Fig. 2). Considering magnetic field conservation, we believe that the magnetic field will change along the axis of the hourglass. We assumed that the length of the inner blob along the jet axis may be smaller than the radius $R_s$ of its cross section, which may be subparsec-scale, and may remain unchanged (the blob may be constrained in a series of shocks along the jet axis). If the axis coordinate and radius of the hourglass nozzle are $x_n$ and $R_n$, respectively, then the layer radius $R$ (we assumed that $R\sim R_s$) as a broken exponential function of the axis coordinate is \begin{equation} R=\left\{ \begin{array}{ll} R_n e^{k(x_n-x)},&\mbox{~$x<x_n$;}\\ R_n e^{k^{'}(x-x_n)},&\mbox{~$x>x_n$,}\\ \end{array} \right. \end{equation} \noindent where $k$ and $k^{'}$ are constant. Considering magnetic field conservation, $B\propto R^{-2}$, the magnetic field $B$ along the axis of the hourglass will be \begin{equation} B=\left\{ \begin{array}{ll} B_n e^{-2k(x_n-x)},&\mbox{~$x<x_n$;}\\ B_n e^{-2k^{'}(x-x_n)},&\mbox{~$x>x_n$,}\\ \end{array} \right. \end{equation} \noindent where $B_n$ is the magnetic field of the hourglass nozzle. When a blob travels down the axis of the first bulb in the hourglass, because of magnetic field conservation, its cross section experiences an adiabatic compression (compression velocity $u=-(dR/dt)=R_n k v e^{k(x_n-x)}$), which results in particle acceleration and the brightening of HST-1 (this may explain why Perlman et al. 2011 found no evidence for the motion of the flaring blob of HST-1 in their optical data from \emph{Hubble Space Telescope} (\emph{HST}) observations but Giroletti et al. 2012 found no prominent stationary components in their radio data from VLBA and EVN observations). When the blob moves into the second bulb of the hourglass, considering magnetic field conservation, the dimming of the knot will occur along with an adiabatic expansion of its cross section ($u=-(dR/dt)=-R_n k^{'} v e^{k^{'}(x-x_n)}$). When a blob approaches the hourglass nozzle, the energy gains of the particles in the blob are $(\partial E/\partial t)=\alpha_2 E$, where $\alpha_2=(2/3)(u/R)=(2/3)kv>0$ (Pacholczyk 1970). The compression timescale $\tau=(R/u)=(2/3\alpha_2)$. A modified continuous injection model, which considers that source particles (a broken power law which could be interpreted as partial loss or escape of high-energy particles in the acceleration region) are continuously injected into an adjacent radiation region from an acceleration region, can be fit to the spectral energy distribution of the outer jet in M87 (Liu \& Shen 2007; Sahayanathan 2008 presented a similar model with two spectral indices). We suppose that the radiation mechanism of the inner jet is the same as that of the outer jet. When the source particles (a broken power law) have been injected for the time interval of $t_1$, the radiation region of a blob in HST-1 moves into the hourglass layer. Then, the initial condition of the particle spectra (this is a relic of a past process and may not be ignored for the next process) in HST-1 at the beginning of the compression process is (Liu \& Shen 2007; Sahayanathan 2008) \begin{equation} N^{*}(E, \theta, 0)=\left\{ \begin{array}{ll} q_1^{*} t_1 E^{-p_1},&\mbox{~$E\ll E_b$;}\\ q_2^{*} t_1 E^{-(p_2+1)},&\mbox{~$E_b\ll E\ll \frac{1}{\beta_0 t_1}$,}\\ \end{array} \right. \end{equation} \noindent where $N^{*}$, $q_1^{*}$, and $q_2^{*}$ refer to the entire radiation region in the blob taken as an entity (but $N$, $q_1$, and $q_2$ are applied to a unit volume); $\theta$ is the pitch angle between the magnetic field and the particle; $E_b$ is the break energy of a broken power law (here, we assumed that $E_b<(1/\beta_0 t_1)$; Sahayanathan 2008); $p_1$ and $p_2$ denote particle spectrum indices that may be different (e.g., Sahayanathan 2008); $\beta_0=bB^2_{\perp}$; $b$ is a constant; and $B_{\perp}$ represents the component of the magnetic field perpendicular to the velocity of the particle. Because the acting timescale of the compression process is shorter than that of the particle injection process, which may be comparable with the kinetic timescale of HST-1 relative to the M87 core, we can ignore the influence of particle injection in the compression process. Next, we consider synchrotron radiation and an adiabatic compression of the radiation region in the blob, and the kinetic equation is (Kardashev 1962) \begin{equation} \frac{\partial N^{*}}{\partial t}=-\alpha_2 \frac{\partial}{\partial E}(EN^{*})+\beta \frac{\partial}{\partial E}(E^2 N^{*}), \end{equation} Under the initial conditions given in Equation (3), we have \begin{equation} N^{*}(E, \theta, t)= q^{*}t_1 E^{-\lambda }[1-E e^{-a_2}\int_{0}^{t} \beta e^{a_2} dt]^{\lambda-2}e^{(\lambda-1)a_2}, \end{equation} \noindent where $a_2=\int_{0}^{t} \alpha_2 dt$, $q^{*}$ represents $q_1^{*}$ or $q_2^{*}$, $\lambda$ is a substitute for $p_1$ or $p_2+1$, and we have assumed that $E\ll (e^{a_2}/\int_{0}^{t} \beta e^{a_2} dt)$ and $\alpha_2=$const in the formula (5). Hence, \begin{equation} N^{*}(E, \theta, t)=\left\{ \begin{array}{ll} q_1^{*} t_1 E^{-p_1}e^{(p_1-1)\alpha_2 t}[1-\frac{\beta}{7\alpha_2 }E(1-e^{-7\alpha_2 t})]^{p_1-2},&\mbox{~$E\ll E_b$;}\\ q_2^{*} t_1 E^{-(p_2+1)}e^{p_2\alpha_2 t}[1-\frac{\beta}{7\alpha_2}E(1-e^{-7\alpha_2 t})]^{p_2-1},&\mbox{~$E_b\ll E\ll \frac{7\alpha_2}{\beta (1-e^{-7\alpha_2 t})}$.}\\ \end{array} \right. \end{equation} Considering the conservation of magnetic flux, $B\propto R^{-2}$, $R=R_0 e^{-k v t}=R_0 e^{-(3/2)\alpha_2 t}$, where $R_0$ is the initial radius of the blob. Hence, $\beta=\beta_0 e^{6\alpha_2 t}$. The factors $[1-(\beta/7\alpha_2)E(1-e^{-7\alpha_2 t})]^{p_1-2}$ and $[1-(\beta/7\alpha_2)E(1-e^{-7\alpha_2 t})]^{p_2-1}$ in Formula (6) are close to 1, because of the following reasons. (1) $E\ll (7\alpha_2/\beta (1-e^{-7\alpha_2 t}))$, and so $1-(\beta/7\alpha_2)E(1-e^{-7\alpha_2 t})\rightarrow 1$. (2) $p_1, p_2\sim 2$ in the M87 jet (Perlman \& Wilson 2005; Liu \& Shen 2007; Perlman et al. 2011), hence $p_1-2\sim 0$ and $p_2-1\sim 1$. Formula (6) is thus close to \begin{equation} N^{*}(E, \theta, t)\approx \left\{ \begin{array}{ll} q_1^{*} t_1 E^{-p_1}e^{(p_1-1)\alpha_2 t},&\mbox{~$E\ll E_b$;}\\ q_2^{*} t_1 E^{-(p_2+1)}e^{p_2\alpha_2 t},&\mbox{~$E_b\ll E\ll \frac{7\alpha_2}{\beta (1-e^{-7\alpha_2 t})}$,}\\ \end{array} \right. \end{equation} and when this is converted to unit volume (entire volume $V\propto R^2 \propto e^{-3\alpha_2 t}$), \begin{equation} N(E, \theta, t)=\frac{N^{*}}{V}\approx \left\{ \begin{array}{ll} q_1 t_1 E^{-p_1}e^{(p_1+2)\alpha_2 t},&\mbox{~$E\ll E_b$;}\\ q_2 t_1 E^{-(p_2+1)}e^{(p_2+3)\alpha_2 t},&\mbox{~$E_b\ll E\ll \frac{7\alpha_2}{\beta (1-e^{-7\alpha_2 t})}$.}\\ \end{array} \right. \end{equation} Next, we consider the increase in magnetic field strength and the time dependence of blob length along the line of sight ($\propto e^{-(3/2)\alpha_2 t}$). If the distributions of particles are isotropic, we can derive the flux formula of the synchrotron model: \begin{equation} I_{\nu}\propto\left\{ \begin{array}{ll} e^{(5p_1+4)\alpha_2 t/2}\nu^{-(p_1-1)/2},&\mbox{~$\nu\ll \nu_{B1}$;}\\ e^{(5p_2+9)\alpha_2 t/2}\nu^{-p_2/2},&\mbox{~$\nu_{B1}\ll \nu\ll \nu_{B2}$,}\\ \end{array} \right. \end{equation} \noindent where $\nu_{B1}$ and $\nu_{B2}$ are break frequencies. When the radiation region of the blob in HST-1 moves away from the hourglass and into the adjacent bulb, deceleration of particles in the blob may occur. Then, $\alpha_2 ^{'}=-(2/3)k^{'}v<0$, the expansion timescale $\tau^{'}=-(2/3\alpha_2 ^{'})$, and the formula for the particle spectrum will become \begin{equation} N^{'}(E, \theta, t)\approx \left\{ \begin{array}{ll} q_1 t_1 E^{-p_1}e^{(p_1+2)\alpha_2 t_2}e^{(p_1+2)\alpha_2 ^{'} (t-t_2)},&\mbox{~$E\ll E_b$;}\\ q_2 t_1 E^{-(p_2+1)}e^{(p_2+3)\alpha_2 t_2}e^{(p_2+3)\alpha_2 ^{'} (t-t_2)},&\mbox{~$E_b\ll E\ll \frac{7\alpha_2^{'}}{\beta^{'} [1-e^{-7\alpha_2^{'}} (t-t_2)]}$,}\\ \end{array} \right. \end{equation} \noindent where $t_2$ is the acting timescale of the compression process, $\beta^{'}\propto e^{6 \alpha_2^{'} (t-t_2)}$. Now, we consider the decrease of magnetic field strength and the time dependence of blob length along the line of sight ($\propto e^{-(3/2)\alpha_2^{'} (t-t_2)}$). With this consideration, the flux expression would be changed to \begin{equation} I_{\nu}^{'}\propto\left\{ \begin{array}{ll} e^{(5p_1+4)\alpha_2 ^{'} t/2}\nu^{-(p_1-1)/2},&\mbox{~$\nu\ll \nu_{B1}$;}\\ e^{(5p_2+9)\alpha_2 ^{'} t/2}\nu^{-p_2/2},&\mbox{~$\nu_{B1}\ll \nu\ll \nu_{B2}^{'}$,}\\ \end{array} \right. \end{equation} \noindent where $\nu_{B1}$ and $\nu_{B2}^{'}$ are break frequencies. According to Formulae (9) and (11), the flux $f$ changes exponentially with time in our model, i.e., for low frequency, \begin{equation} f\propto\left\{ \begin{array}{ll} e^{(5p_1+4) \alpha_2 t/2},&\mbox{~$t<t_{\mbox{\scriptsize peak}}$;}\\ e^{(5p_1+4) \alpha_2 ^{'} t/2},&\mbox{~$t>t_{\mbox{\scriptsize peak}}$,}\\ \end{array} \right. \end{equation} \noindent where $t_{\mbox{\scriptsize peak}}$ denotes the peak time of the light curve. Further, for high frequency, \begin{equation} f\propto\left\{ \begin{array}{ll} e^{(5p_2+9) \alpha_2 t/2},&\mbox{~$t<t_{\mbox{\scriptsize peak}}$;}\\ e^{(5p_2+9)\alpha_2 ^{'} t/2},&\mbox{~$t>t_{\mbox{\scriptsize peak}}$.}\\ \end{array} \right. \end{equation} \section{FITTING RESULTS AND DISCUSSION} Now, we apply the above hourglass model to the multi-wavelength light curves of HST-1 in the M87 jet. Here, the beaming factor of a blob is assumed to be constant. The data we used are plotted in Fig. 1. The units of the radio data are 1 Jy. VLBA 15 GHz radio data from Chang et al. (2010), in which relative uncertainty is assumed to be $5\%$, are plotted as up triangles and down triangles that show the upper limits of fluxes. Very Large Array (VLA) 15 GHz radio data from Harris et al. (2009) and Abramowski et al. (2012) are plotted as squares. The units of the UV data are 1 mJy (Madrid 2009); these data are plotted as circles. For the X-ray data, we use the flux density integrated from 0.2 to 6 keV (Harris et al. 2006, 2009), and the units are $10^{-11}$ erg cm$^{-2}$ s$^{-1}$; these data are plotted as diamonds. The X-ray data after 2005 August 6 were estimated by assuming that the correction factor for the (unknown) effective area (Harris et al. 2006) is the same as that for 2005 August 6. Harris et al. (2006) estimate that the resulting uncertainties are of the order of $15\%$. Note that because of the unknown effective area, the uncertainties of the X-ray fluxes are larger than those in the counts, which may reduce (or smooth) the individual oscillations of the X-ray fluxes. In our model, as the time dependence of the effective area is considered, the resulting uncertainties of the X-ray fluxes may be larger than $15\%$. Based on the shape trends of the radio, UV, and X-ray light curves in HST-1, as shown in Fig. 1, we select a common shape section (from about the year 2003 to 2006.60), which may be the main peak in our fitting area. We used the weighted least-squares method to fit our model to this main peak, with Equation (12) corresponding to the radio and UV light curves and Equation (13) corresponding to the X-ray light curve. Although there is a peak time in our model, its exact position was unknown in the fitting of our model to the light curve. Hence, we first arbitrarily divided the light curve data during the chosen period into two groups to perform the least-squares method, in which the sum of the reduced chi square $\chi^2_{\nu}$ for the two parts is the least. Then we can calculate the corresponding least $\chi^2_{\nu}$ by changing the division of the two groups. All reasonable combinations of the two groups (e.g., the fitting peak time should lie between the two groups) are considered before we obtain the best fit with a minimal $\chi^2_{\nu}$. These best-fit parameters for each light curve are shown in Fig. 1 by blue solid lines. Note that the reduced chi square value depends on uncertainties of the data. Our model can well fit the main peak in the multi-wavelength light curves of HST-1 as shown in Fig. 1. It satisfies the aforementioned first constraint for the multi-wavelength light curves of HST-1: the main peak time is around the year 2005.30, as shown in Fig. 1 and Table 1. The stratified effect of radiation regions in the outer knots of M87 has been verified by Perlman et al. (1999), Marshall et al. (2002), and Perlman $\&$ Wilson (2005). The flattened peak section in the radio light curve can be explained by the greater length of the radio radiation region along the jet axis than the UV and X-ray region. Abramowski et al. (2012) showed that the VLA 22 GHz radio light curve of HST-1 is consistent with the 15 GHz light curve within the error range, which implies that our model can also explain the VLA 22 GHz light curve. In other words, the observation verifies our prediction based on the model fit to the main peak of the VLA 15 GHz light curve. The aforementioned case is a single component; however, a complex structure that contains multiple components may be more reasonable, as in this case, each peak corresponds to a component passing through the hourglass nozzle. We also use the aforementioned method to fit our model to the possible secondary peak of the multi-wavelength light curves. The best fits for our model are plotted in Fig. 1 in red dotted and dot-dashed lines. A similar broken exponential function could fit the TeV peaks in M87 (Abramowski et al. 2012), which may imply a correlation between the TeV flares of M87 and the light curves from radio to X-ray in HST-1. Further, the maximum of the TeV flares of M87 was coincident with the peak of light curves from radio to X-ray in HST-1 observed in 2005 (Cheung et al. 2007); this may imply that the observable TeV flux density was produced through inverse Compton as a blob passed through the hourglass nozzle. The detailed generation process of the TeV flare may be very complicated. The estimates of $p_1$ and $p_2$ in HST-1 require simultaneous broad observational data, which are scarce. Liu \& Shen (2007) found that the averaged spectral index of outer knots in M87 is about 2.36 through model fitting. If we assumed that the spectral index of HST-1 is similar to that of outer knots (i.e., $p_1, p_2\approx 2.36$), we could derive some physical parameters of HST-1 from our model fits, as shown in Table 2. Based on the timescale of $\sim$ 5.6 yr, the local size of the component is smaller than a parsec. The derived $\alpha_2$ agrees well with the derived $\alpha_2^{'}$, and the common value is around 0.12, which implies that the hourglass-shaped layer may be symmetrical with respect to the nozzle. \section{CONCLUSION} We propose a modified two-zone system of TG08: a fast blob passes through a slow hourglass-shaped layer that is connected by two revolving exponential surfaces. Mainly, the emissions from radio to X-ray are radiated by the fast blob. Because of magnetic flux conservation, the brightening and dimming of HST-1 could be explained as adiabatic compression and expansion, respectively, of a blob passing through the outer layer. The observable TeV flux density may be produced through inverse Compton as a blob passes through the nozzle. The VLA 22 GHz radio light curve of HST-1 was used to verify our model. \section{ACKNOWLEDGMENT} We are grateful to Prof. D. E. Harris for his help in the VLA 15 GHz radio light curve of HST-1. We acknowledge the support from the National Natural Science Foundation of China (NSFC) through grants U1231106 and 11273042, the Science and Technology Commission of Shanghai Municipality (12ZR1436100), and the Scientific Research Foundation of the Civil Aviation University of China (09QD15X). This work is partly supported by the China Ministry of Science and Technology under the State Key Development Program for Basic Research (2012CB821800), NSFC (grants 10625314, 11121062, 11173046, 11033007, 10973012, and 11073019), the CAS/SAFEA International Partnership Program for Creative Research Teams, the Strategic Priority Research Program on Space Science, the Chinese Academy of Sciences (Grant No. XDA04060700), and 973 program 2007CB815405. \clearpage
\section{Introduction} Randomness and computation are related to two basic elements of boun\-ded rationality modeling: (a) players cannot generate truly random actions; (b) they cannot implement arbitrarily complex strategies. When rationality is computationally constrained, true randomness can be replaced by pseudo-randomness: one need not generate truly random actions, just actions whose pattern is too complicated for any rationally bounded agent to be able to recognize. Algorithmic randomness is the field of computer science that studies the relations between computability and randomness. It provides the rigorous definitions and analytic tools to address questions involving the tension between the two above elements of bounded rationality. On the other hand, the present paper applies game theory to solve a problem from algorithmic randomness by recasting the problem as an extensive form game. \subsection{Gambling house game} Gambler 0, the cousin of the casino owner, repeatedly bets on red/black. She is only allowed to wager even (positive) integers, i.e., she can bet 4 dollars on black, then 10 on red, then 2 on black, etc. (``betting 4 on black'' means that you gain 4 if black occurred, and lose 4 if red did). Then the \emph{regular} gamblers make their bets, and they are only allowed to wager odd integers. The red/black outcome is not decided by a lottery; rather, the casino owner chooses red or black. He attempts to aid gambler 0 by his choices.\footnote{Following the standard algorithmic randomness formalism, it is assumed that the gamblers do not observe each other's bets or gains, but only observe the casino sequence of reds and blacks. This assumption turns out to be immaterial for our results.} All gamblers start with some finite, not necessary equal, wealth. Going into debt is not allowed, i.e., you cannot bet more than you have. Say that a gambler ``succeeds'' if, along the infinite stream of bets, her wealth tends to infinity. The ``home team'', consisting of the casino owner and gambler 0, wins if and only if gambler 0 succeeds while the other gamblers do not. \par In the case depicted above, gambler 0 uses even integers as wagers, and the others use odd wagers. We will see that the home team can guarantee a win in this case and that they could not had it been the other way round, namely, odd integers for gambler 0 and even integers for the others. In general, let $A$ be the wagers allowed to gambler 0, and $B$ those allowed to the other gamblers ($A,B$ being any pair of subsets of the positive real numbers). We ask for which $A$ and $B$ the home team can \emph{guarantee a win}, in the following sense: gambler 0 can choose a strategy that guarantees that for any strategies that the other gamblers choose the casino can choose a sequence so that the home team wins. \par As a simple example, suppose that apart from gambler 0 there is only one more, regular, gambler (call him gambler 1), and first, let $A=\{2\}$ and $B=\{1\}$. Then the home team cannot guarantee a win: for any pure strategy of gambler 0, let gambler 1 employ the strategy that is a copy of it, only wagering 1 instead of 2. Then the gains (or losses) of 1 are exactly half the gains of 0. Therefore, if gambler 0 succeeds, so does gambler 1. Second, let $B$ be as before, and let $A=\{1,2\}$. Then the home team can guarantee a win as follows. The strategy of gambler 0 is simple: she always bets on red; at the start she wagers two dollars, and keeps doing this as long as she wins. Then, after the first loss, she switches to wagering one dollar ad infinitum. The casino strategy chooses, in a first phase, red every time, until the wealth of 0 exceeds 1's wealth by more than three dollars (note that this is bound to eventually happen, since in the first phase 0 gains two every time, while 1 gains at most one). Then it chooses black once (after which the wealth of 0 still exceeds the wealth of 1, as 0 lost two dollars and 1 gained no more than one dollar). Then comes the second phase, where the casino always chooses the opposite color to what gambler 1 bet on, and chooses red if 1 did not bet anything. This strategy of the casino ensures that, during the second phase, 1 cannot make a non-zero bet more than a finite number of times before he goes broke (and if 1 does go broke, 0 still does not). Hence, whatever strategy he chooses, 1 will not succeed, while 0 will. \subsection{Computability, randomness and unpredictability} \par This casino game is related to the notions of randomness and computability. In the theory of \emph{algorithmic randomness}, a sequence of zeros and ones is called \emph{computably random} if there exists no computable strategy that succeeds against it (thus, the sequence is in some sense unpredictable). For a set $A$ of positive real numbers, a sequence is \emph{$A$-valued random} if there exists no computable gambling strategy (aka ``martingale''), with wagers in $A$, that succeeds against it. Hence we may compare two sets of reals $A$ and $B$, and ask whether $B$-valued randomness implies $A$-valued randomness. In other words, whether for any casino sequence $x$, and a computable $A$-martingale (gambling strategy with wagers in $A$) that succeeds on $x$, there exists a $B$-martingale that succeeds on $x$. And in terms of our casino game, suppose that \emph{every} computable $B$-martingale is employed by some regular gambler. Then the question is whether there exist a computable $A$-martingale and a casino sequence, such that the home team wins. \par There are countably many computable strategies. In many cases this countability, rather than the specific type of admissible strategies, is the essential point in the analysis. This naturally leads to the following setting (Definition \ref{def countable anticipation}): gambler 0's wagers are in $A$, and there are countably many regular players, whose wagers are in $B$. We will say that $A$ \emph{evades} $B$ if the home team can guarantee a win. This is the central notion in our paper. Thus, we are not directly concerned with computability, or other complexity considerations\footnote{Rod Downey (private communication) pointed out that in the case where $A$ and $B$ are recursive sets of rational numbers our analysis applies to the question whether every $B$-valued random is $A$-valued random.}, but rather take a more abstract view: we only require that the number of strategies against which the home team needs to concurrently combat is countable (See also Remark \ref{rem:effective} below). One may consider a few variants of this settings (see below). \par If $A$ does not evade $B$, we say that $B$ \emph{anticipates} $A$. If $B$ contains $A$, then clearly $B$ anticipates $A$, because the strategy of gambler 1 can be an exact copy of 0's strategy (and there is even no need for countably many regular gamblers, one is enough). Therefore, if 0 succeeds, so does 1. Similarly, if $B$ contains a multiple of $A$, i.e., $B \supseteq r A$ for some $r > 0$, then $B$ anticipates $A$: as in the example above, the $B$-strategy can be the same as the $A$-strategy up to the multiple $r$ (i.e., the wagers are multiplied by $r$), and therefore the gains are the same, multiplied by $r$. For example, the even integers anticipate the odd integers (or the whole set of integers, for that matter). \par Thus, $B$ containing a multiple of $A$ is a sufficient condition for $B$ to anticipate $A$. Is it also a necessary condition? \cite{chalcraft12} showed that if $A$ and $B$ are finite, then it is necessary. They asked whether this characterization extends to infinite $A$ and $B$ (note that their framework is that of algorithmic randomness, i.e., it only allows for computable martingales). In other words: \begin{itemize} \item[$(*)$] Does $B$ not containing a multiple of $A$ imply that $A$ evades $B$? \end{itemize} \par \subsection{Our contribution} Theorem~\ref{thm:suffic} implies a negative answer to this question.\footnote{We restrict our attention to closed sets, since every set anticipates its closure (Lemma~\ref{lem close}).} Thus, for example, the segment $[0,1]$ anticipates the set $\mathbb R_+$, although it does not contain any multiple of it. Still, although $(*)$ does not hold in general, we will see classes of pairs $A,B$ where $(*)$ does hold. Theorem~\ref{theorem bounded} says that if $A$ is bounded, and $B\setminus \{0\}$ is bounded away from 0, then $(*)$ holds. Thus, for example, the set $\{1,\pi\}$ evades the positive integers. Theorem \ref{theorem well ordered} says that if $B$ is well-ordered, then $(*)$ holds. Thus, for example, the set of even integers evades the odds. \par Now let us consider the following version of our casino game, whose description is more like that of a ``classic'' repeated game. At each stage of the repeated game, first gambler 0 makes a bet; then, after observing it, the (team of) regular gamblers each make their bets; and then the casino chooses a red/black outcome. All of our results apply to this version as well. \par The difference between our definition of evasion and the description of this game is that in the former setting, the regular gamblers knew the \emph{strategy} of gambler 0 (i.e., including her future actions), gambler 0 did not observe their past actions, and the casino knew the strategies of all gamblers. One may consider many such variations to the settings, and our results are robust to all of them -- whether gambler 0 observes anything or not, whether the regular gamblers know her future bets or not, and whether the casino knows the future bets or not. \par Previous work on martingales with restricted wagers (\cite{bienvenu12}, \cite{teutsch13}, \cite{peretz13}) employed various solutions in various specific situations. The present paper proposes a systematic treatment that applies to most of the previously studied situations, as well as many others. Our proofs are elementary and constructive. Also, the proposed construction is recursive: it does not rely on what the strategies do in the future; therefore we believe our method can be useful in many frameworks where computational or other constraints are present. \par The rest of the paper is organized as follows. Section \ref{sec:Martingales} presents the definitions, results, and a few examples, as well as a discussion of related previous work. The next two sections contain proofs. Section \ref{sec:Discussion} points to open problems and further directions. \section{Definitions and Results}\label{sec:Martingales} A \emph{martingale} is a gambling strategy that bets on bits of a binary sequence. Formally, it is a function $M:\{\head,\tail\}^{*}\to\mathbb R$ that satisfies \[ M(\sigma)=\frac{M(\sigma\head)+M(\sigma\tail)}{2}, \] for every string $\sigma\in\{\head,\tail\}^{*}$. The \emph{increment} of $M$ at $\sigma\in\{\head,\tail\}^{*}$ is defined as \[ M'(\sigma)= M(\sigma\head)-M(\sigma). \] For $A\subset\mathbb R_+$, we say that $M$ is an \emph{$A$-martingale} if $|M'(\sigma)|\in A$, for every $\sigma\in\{\head,\tail\}^{*}$. The empty string is denoted $\varepsilon$ and $M(\varepsilon)$ is called the \emph{initial value} of $M$. Note that a martingale is determined by its initial value and its increments. The initial sub-string of length $n$ of a binary sequence, $X\in\{\head,\tail\}^{\infty}$, is denoted $X\restriction n$. A martingale $M$ \emph{succeeds} on $X$, if $\lim_{n\to\infty}M(X\restriction n)=\infty$ and $M(X\restriction n)\geq |M'(X\restriction n)|$, for every $n$. The latter condition asserts that $M$ doesn't bet on money it doesn't have. The set of sequences on which $M$ succeeds is denoted $\Succ(M)$. A martingale $N$ \emph{dominates} $M$ if $\Succ(N)\supseteq\Succ(M)$, and a set of martingales $\mathcal N$ dominates $M$ if $\bigcup_{N\in\mathcal N}\Succ(N)\supseteq\Succ(M)$. The following are non-standard definitions. \begin{definition}\label{def single anticipation} A set $B\subseteq\mathbb R_+$ \emph{singly anticipates} a set $A\subseteq\mathbb R_+$, if every $A$-martingale is dominated by some $B$-martingale. If $A$ singly anticipates $B$ and $B$ singly anticipates $A$, we say that $A$ and $B$ are \emph{strongly equivalent}. \end{definition} \begin{definition}\label{def countable anticipation} A set $B\subseteq\mathbb R_+$ \emph{countably anticipates} (\emph{anticipates}, for short) a set $A\subseteq\mathbb R_+$, if every $A$-martingale is dominated by a countable set of $B$-martingales. If $A$ anticipates $B$ and $B$ anticipates $A$, we say that $A$ and $B$ are \emph{(weakly) equivalent}. If $B$ does not anticipate $A$ we say that $A$ \emph{evades} $B$. \end{definition} Note that both ``singly anticipates'' and ``anticipates'' are reflexive and transitive relations (namely, they are preorders),\footnote{The evasion relation is anti-reflexive and is not transitive.} and that single anticipation implies anticipation. Also, if $A\subseteq B$ then $B$ singly anticipates $A$. \begin{remark}\label{rem:effective} The motivation for the definition of countable anticipation comes from the study of algorithmic randomness (see, e.g., \cite{downey-online}), where it is natural to consider the countable set of all $B$-martingales that are computable relative to $M$ (see \cite{peretz13}). Formally, a set $B\subseteq\mathbb R_+$ \emph{effectively anticipates} a set $A\subseteq\mathbb R_+$, if for every $A$-martingale $M$ and every sequence $X\in\Succ(M)$, there is a $B$-martingale, computable relative to $M$, that succeeds on $X$. More generally, one could define in the same fashion a ``$\mathcal C$-anticipation'' relation with respect to any complexity class $\mathcal C$. Our main focus will be on countable anticipation, and specifically on sets $A$ and $B$ such that $B$ does not countably anticipate $A$; and therefore $B$ does not $\mathcal C$-anticipate $A$ for any complexity class $\mathcal C$. When we present cases in which anticipation does hold, the dominating martingales will usually be fairly simple relative to the dominated martingales. We do not presume to rigorously address the computational complexity of those reductions, though. \end{remark} The topological closure of a set $A\subseteq\mathbb R_+$ is denoted $\bar A$. The following lemma says that we can restrict our attention to closed subsets of $\mathbb R_+$. \begin{lemma}\label{lem close} Every subset of $\mathbb R_+$ is strongly equivalent to its closure. \end{lemma} \begin{proof}{} Let $A\subset\mathbb R_+$ and let $M$ be an $\bar A$-martingale. Define an $A$-martingale $S$ by \begin{align*} S(\varepsilon)&=M(\varepsilon)+2,\\ S'(\sigma)&\in A\cap (M'(\sigma)-2^{-|\sigma|},M'(\sigma)+2^{-|\sigma|} ), \end{align*} where $|\sigma|$ is the length of $\sigma$. Clearly, $S(\sigma) > M(\sigma)$, for every $\sigma\in\{\head,\tail\}^*$; therefore $\Succ(S)=\Succ(M)$. \end{proof} Another simple observation is that for every $A\subseteq\mathbb R_+$ and $r>0$, $A$ is strongly equivalent to $rA\:=\{ra: a\in A\}$. This observation leads to the next definition. \begin{definition} Let $A,B\subseteq\mathbb R_+$. We say that $A$ and $B$ are \emph{proportional}, if there exists $r>0$ such that {$r A=B$}. If we only require that $rA\subseteq \bar B$, then $A$ is proportional to a subset of the closure of $B$. In that case we say that $A$ \emph{scales} into $B$. \end{definition} From the above and the fact that $A\subseteq B$ implies that $B$ singly anticipates $A$, we have the following lemma. \begin{lemma}\label{lem similarity} If $A$ scales into $B$, then $B$ anticipates $A$, for every $A,B\subseteq\mathbb R_+$. \end{lemma} The next two theorems provide conditions under which the converse of Lemma~\ref{lem similarity} also holds. \begin{theorem}\label{theorem bounded} For every $A,B\subseteq\mathbb R_+$, if $\sup A<\infty$ and $0\not\in \overline{B\setminus\{0\}}$, then $B$ anticipates $A$ only if $A$ scales into $B$. \end{theorem} \begin{theorem}\label{theorem well ordered} For every $A,B\subseteq\mathbb R_+$, if $B$ is well ordered, namely, $\forall x\in\mathbb R_+\ x\not\in \overline{B\setminus[0,x]}$, then $B$ anticipates $A$ only if $A$ scales into $B$. \end{theorem} \citet{chalcraft12} studied effective anticipation between finite subsets of $\R$. They showed that (effective) anticipation is equivalent to containing a proportional set on the domain of finite sets. They further asked whether their result extends to infinite sets. In particular, they asked if $\mathbb Z_+$ anticipates the set $V=\{0\}\cup[1,\infty)$. \cite{peretz13} showed that the set $\{1+\frac{1}{n}\}_{n=1}^\infty\subset V$ evades (i.e., is not anticipated by) $\mathbb Z_+$, and the set $\{\frac{1}{n}\}_{n=1}^\infty$ evades $V$. All of the above results follow immediately from Theorem \ref{theorem bounded}. Furthermore, any set that contains two $\mathbb Q$-linearly independent numbers (e.g., $\{1,\pi\}$) evades $\mathbb Z_+$. In light of Lemma \ref{lem close}, one may rephrase the above question of \citet{chalcraft12} and ask whether, for infinite sets, anticipation is equivalent to scaling. The answer is still negative (Theorem \ref{thm:suffic}), although Theorems \ref{theorem bounded} and \ref{theorem well ordered} above gave some conditions under which the equivalence does hold. Theorem \ref{theorem well ordered} says that if, for example, $A = \mathbb Z_+$ and $B$ is a subset of $\mathbb Z_+$ whose density is zero, then $A$ evades $B$. This is because $B$ is well ordered, and $A$ does not scale into $B$ (by the zero density). Another example is when $B$ is the set of all odd integers where, again, $A$ evades $B$. Note that the set of even integers is proportional to $\mathbb Z_+$ and hence is equivalent to $\mathbb Z_+$. Therefore the even integers evade the odd integers, but not vice versa. Furthermore, we can take any subset of $\mathbb Z_+$ that does not contain an ``ideal'' (i.e., all the integer multiples of some number, namely, a set proportional to $\mathbb Z_+$). If $B$ is of the form $B = \mathbb Z_+ \setminus \event{n \cdot \phi(n)}_{n=1}^\infty$ for some function $\phi : \mathbb N \to \mathbb N$, then $\mathbb Z_+$ evades $B$ even when, for example, the function $\phi$ grows very rapidly. In particular, the density of $B$ could equal one. The previous theorems gave some necessary conditions for anticipation. The next theorem gives a sufficient condition. \begin{theorem}\label{thm:suffic} Let $A,B\subset\R_+$\,. For $x>0$, let $$ P(x)= \{ t \geq 0 : t\cdot (A\cap [0,x]) \subseteq \bar{B}\cup \event{0} \,\}. $$ For any $M \geq 0$ \,let $q_M (x) = \max (P(x) \cap [0,M])$. If for some $M$, $\int_0^\infty q_M(x)dx = \infty$, then $B$ singly anticipates $A$. \end{theorem} This is equivalent to the following seemingly stronger theorem. \newtheorem*{nt}{Theorem \ref{thm:suffic}$^*$} \begin{nt}\label{lemma f} Let $A,B\subset\R_+$. Suppose there is a non-increasing function $f\colon\R_+\to\R_+$, such that \begin{enumerate} \item $\bar{B}\supset f(x)\left(A\cap[0,x]\right)$, for every $x\in\R_+$; and \item $\int_0^\infty f(x)\,\mathrm d x=\infty$. \end{enumerate} Then, $B$ singly anticipates $A$. \end{nt} \begin{proof}[Proof of Theorem \ref{thm:suffic}$^*$] Let $M$ be an $A$-martingale. By Lemma~\ref{lem close} we may assume that $B$ is closed. Define a $B$-martingale, $S$, by \begin{figure} \centering \includegraphics[width=\textwidth,keepaspectratio=true,trim=100pt 103pt 32pt 79pt,clip=true]{thm8.eps} \caption{Inequality \eqref{eq integral}} \end{figure} \begin{align*} S(\varepsilon) &= f(0)M(\varepsilon),\\ S'(\sigma)&=f(M(\sigma))M'(\sigma). \end{align*} Let $X\in\{\head,\tail\}^{\mathbb N}$ such that $X \in \Succ(M)$, and let $n\in\mathbb N$. Since $f$ is non-increasing, \begin{equation}\label{eq integral} S(X\restriction n+1)-S(X\restriction n)\geq\int_{M(X\restriction n)}^{M(X\restriction n+1)}f(x)\,\mathrm{d}x. \end{equation} It follows by induction that \[ S(X\restriction n)\geq\int_{0}^{M(X\restriction n)}f(x)\,\mathrm{d}x, \] which concludes the proof, since $M(X\restriction n) \to \infty$ and $\int_{0}^{\infty}f(x)\,\mathrm{d}x=\infty$. \end{proof} If, for example, $A$ scales into $B$, namely, $r A \subseteq \bar{B}$, the function $f$ in Theorem \ref{thm:suffic}$^*$ can be taken to be simply $f(x)=r$. Also note that when $A$ and $B$ are finite, such a function $f$ as in the theorem exists if and only if $A$ scales into $B$. The theorem tells us, for example, that although $\R_+$ does not scale into the interval $[0,1]$, these two sets are (strongly) equivalent: to see that $[0,1]$ singly anticipates $\mathbb R_+$, apply Theorem \ref{thm:suffic}$^*$ with $f(x)=\min\{\frac 1 x ,1 \}$. Another example is the set $A = \{2^n\}_{n=-\infty}^{+\infty}$ being (strongly) equivalent to $B = \{2^n\}_{n=-\infty}^0$, although $A$ does not scale into $B$. To see this, apply Theorem \ref{thm:suffic}$^*$ with $f(x)=\min\{1/ 2^{\lfloor\log_2 x\rfloor} ,1 \}$. We previously saw, by Theorem \ref{theorem well ordered}, that $A = \mathbb Z_+ $ evades $B = \mathbb Z_+ \setminus \event{n \cdot \phi(n)}_{n=1}^\infty$. Now look at this example in the context of Theorem \ref{thm:suffic}, to see that there is no contradiction. Suppose $\phi(n)$ is increasing, and $\phi(n) \to \infty$. Then for any $x > 0 $, the set $P(x)$ in the theorem is unbounded, i.e., $\sup P(x)=\infty$. Nevertheless, fix any choice of $M$, and note that for every $x$ large enough, $P(x)$ does not contain any nonzero number smaller than $M$, i.e., $P(x) \cap [0,M] = \event{0}$. Thus, for any choice of $M$, $q_M(x) = 0$ for every $x$ large enough. In particular, $\int_0^\infty q_M(x)dx < \infty$. \section{Proof of Theorem~\ref{theorem bounded}} Throughout this section $A,B\subset\mathbb R_+$ are two sets satisfying \begin{align*} \sup A &<\infty,\\ 0&\not\in \overline {B\setminus\{0\}},\\ A&\text{ does not scale into } B. \end{align*} We must show $A$ (countably) evades $B$. Since $A$ does not scale into $B$, one thing that $B$-martingales cannot do in general is to mimic $A$-martingales, not even up to a constant ratio. We use this idea in order to construct a sequence of heads and tails that will separate between the two types of martingales. \subsection{Ratio minimization} Let $N$ and $M$ be martingales with $N$ non-negative. We say that $x\in\{\head,\tail\}$ is the \emph{$N/M$-ratio-minimizing outcome} at $\sigma\in\{\head,\tail\}^{<\infty}$ (assuming $M(\sigma)> 0$) if either \begin{enumerate} \item $\frac{N(\sigma)}{M(\sigma)}> \frac{N(\sigma x)}{M(\sigma x)}$, or \item $\frac{N(\sigma)}{M(\sigma)}= \frac{N(\sigma x)}{M(\sigma x)}$ and $M(\sigma x)> M(\sigma)$, or \item $M'(\sigma)=N'(\sigma)= 0$ and $x=\head$. \end{enumerate} In words: our first priority is to make the ratio $N/M$ decrease; if this is impossible (i.e., the increments of $N$ and $M$ are proportional to their value at $\sigma$, and so $N/M$ doesn't change), then we want $M$ to increase so as to insure that $M(\sigma x)>0$; if that is impossible as well (i.e., both increments are 0), we set $x$ to be $\head$, for completeness of definition only. Our definition extends to finite/infinite extensions of $\sigma$ by saying that $X$ is the length $|X|$ (possibly $|X|=\infty$) \emph{$N/M$-ratio-minimizing extension} of $\sigma$, if $X_{t+1}$ is the $N/M$-ratio-minimizing outcome at $X\restriction t$, for every $|\sigma|\leq t < |X|$. For any such $N$, $M$, and $\sigma$ the infinite $N/M$-ratio-minimizing extension of $\sigma$, $X$, makes the ratio $N(X\restriction t)/M(X\restriction t)$ monotonically converging to a limit $L\in \mathbb R_+$, as $t\to\infty$. The next lemma will help us argue that the ratio between the increments ${N'(X\restriction t)}/M'(X\restriction t)$ also converges to $L$ in a certain sense. \begin{lemma}[Discrete l'H\^opital rule]\label{lem:lhopital} Let $(a_n)_{n=1}^\infty$ and $(b_n)_{n=1}^\infty$ be sequences of real numbers. Assume that $a_n> 0$, for every $n$. If $b_n/{a_n}$ monotonically converges to a limit $L\in\mathbb R$, and $\sup\{\frac 1 n \sum_{k=1}^n |a_{k+1}-a_k|\}<\infty$, then \begin{equation*} \lim_{n\to\infty} \frac 1 n \sum_{k=1}^{n} |(b_{k+1}-b_k) - L(a_{k+1}-a_k)|=0. \end{equation*} \end{lemma} \begin{proof} Since $\frac 1 n \sum_{k=1}^n |a_{k+1}-a_k|$ is bounded, $\frac 1 n \sum_{k=1}^n (a_{k+1}-a_k)$ is bounded, too; therefore \begin{equation}\label{eq weak} \lim_{n\to\infty} \frac 1 n \sum_{k=1}^{n} (b_{k+1}-b_k) - L(a_{k+1}-a_k)=0. \end{equation} It remains to prove that \begin{equation}\label{eq strong'} \lim_{n\to\infty} \frac 1 n \sum_{k=1}^{n} \left[b'_k-L a'_k\right]_+=0, \end{equation} where $a'_k=a_{k+1}-a_k$ and similarly $b'_k=b_{k+1}-b_k$. We may assume w.l.o.g. that $\frac{b_n}{a_n}\searrow L$ (otherwise consider the sequence $(-b_n)_{n=1}^\infty$). Namely, $\frac{b_k}{a_k}\geq \frac{b_{k+1}}{a_{k+1}}$, which implies that $b'_k\leq \frac{b_k}{a_k}a'_k$; hence \[ \left[b'_k-L a'_k\right]_+ \leq \left[\left(\frac{b_k}{a_k}-L\right)a'_k\right]_+\leq \left(\frac{b_k}{a_k}-L\right)|a'_k|. \] Now \eqref{eq strong'} follows since $\frac{b_k}{a_k}$ converges to $L$ and $\frac 1 n \sum_{k=1}^n |a'_k|$ is bounded. \end{proof} \begin{cor}\label{cor ratio} Let $M$ and $N$ be a pair of martingales and $\sigma\in \{\head,\tail\}^{*}$. Assume that $N$ is non-negative, $M(\sigma)>0$, and $M'$ is bounded.\footnote{The assumption that $M'$ is bounded can be relaxed by assuming only that $\frac{1}{N}\sum_{t=0}^{N-1}|M'(X\restriction t)|$ is bounded.} Let $X$ be the infinite $N/M$-ratio-minimizing extension of $\sigma$ and $L=\lim\limits_{t\to\infty}\frac{N(X\restriction t)}{M(X\restriction t)}$. For every $\epsilon>0$ the set \[ \left\{t:\left\vert N'(X\restriction t) -L\cdot{M'(X\restriction t)} \right\vert > \epsilon \right\} \] has zero density. \end{cor} \begin{proof} Note that \begin{multline*} \left\vert N'(X\restriction t) -L\cdot M'(X\restriction t)\right\vert =\\ \left\vert (N(X\restriction t+1)-N(X\restriction t)) -L\cdot(N(X\restriction t+1)-N(X\restriction t))\right\vert. \end{multline*} By Lemma~\ref{lem:lhopital}, we have \[ \lim_{n\to\infty} \frac 1 n \sum_{k=1}^{n} |N'(X\restriction t) -L\cdot M'(X\restriction t)|=0, \] which implies that the density of the set \[ \left\{t:\left\vert N'(X\restriction t) -L\cdot{M'(X\restriction t)}\right\vert > \epsilon \right\} \] is zero, for every $\epsilon>0$. \end{proof} The next step is to construct the history-independent $A$-martingale prescribed by Theorem~\ref{theorem bounded}. We formalize the properties of this martingale in the following lemma. \begin{lemma}\label{lemma 0 ratio} Let $A,B\subset\mathbb R_+$. Suppose that $\sup A< \infty$ and $A$ does not scale into $B$; then there exists a history-independent $A$-martingale with positive initial value, $M$, such that for every non-negative $B$-martingale, $N$, and every $\sigma\in\{\head,\tail\}^*$ such that $M(\sigma)>0$, the infinite $N/M$-ratio-minimizing extension of $\sigma$, $X$, satisfies \[ \lim_{t\to\infty}\frac{N(X\restriction t)}{M(X\restriction t)}=0. \] \end{lemma} Note that the $A$-martingale, $M$, does not depend on $N$ or $\sigma$, so the same $M$ can be used against any $N$ at any $\sigma$ that leaves $M(\sigma)$ positive. \begin{proof}[Proof of Lemma \ref{lemma 0 ratio}] Let $\{a_n\}_{n=0}^\infty$ be a countable dense subset of $A\setminus\{0\}$. For every positive integer $t$, let $n(t)\in\mathbb Z_+$ be the largest integer such that $2^{n(t)}$ divides $t$. Define $x_t\:=a_{n(t)}$. The sequence $\{x_t\}_{t=1}^\infty$ has the property that the set \begin{equation}\label{eq positive density} \{t:|x_t-a|<\epsilon\} \end{equation} has positive density, for every $\epsilon>0$ and $a\in A$. Let $M$ be a history-independent martingale whose increment at time $t$ is $x_t$, for every $t\in\mathbb Z_+$ (with an arbitrary positive initial value). Let $N$ be an arbitrary non-negative $B$-martingale. Suppose that $M(\sigma)> 0$ and let $X$ be the infinite $N/M$-ratio-minimizing extension of $\sigma$ and let $L=\lim_{t\to\infty}\frac{N(X\restriction t)}{M(\restriction t)}$. Corollary \ref{cor ratio} and \eqref{eq positive density} guarantee that $L\cdot A\subset \bar B$. By the assumption that $A$ does not scale into $B$, we conclude that $L=0$. \end{proof} \subsection{The casino sequence} In the rest of this section we assume that $\sup A=\inf (B\setminus \{0\})=1$. This is w.l.o.g. since proportional sets are (strongly) equivalent. We begin with an informal description of the casino sequence. Lemma \ref{lemma 0 ratio} provides a history-independent $A$-martingale, $M$, that can be used against any $B$-martingale. Given a sequence of $B$-martingales, $N_1,N_2,\ldots$, we start off by ratio-minimizing against $N_1$. When $M$ becomes greater than $N_1$, we proceed to the next stage. We want to make sure that $N_1$ no longer makes any gains. This is done by playing adversarial to $N_1$ whenever he wagers a positive amount. At times when $N_1$ wagers nothing (i.e., $N_1'=0$), we are free to choose either $\head$ or $\tail$ without risking our primary goal. At those times we turn to ratio-minimizing against $N_2$, while always considering the goal of keeping $N_1$ from making gains a higher priority. Since $\inf \left\{B\setminus\{0\}\right\}=1>0$, it is guaranteed that at some point we will no longer need to concern $N_1$, and hence, at some even further point, $M$ will become greater than $N_1+N_2$. The process continues recursively, where at each stage our highest priority is to prevent $N_1$ from making gains, then $N_2$, $N_3$, and so on until $N_k$; and if none of $N_1,\ldots,N_k$ wagers any positive amount, we ratio-minimize against $N_{k+1}$. When a positive wager of some $N_i$, $i\in\{1,\ldots k\}$, is answered with an adversarial outcome, a new index $k'$ must be calculated, so that $M$ is sufficient to keep $N_1,\ldots,N_{k'}$ from making gains. That is, $M>N_1+\cdots+N_{k'}$. An inductive argument shows that for every fixed $k$, there is a point in time beyond which none of $N_1,\ldots,N_k$ will ever wager a positive amount; therefore, at some even further point, $M$ becomes greater than $N_1+\cdots+N_{k+1}$; hence the inductive step. The above explains how the value of each $N_i$ converges to some $L_i \in \mathbb R_+$, and the limit inferior of the value of $M$ is at least $\sum_{i=1}^{\infty}L_i$. In order to make sure that $M$ goes to infinity we include, among the $N_i$s, infinitely many martingales of constant value $1$. We turn now to a formal description. As mentioned above, we assume without loss of generality that $\sup A=\inf \{B\setminus \{0\}\}=1$. We additionally assume that $0\in B$, and so we can convert arbitrary $B$-martingales to non-negative ones by making them stop betting at the moment they go bankrupt. Let $M$ be a history-independent $A$-martingale provided by Lemma \ref{lemma 0 ratio}. Let $N_1,N_2,\ldots$ be a sequence of non-negative $B$-martingales. Assume without loss of generality that infinitely many of the $N_i$s are the constant $1$ martingale. We define a sequence $X\in\{\head,\tail\}^\infty$ recursively. Assume $X\restriction t$ is already defined. First we introduce some notation. Denote the value of $M$ at time $t$ by $m(t)=M(X\restriction t)$, and similarly $n_i(t)=N_i(X\restriction t)$, for every $i\in\mathbb N$. Let \begin{align*} S_i(t)&=\sum_{j=1}^in_j(t),\\ k(t)&=\max\{i:S_i(t) < m(t)\},\text{ and}\\ S(t)&=S_{ k(t) }(t). \end{align*} Note that the maximum is well defined, since the $n_i(t)$ include infinitely many 1s. We are now ready to define $X_{t+1}$. We distinguish between two cases: Case I: there exists $1\leq j\leq k(t)$ such that $N_j'(X\restriction t)\neq 0$; Case II: $N_1'(X\restriction t)=\cdots=N_{k(t)}'(X\restriction t)= 0$. In Case I, let $i=\min\{j:N_j'(X\restriction t)\neq 0\}$ and define \[ X_{t+1} = \begin{cases} \tail &\text {if $N_{i}'(X\restriction t)>0$,}\\ \head &\text {if $N_{i}'(X\restriction t)<0$.} \end{cases} \] In Case II, $X_{t+1}$ is the $\frac{N_{k(t)+1}}{M-S(t)}$-ratio-minimizing outcome at $X\restriction t$. Explicitly, \[ X_{t+1} = \begin{cases} \tail &\text {if $\frac{N_{k(t)+1}'(X\restriction t)}{M'(X\restriction t)}>\frac{n_{k(t)+1}}{m(t)-S(t)}$,}\\ \head &\text {if $\frac{N_{k(t)+1}'(X\restriction t)}{M'(X\restriction t)}\leq\frac{n_{k(t)+1}}{m(t)-S(t)}$.} \end{cases} \] Consider the tuple $\alpha(t)=(\lfloor n_1(t)\rfloor,\ldots,\lfloor n_{k(t)}(t)\rfloor)$. In Case I, $\alpha(t+1)$ is strictly less that $\alpha(t)$ according to the lexicographic order. In Case II, $\alpha(t)$ is a prefix of $\alpha(t+1)$, and so under a convention in which a prefix of a tuple is greater than that tuple, we have that $\{\alpha(t)\}_{t=1}^\infty$ is a non-increasing sequence.\footnote{Alternatively, one can use the standard lexicographic order where $\alpha(t)$ is appended with an infinite sequence of $\infty$ elements.} Let $k=\liminf\limits_{t\to\infty} k(t)$. It follows that from some point in time, $\alpha$ consists of at least $k$ elements; therefore the first $k$ elements of $\alpha$ must stabilize at some further point in time. Namely, for $t$ large enough we have $\lfloor n_i(t)\rfloor=\lim\limits_{t'\to\infty}\lfloor n_i(t')\rfloor <\infty$, for every $i\leq k$. Since the increments of $n_i(t)$ are bounded below by 1, $n_1(t),\ldots,n_k(t)$ stabilize, too. Also, since $m(t)> S(t)$, we have $\liminf\limits_{t\to\infty}m(t)\geq \lim_{t\to\infty} S_i(t)$, for every $i\leq \liminf\limits_{t\to\infty} k(t)$. Since there are infinitely many $i$'s for which $n_i(t)$ is constantly $1$, the proof of Theorem \ref{theorem bounded} is concluded by showing that $\liminf\limits_{t\to\infty} k(t)=\infty$. Assume by negation that $\liminf\limits_{t\to\infty} k(t)=k<\infty$. There is a time $T_0$ such that $n_k(t)=\lim\limits_{t'\to\infty}n_k(t')$ and $k(t)\geq k$, for every $t>T_0$. There cannot be a $t>T_0$, for which $k(t)>k$ and $k(t+1)=k$. That would mean a Case I transition from time $t$ to $t+1$ and we would have $n_i(t+1)<n_i(t)$, for some $i \leq k$. It follows that $k(t)=k$, for every $t>T_0$. From time $T_0$ ratio-minimization against $n_{k+1}$ takes place. Let $l=\linebreak\liminf\limits_{t\to\infty}n_{k+1}(t)$. If $l=0$, then $n_{k+1}(t)<1$, for some $t>T_0$; at this point $n_{k+1}$ stabilizes (otherwise $N_{k+1}$ would go bankrupt); therefore $n_{k+1}(t)=0$; therefore $k(t)>k$, which is not possible. If $l>0$, then by Lemma~\ref{lemma 0 ratio}, there must be some time $t>T_0$ in which $m(t)>n_{k+1}(t)+S_k(T_0)=n_{k+1}(t)+S(t)$, which contradicts the definition of $S(t)$. \section{Proof of Theorem~\ref{theorem well ordered}} To show that if $B$ is well ordered and $A$ does not scale into $B$, then $A$ evades $B$, we construct an $A$-martingale $M$, s.t. for any $B$-martingales $N_1,N_2,\ldots$ we construct a sequence $X$ on which $M$ succeeds, while every $N_i$ does not. We begin with a rough outline of the proof ideas. $M$ always bets on ``heads.'' Before tackling every $N_i$, we first gain some money and ``put it aside.'' Then we ratio-minimize against $N_i$. It will eventually make $M$ sufficiently richer than $N_i$, so that we can declare $N_i$ to be ``fragile'' now. This means that from now on the casino can make $N_i$ lose whenever it is ``active'' (i.e., makes a non-zero bet), since $M$ can afford losses until $N_i$ is bankrupt. When $N_i$ is not active, we can start tackling $N_{i+1}$, while constantly making sure that we have enough money kept aside for containing the fragile opponents. An important point is to show that once some $N_i$ becomes fragile, it remains fragile unless a lower-index martingale becomes active. \par Let $(a_n)$ be a sequence that is dense within $A\setminus \event{0}$, and such that each number in the sequence appears infinitely many times. For example, given a dense sequence $(x_n)$ in $A\setminus \event{0}$, the sequence \[ x_1;\ x_1,x_2;\ x_1,x_2,x_3;\ x_1,x_2,x_3,x_4; \ldots \] can be used. \par Since multiplying $A$ or $B$ by a positive constant does not make a difference, we may assume w.l.o.g. that $a_1=1$, and that $\inf (B\setminus \event{0}) =1$ ($B$ is well ordered; hence in particular $B\setminus \event{0}$ is bounded away from $0$). \par We construct $M$ and $X$ as follows. Denote $m(t) = M(X\restriction t)$, and similarly $m'(t) = M'(X\restriction t)$. Take integers $$f(t,k) \geq \max \event{m(t),\abs{(a_{k+1}-a_k)/a_k}}.$$ We take $M(\ve) = a_1$. For the first $f(0,1)$ stages, $M' = a_1$. Then, after stage $t = f(0,1)$, $M' = a_2$ for the next $f(t,2)$ periods, and after stage $t' = t + f(t,2)$, $M' = a_3$ for $f(t',3)$ stages, and so on. But this goes on only as long as no ``tails'' appears. Whenever a ``tails'' appears, namely, at a stage $t$ where $X_t = \tail$, we revert to playing from the beginning of the sequence, i.e., $M' = a_1$ for the next $f(t,1)$ stages (or until another ``tails'' appears), then $M' = a_2$, etc. \par For $t \geq 0$ we define the function $\g(t)$, which is similar to $m'(t)$, but modifies sudden increases of $m'$ into more gradual ones. \begin{figure} \centering \includegraphics[width=\textwidth,keepaspectratio=true,trim=56pt 165pt 90pt 198pt,clip=true]{thm7.eps} \caption{$\g$ and the wagers} \end{figure} At the beginning of a block of stages where $a_k$ is wagered (i.e., where $m'(\cdot)=a_k$), $\g$ equals $a_k$. If $a_{k+1} \leq a_k$, then $\g$ remains $a_k$ throughout this block. Otherwise, it linearly increases until reaching $a_{k+1}$ exactly at the beginning of the next block. I.e., suppose $m'(t) = a_k$, and let $s \leq t$ be the beginning of the block (of length $f(s,k)$) of $a_k$ wagers. If $a_{k+1} \leq a_k$ then $\g(t) = a_k$. Otherwise, $$\g(t) = \frac{(s+ f(s,k) - t) \,a_k + (t-s) a_{k+1}}{f(s,k)}.$$ \begin{align*} \text{Note:\qquad} &(i)\,\g(t) \geq m'(t),\\ &(ii) \text{\,If } X_{t+1} = \tail \text{ \,then } \g(t+1) = a_1 = 1,\\ &(iii) \text{\,If } X_{t+1} = \head \text{ \,then } \g(t+1)- \g(t) \leq m'(t). \end{align*} The last one follows from $m'(t) = a_k$, \,$\g(t+1)- \g(t) \leq (a_{k+1} - a_k) / f(s,k)$, and $f(s,k) \geq \abs{(a_{k+1}-a_k)/a_k}$\,, by the definition of $f$. \par Let $N_1,N_2,\ldots$ be $B$-martingales (and assume these martingales never bet on money that they do not have). Denote $n_i(t) = N_i(X\restriction t)$. To define the sequence $X$, denote $\nu_k(t) = k + n_1(t)+\ldots+n_k(t)$, and let $p=p(t) \geq 0$ be the largest integer such that $ m(t) - (\g(t) - 1) > \nu_p(t). $ $N_1,\ldots,N_{p(t)}$ are the ``fragile'' martingales at time $t$. Define $$\mu(t) = m(t)- (\nu_p(t) + 1)$$ and consider two cases. (i) If there exists some index $1 \leq j \leq p(t)$ s.t. $n'_j(t) \neq 0$, let $i$ be the smallest such index, and $X_{t+1}$ is chosen adversely to $n'_i$. (ii) Otherwise, $X_{t+1}$ is chosen by $\mu / N_{p+1}$-ratio-minimizing, i.e., if $\mu(t) > 0 $ and $n'_{p+1}(t) / m'(t) > n_{p+1}(t) / \mu(t)$ then $X_{t+1}= \tail$, and otherwise $X_{t+1}= \head$\,. \par We now show that these $M$ and $X$ indeed work. \begin{lemma}\label{lem:fragile} For any $t$, if $p(t) \geq i $ and $n'_j(t)=0$ for every $j < i$, then $p(t+1) \geq i$\,. \end{lemma} \def \lside{\mathcal{L}} \begin{proof}[Proof of Lemma \ref{lem:fragile}] We prove the following equivalent claim:\\(I) If $i \leq p(t)$ is the smallest index such that $n'_i(t) \neq 0$, then $i \leq p(t+1)$. (II) If $n'_j(t) = 0$ for any $j \leq p(t)$, then $p(t) \leq p(t+1)$. \par In case (II), denote $i=p(t)$. Then in both cases $$m(t) - (\g(t) - 1) > \nu_i(t)$$ is known. Let $\lside(t)$ designate the LHS of this inequality. We need to show that $\lside(t+1) > \nu_i(t+1)$. Note that for any $j< i$, $n_j(t+1) = n_j(t)$, and that if $X_{t+1} = \head$ \,then $\lside(t+1) \geq \lside(t)$, since $m(t+1) = m(t)+ m'(t)$ and $\g(t+1) \leq \g(t) + m'(t)$. \par (I) In this case the casino makes $i$ lose, hence $n_i(t+1) \leq n_i(t)-1$ (recall that $\inf (B\setminus \event{0})=1$); therefore $\nu_i(t+1) \leq \nu_i(t)-1$. If $X_{t+1} = \head$ \,we are done. If $X_{t+1} = \tail$ \,then $m(t+1) = m(t) - m'(t)$, and $\g(t+1)=1$; therefore, $\lside(t+1) = \lside(t) - m'(t) + (\g(t) -1 ) \geq \lside(t) - \g(t) + (\g(t) -1 ) = \lside(t) -1$. \par (II) In this case $n_i(t+1) = n_i(t)$; hence $\nu_i(t+1) = \nu_i(t)$. If $X_{t+1} = \head$ \,we are done. If $X_{t+1} = \tail$ \,then $m(t+1) = m(t) -m'(t)$ and $\g(t+1) = 1$. But $X_{t+1} = \tail$ \,also implies (by the definition of $X$) that $\mu(t) > 0$ and $n'_{i+1}(t) / m'(t) > n_{i+1}(t) / \mu(t)$. Since $n'_{i+1}(t)$ is always $\leq n_{i+1}(t)$, we get that $\mu(t) > m'(t)$. Now, $\lside(t+1) = (m(t) - m'(t)) - (1-1 ) = m(t) - m'(t) > m(t) - \mu(t) = \nu_i(t)+1$, because $\mu(t)$ is \,$m(t) - (\nu_i(t) +1)$. Thus, $\lside(t+1) > \nu_i(t)+1 > \nu_i(t) = \nu_i(t+1)$. \end{proof} Remark: The above argument also proves that $M$ is never bankrupt, i.e., $m(t) \geq m'(t)$, and moreover $m(t) \geq \g(t)$, as follows. \par In the beginning $1=m(0)\geq \g(0)=1$. As long as $p(t)=0$ we are in case (II). In this case, if $X_{t+1}=\head$ \,then $\lside(t)$ does not decrease; hence $m(t)-\g(t)$ does not decrease. And if $X_{t+1}=\tail$ \,we just saw that actually $\lside(t+1) > 1 + \nu_i(t+1)$, which implies that $m(t)-\g(t) > 0$. \par Once $p(t)>0$, then $\nu_i(t) \geq 1$, hence $\lside(t) > \nu_i(t)$ implies that $m(t)-\g(t) > 0$; and Lemma \ref{lem:fragile} implies that $p(t)$ remains $> 0$. \begin{lemma}\label{lem:inductive} For any $i$ there exists a stage $T_i$ \,s.t. for any $t > T_i$\,, $n'_1(t) = n'_2(t) = \ldots = n'_i(t) = 0$, and $p(t) \geq i$. \end{lemma} \begin{proof}[Proof of Lemma \ref{lem:inductive}] We proceed by induction over $i \geq 0$; namely, the induction hypothesis is that the lemma holds for $i-1$. Note that the induction base case \,$i=0$ holds vacuously.\footnote{Incidentally, the inequality defining fragility always holds for $p(t)=0$, as $m(t) \geq \g(t)$ and $\nu_0(t) = 0$ imply that $m(t) - (\g(t)-1) > \nu_0(t)$.} \par If $p(t_0) \geq i$ for some stage $t_0 > T_{i-1}$, then $p(t) \geq i$ for every $t \geq t_0$, by lemma \ref{lem:fragile}. From this stage on, the casino chooses adversely to $i$ whenever $i$ is active (because the lower-index players are not active). Therefore, $i$ will be active at no more than $n_i(t_0)$ stages after $t_0$, since afterwards $i$ has nothing to wager, and we are done. \par So assume by way of contradiction that $p(t) = i-1$ for every $t > T_{i-1}$\,. Then $X_{t+1}$ is $\mu/N_i$-ratio-minimizing. As long as $\mu(t) \leq 0$ we get ``heads''; therefore, from some stage on, $\mu > 0$ (as every $a_k$ appears infinitely many times, the sum of the wagers will not converge). Denote $q(t) = n_i(t)/ \mu(t)$. $q(t) \geq 0$ is non-increasing and therefore converges to a limit $L$. Denote $0 \leq r(t) = n_i(t) - L \mu(t) $. Suppose there exists some $k$ s.t. the wagers $m'(t)$ never reach beyond $a_1,\ldots,a_k$\,. Hence, there are infinitely many stages $t$ where $i$ over-bets, i.e., $n'_i(t) > q(t)m'(t)$. For $1 \leq j \leq k$, let $x_j = L \,a_j$. Since $B$ is well ordered, there exists some $\d_j > 0$ \,s.t. $(x_j,\,x_j+\d_j) \cap B = \emptyset$\,. Since $q(t) \to L$\,, $q(t) < L + \min_{1\leq j \leq k}(\d_j/a_j)$ \,for $t$ large enough. \par When $i$ over-bets and $m'(t) = a_j$, then $r(t+1) = n_i(t+1) - L \mu(t+1) \leq n_i(t) - (La_j + \d_j) - L (\mu(t) - a_j )= r(t) - \d_j$. When $i$ does not over-bet, then $n'_i(t) \leq x_j = L a_j$\,, and $r(t+1) = n_i(t+1) - L \mu(t+1) \leq (n_i(t)+La_j) - L(\mu(t) + a_j) = r(t) $. Therefore $r(t)$ does not increase, and infinitely many times it decreases by at least $\d = \min_{1 \leq j\leq k}\d_j > 0$; hence eventually $r(t) < 0$, which is a contradiction. \par Therefore, there does not exist an index $k$ as above. This implies that for any $j$, there is a stage $t$ after which $a_j$ is wagered $f(t,j)$ consecutive times, and $N_i$ does not over-bet (otherwise $a_{j+1}$ cannot be reached). Now suppose that $L>0$\,. Let $A_0 = \event{a_1,a_2,\ldots}$ be the set of all the values that the sequence $(a_n)$ takes. $A_0$ is dense in $A$, and $L\cdot A \nsubseteq \bar B$ (since $A$ does not scale into $B$); therefore, also $L\cdot A_0 \nsubseteq \bar B$. Hence, there exists an $a \in A_0$ s.t. the distance between $La$ and $\bar B$ is $\d > 0$. \par Let $\D = \min \event{\d/a, \d}$. For $t$ larger than some $T_\D$, \,$q(t) < L + \D$. Since $a$ appears infinitely many times in the sequence $(a_n)$, there exist $T_a >T_\D$ and an index $j$, s.t. $a_j = a$ is wagered $f(T_a,j)$ consecutive times, and at each of these times $n'_i(t) \leq La-\d$, as otherwise $n'_i(t) \geq La + \d$, but that is over-betting since $(La + \d )/ a = L + \d/a > q(t)$. Hence, $r(t+1) \leq n_i(t)+La-\d - L(\,\mu(t) + a) = r(t) - \d$. But $q(T_a) < L + \d$; therefore $n_i(T_a) < (L+\d)\,\mu(T_a)$; hence $r(T_a) < \d \mu(T_a)$. By the definition of $f$, $f(T_a,j) \geq m(T_a) \geq \mu(T_a)$; therefore after those $f(T_a,j)$ times, $r < 0$\,. This cannot be; therefore $L=0$. \par As $q(t) \to 0$, surely $q(t) < 1$ for large enough $t$, namely, $\mu(t) > n_i(t)$. Since $\mu(t) = m(t) - (\nu_{i-1}(t)+1)$, we get $m(t) > n_i(t) + \nu_{i-1}(t) + 1 = n_i(t) + ((i-1) + n_1(t) + \ldots + n_{i-1}(t)) + 1 = i + n_1(t) + \ldots + n_i(t) = \nu_i(t)$. At some stage $t$, $M$ starts wagering $1$. For this $t$, $\g(t) = m'(t) =1$; hence $m(t) - (\g(t)-1) = m(t) > \nu_i(t)$, contradicting our assumption that $i$ is not fragile. \end{proof} Lemma \ref{lem:inductive} states that any $N_i$ is only active a finite number of times, and therefore it is bounded; it also states that for any $i$ and large enough $t$, $m(t) - (\g(t)-1) > \nu_i(t)$, hence $m(t) > \nu_i(t) + (\g(t)-1) > \nu_i(t) -1 \geq \,i -1$, and therefore $m(t) \to \infty $. \section{Extensions and Further Research}\label{sec:Discussion} \subsection{Extensions} The definition of anticipation corresponds to the game where gambler 0 first announces her $A$-martingale, then the regular gamblers announce their $B$-martingales, and then the casino chooses a sequence. Thus, the regular gamblers know the future actions of gambler 0, the casino knows the future actions of all gamblers, and gambler 0 does not even observe the past bets of other gamblers. \par Our proofs, however, do not rely on this state of affairs. One may consider variants of this game in which the casino does not know the future actions, or the regular gamblers do not know the future actions of gambler 0, or gambler 0 does observe the bets of others; or any combination thereof. Our results hold for all these variants. \par One of these variants looks more like a ``classic'' repeated game: at each stage of the game, first gambler 0 makes a bet, then the regular gamblers make bets, and then the casino chooses red or black (and everything is observed by all). \par Another generalization is that the set $B$ need not be the same for all regular gamblers. That is, the results still apply for sets $B_1,B_2,B_3\ldots$ of wagers for gamblers $1,2,3\ldots$, if the conditions hold for each $B_i$ separately (of course, the conditions may not hold for their union). \par \cite{bienvenu12} consider the case of computable martingales, with $A$ being the set of all integers, and $B_i$ finite sets of integers. They use a probabilistic argument to show evasion. Our proof provides, in particular, a construction of the casino sequence for this case. The constructed sequence is computable relative to the history of bets. \subsection{Further Research} It seems possible that our proof of Theorem~\ref{theorem well ordered} could be modified so as to avoid the assumption that $B$ is well ordered. We conjecture that Theorems~\ref{theorem bounded} and \ref{theorem well ordered} can be unified as the following statement. \begin{conjecture} Let $A,B\subset \mathbb R_+$. If $0\not\in\overline {B \setminus \{0\}}$, then $B$ anticipates $A$ only if $A$ scales into $B$. \end{conjecture} The present paper strove to understand the effect of restricting the wager sets on the prediction power of martingales. As is often the case, understanding one thing brings up many new questions. We list just a few. \begin{itemize} \item Under the assumptions of Theorem~\ref{theorem bounded}, $A$ can evade $B$ through a history-independent martingale. Is this also the case under the assumptions of Theorem~\ref{theorem well ordered}? \item Are single anticipation and countable anticipation different? That is, are there sets $A,B\subset\mathbb R_+$, such that $B$ countably, but not singly, anticipates $A$? \item What can be said about the anticipation relation between sets that \emph{do} include 0 as an accumulation point, for example, $\{2^{-n}\}_{n=1}^\infty$, $\{\frac 1 n\}_{n=1}^\infty$, and $\mathbb R_+$? \item \cite{buss} introduced martingales defined by probabilistic strategies. How do these martingales behave in our framework? \end{itemize} \section{Acknowledgement} We wish to thank Abraham Neyman and Bernhard von Stengle for their useful comments.
\section{Introduction} The entanglement entropy is an useful quantity to characterize the non-local property of a quantum field theory. Given a subsystem A, the degree of freedom (DOF) that is only accessible to an observer in A can be described by the density matrix $\rho_A$, which is defined by taking trace over the DOFs outside the region $A$, i.e., $\rho_A=tr_{\bar A}\rho$, where $\bar A$ is the complementary part of the subsystem A, $\rho$ is the density matrix for the whole system. The entanglement entropy is defined as the von Neumann entropy of the reduced density $\rho_A$: $S_A=-tr(\rho_A\log \rho_A)$, its holographic calculation for a CFT with a gravity dual was proposed in \cite{Ryu:2006bv}. The entanglement entropy in d-dimensional boundary theory for a spatial region A on a constant time slice is calculated by \begin{eqnarray}\label{HolographicFormula} S_A=min[\frac{s(A)}{4G}], \end{eqnarray} where $s(A)$ is the area of the codimension-2 surface in the bulk spacetime which is homologous to the boundary of spatial region $A$.\\ The \ren entropy can also be used as a quantity to describe the entanglement property. Given the reduced normalized density matrix $\rho_A$, the entanglement \ren entropy is a series of quantities defined as \begin{eqnarray}\label{RenyiFormula} S_n=\frac{\log tr[\rho_A^n]}{1-n}, \end{eqnarray} where $0<n<\infty$. The entanglement \ren entropy contains much more information about entanglement. In fact, knowing the \ren entropy $S_n$ only for all integers $n$ we can calculate the full eigenvalues of the reduced density matrix $\rho_A$. The standard way to calculate the entanglement \ren entropy is using the ``replica trick'', requiring evaluating the partition function on a singular manifold. The exact results are only known for some simply cases. The entanglement entropy is just the entanglement \ren entropy by taking the limit $n\to 1$ in (\ref{RenyiFormula}). When considering the holographic approach for the field theory living on the boundary with a gravity dual, it seems that we need to construct the dual bulk geometry for the singular manifold after performing the replica trick by using the dual relation to calculate the partition function. In \cite{Fursaev:2006ih} the author attempts to use the singular bulk space which has the required boundary singular manifold structure to prove the holographic entanglement entropy proposal (\ref{HolographicFormula}). However, it is pointed out in \cite{Headrick:2010zt} that the approach in \cite{Fursaev:2006ih} would produce an incorrect result of the bulk solution, and the entanglement \ren entropy is also independent of $n$. \\ On the other hand, in \cite{Casini:2011kv} the authors find the equivalence between the reduced density matrix $\rho_A$ of a ball in d-dimensional flat spacetime and the thermal density matrix in background $R\times H^{d-1}$ for CFT living on the boundary. This approach eliminates a singular boundary. One can calculate the thermal entropy in the bulk side, which is a topological black hole with hyperbolic horizon. The entanglement \ren entropy is also calculable following the same step, which is consistent with the known result in 2-dimensions \cite{Hung:2011nu}. But it is not easy to generalize this method to arbitrary entangling surface. \\ A recent paper \cite{Lewkowycz:2013nqa} establishes a method, named generalized gravitational entropy, to calculate entropy of an Euclidean black hole solution without a Killing vector. The entanglement entropy of an arbitrary spatial subregion for the CFT with a gravity dual is an example of this general case by the gauge/gravity duality. This argument would lead to the proposal (\ref{HolographicFormula}) beyond the spherical shape \cite{Casini:2011kv}\footnote{The subsequent papers \cite{Chen:2013qma}\cite{Bhattacharyya:2013jma} point out there are some mismatches for higher derivative gravity theory, but the mismatch is removed by the paper \cite{Dong:2013qoa}}. In this paper we first discuss the proof for the spherical entangling surface case, and use the result to calculate the entanglement \ren entropy, the result is consistent with that in \cite{Hung:2011nu}. We also notice the analogy between these two approaches. The generalized gravitational entropy \cite{Lewkowycz:2013nqa} provides us a method to study the entanglement entropy for a general entangling surface. We find that the area law of the entanglement entropy is closely related with the local property of the entangling surface. \\ This paper is organized as follows. In section \uppercase\expandafter{\romannumeral2} we briefly review the generalized gravitational entropy method and construct the dual bulk geometry for the spherical entangling surface. By using this result we study the entanglement \ren entropy with a spherical entangling surface in section \uppercase\expandafter{\romannumeral3}. We also notice the relation between the generalized gravitational entropy method and the previous approach \cite{Casini:2011kv}. In section \uppercase\expandafter{\romannumeral4} the area law of the entanglement entropy is derived for a general entangling surface. Section \uppercase\expandafter{\romannumeral5} contains some discussion about the gravity theory without a dual field theory on the boundary and the conclusion. \section{The Generalized Gravitational Entropy Method} The argument of the generalized gravitational entropy begins with the assumed underlying density matrix $\rho$ in the full quantum gravity theory. In the classical approximation, the trace of $\rho$ can be explained as Euclidean evolution of the gravity system on a ``time'' $\tau$ circle with period $2\pi$. The trace of $\rho^n$ is considered as the evolution on a circle with period $2n\pi$. Then the entropy of the system can be calculated by the ``replica trick''. The boundary condition of the system does not necessarily have a $U(1)$ symmetry along the time circle. We get $\log tr\rho$ by the saddle point approximation, which is the on-shell Euclidean action with the period of the time circle being $2\pi$. $\log tr\rho^n$ is also the on-shell Euclidean action, with a new Einstein solution. The new solution is for the spacetime with the time circle $\tau\sim \tau+2n\pi$, while the boundary condition remains the same. The entropy would be one quarter of the area of a special codimension-2 surface where the time circle shrink to zero. It is shown in \cite{Lewkowycz:2013nqa} that the special surface satisfies the minimal area condition, see \cite{Lewkowycz:2013nqa} for more details.\\ The above result is related to the holographic entanglement entropy of subregion $A$ in field theory using the gauge/gravity correspondence. The reduced density matrix $\rho_A$ can be mapped to the corresponding density matrix $\rho$ for the quantum theory in the bulk. The ``replica trick'' is actually the same as we stated above. The entanglement entropy is equal to the gravitational entropy in the bulk, which is related to the special codimension-2 surface. The remaining work is to prove that this codimension-2 surface is homologous to the boundary of subregion $A$. The holographic entanglement entropy proposal (\ref{HolographicFormula}) can be derived then. In the next two subsections we will derive the analytical results for the infinite plane and the sphere. \subsection{An infinite plane} The standard representation of the density matrix $\rho_A$ of a spatial subregion $A$ in a field theory is the path integral over the field configurations on spacetime with a cut on the entangling surface $\partial A$. The entangling surface is defined as the boundary of the subregion $A$ on a constant slice of time. $tr \rho^n$ is the path integral on the spacetime which has a conical defect with a $2n\pi$ at $\partial A$. The key point of the argument in \cite{Lewkowycz:2013nqa} is that one can use a suitable conformal transformation to ensure that the ``time'' circle around $\partial A$ does not shrink to zero size. Here ``time'' is usually not the original time coordinate in the field theory. Now it is safe to use the ``replica trick'' without worrying about the appearance of singularity. \\ Assuming we have a conformal field theory living on the d-dimensional Euclidean space $R^d$. The subregion $A$ is defined by $x>0$. The entangling surface is $t=0, x=0$. We denote the reduced density matrix by $\rho_A$. Now, as is shown in \cite{Lewkowycz:2013nqa}, one can redefine coordinates $r,\tau$, with $t=r\sin\tau$ and $x=r\cos\tau$. The metric of our spacetime is \begin{eqnarray}\label{TransformationPlane} ds^2=\Omega^2 \Big(d\tau^2+\frac{dr^2+\sum_{i=1}^{d-2} dx_i^2}{r^2}\Big), \end{eqnarray} where $\Omega=r$. The original spacetime is mapped to the manifold $R\times H^{d-1}$ by a conformal transformation eliminating the factor $\Omega^2$.\\ The subregion $A$ corresponds to region $\tau=0$,$r>0$, and its complementary part $\bar A$ corresponds to $\tau=\pi$,$r>0$ on the new manifold. $r=0$ defines the entangling surface. The reduced density matrix $\rho_A$ is mapped to an operator $\tilde\rho_A$ on the new manifold by a unitary transformation, $\tilde\rho_A$=$U$ $\rho_A$ $U^{-1}$. The new density matrix $\tilde \rho_A$ is much easier to get if the conformal field theory has a dual gravity theory in the bulk. The manifold is the boundary of $AdS_{d+1}$ with the radius $L=1$. According to AdS/CFT corresponding $\tilde \rho_A$ can be mapped to the density matrix of the bulk, which is the topological black hole with a hyperbolic horizon. The entanglement entropy is equal to $\frac{A}{4G}$, where $A$ is the area of the horizon. The remaining work is to show that the horizon is homologous to the entangling surface $r=0$, the details are presented in appendix $A$. \subsection{A sphere} The entanglement entropy of a sphere is another simple case in which we can get the analytic result by the generalized entropy method. In this section we give the construction of the proof, and the result will be used in section \uppercase\expandafter{\romannumeral3} to study the entanglement \ren entropy. The general idea is same as the infinite plane case.\\ We assume that the CFT lives on the flat d-dimensional Euclidean space $R^d$, with metric \begin{eqnarray}\label{Sphere} ds^2=dt^2+dr^2+r^2d\Omega_{d-2}^2. \end{eqnarray} Our spatial subregion $A$ is a ball with the entangling surface being $S^{d-2}$ with radius $R=1$ on a constant Euclidean time slice, i.e., $t=0$ and $r=R$. The main work is to find new manifold which has the property that the ``time'' coordinate do not shrink to zero at the entangling surface. And the new manifold should be conformal to $R^d$. We find the following coordinate transformation: \begin{eqnarray}\label{Transformation} t=\frac{\sin\tau}{\cosh u+\cos\tau}\quad \quad and \quad \quad r=\frac{\sinh u}{\cosh u +\cos\tau}, \end{eqnarray} where the new coordinate $0<u<\infty$ and $\tau$ has the period $\tau\sim \tau+2\pi$. The subregion $A$, $t=0$ and $r<1$, is mapped to $\tau=0$ and $0<u<\infty$, its complementary part is mapped to $\tau=\pi$ and $0<u<\infty$. Metric (\ref{Sphere}) is transformed to \begin{eqnarray}\label{BoundarySphere} ds^2=\Omega^2\Big(d\tau^2+du^2+\sinh^2ud\Omega_{d-2}^2\Big), \end{eqnarray} where $\Omega=(\cosh u+\cos \tau)^{-1}$, divergent on the entangling surface. Factor $\Omega^2$ can be eliminated by conformal transformation, the new manifold is also $S^1\times H^{d-1}$. The n-th power of the density matrix on $S^1\times H^{d-1}$ is equal to the path integral of fields on manifold $S^n\times H^{d-1}$. The boundary has a well-defined dual bulk geometry. By using the argument in the introduction, the entanglement entropy is obtained, given by $\frac{A}{4G}$, where $A$ is the area of the horizon of the topological black hole. In appendix $A$ we also show that the horizon is homologous to the entangling surface $S^{d-2}$. \section{Holographic \ren Entropy} Actually the method to calculate the holographic entanglement entropy in section \uppercase\expandafter{\romannumeral2} automatically provides us a way to calculate the holographic entanglement \ren entropy for CFT theories with gravity dual \footnote{Although the entanglement entropy is the limit $n\to 1$ of \ren entropy ,it is not necessary to calculate the \ren entropy in order to get the entanglement entropy. The off shell and apparent conical singularities methods are used in \cite{Lewkowycz:2013nqa} to get the result.}. In this section we will choose the entangling surface to be a sphere with a radius $R=1$. As we have noted in section \uppercase\expandafter{\romannumeral2}, the reduced density matrix $\rho _A=U^{-1}\frac{\tilde\rho_A}{tr \tilde\rho_A}U$. The trace of $n-th$ power of $\rho _A$ \begin{eqnarray} tr(\rho_A^n)=\frac{tr\tilde\rho_A^n}{(tr\tilde\rho_A)^n}. \end{eqnarray} According to the definition of \ren entropy (\ref{RenyiFormula}), \begin{eqnarray}\label{RenyiCalculation} S_n=\frac{1}{1-n}\Big(I(n)-n I(1)\Big), \end{eqnarray} where $I(n)$ is the on-shell Euclidean action with the period of Euclidean time $\tau\sim \tau +2\pi n$. Now the problem is how to construct the bulk geometry with the period of the Euclidean time being $2n\pi $. The boundary of the bulk geometry should be $S^n\times H^{d-1}$ as discussed in section \uppercase\expandafter{\romannumeral2}. The theory would admit the following form of metric with the asymptotical boundary $S^n\times H^{d-1}$ \begin{eqnarray}\label{BulkGeometry} ds^2=f(\rho)N^2d\tau^2+\frac{d\rho^2}{f(\rho)}+\rho^2(du^2+\sinh^2ud\Omega^2_{d-2}), \end{eqnarray} $f(\rho)$ is determined by the field equation of the bulk theory we are considering, $N$ is a constant to keep the curvature scale of hyperbolic spatial slices to be $R=1$. $f(\rho)N^2\to \rho^2$ in the limit $\rho\to \infty$, which ensures the boundary is conformal to (\ref{BoundarySphere}).\\ The position of the event horizon, where the time circle $\tau$ shrinks to zero, of the bulk geometry (\ref{BulkGeometry}) is determined by $f(\rho_H)=0$. Now let's define a new coordinate $x^2=f(\rho)N^2$. The metric (\ref{BulkGeometry}) becomes \begin{eqnarray} ds^2=x^2d\tau^2+\frac{4dx^2}{f'(\rho)^2N^2}+\rho^2(du^2+\sinh^2ud\Omega^2_{d-2}). \end{eqnarray} Near the horizon $\rho\to\rho_H$, $x\to 0$. The condition that the period of $\tau$ is $2n\pi$ requires \begin{eqnarray}\label{KeyFormula} \frac{f'(\rho_H)^2N^2}{4}=n^{-2}. \end{eqnarray} Equation (\ref{KeyFormula}) fixes the position of horizon $\rho_H$ and the solution of the bulk geometry.\\ As an exercise to show how this works, let us consider the case when the bulk theory is Einstein gravity firstly. \subsection{Einstein gravity} The $d+1$ dimensional Euclidean action of Einstein gravity is \begin{eqnarray}\label{EinsteinAction} I=-\frac{1}{16G}\int_Md^{d+1}x\sqrt{g}\Big(R+\frac{d(d-1)}{L^2}\Big)-\frac{1}{8G}\int_{\partial M}d^dx\sqrt{h} K, \end{eqnarray} where $M$ denotes the bulk, $\partial M$ is regularized at the infinity boundary $\rho=\rho_{\infty}$, the second integral is the standard Gibbons-Hawking term. To keep with previous notation we set the curvature scale of $AdS$ $L=1$. Using the equation of motion we get \begin{eqnarray}\label{EinsteinSoltion} f(\rho)=\rho^2-1-\frac{C}{\rho^{d-2}}, \end{eqnarray} where $C$ is an integration constant. We notice that the asymptotical condition, $\lim_{\rho\to \infty}f(\rho)=\rho^2$ is satisfied for the above solution (\ref{EinsteinAction}), so $N=1$. By using $f(\rho_H)=0$ we get the relation between $C$ and position of the horizon $\rho_H$, i.e., $C=\rho_H^d-\rho_H^{d-2}$. Formula (\ref{KeyFormula}) determines the position of horizon, we get \begin{eqnarray}\label{QuadraticEquation} d \rho_H^2-(d-2)=2n^{-1}\rho_H, \end{eqnarray} when $n=1$ we have $C=0$, the solution is the hyperbolic version of $AdS$. \\ Next let us turn to (\ref{RenyiCalculation}). We must know how to deal with divergence of the Euclidean action. For $AdS$ spacetime there is a natural way to solve the problem. The Euclidean action should contain the following terms. \begin{eqnarray} I=I_{bulk}+I_{surface}+I_{ct}, \end{eqnarray} where the surface term is regularized at the boundary of the bulk on a constant slice of $\rho$, the counter term depends only on the curvature $\mathcal R$ and its derivatives on the induced boundary \cite{Balasubramanian:1999re}\cite{Emparan:1999pm}\cite{de Haro:2000xn}. The counter terms in this method to regularize the action are referred to as Dirichlet counter terms \cite{Kofinas:2007ns}\cite{Olea:2006vd}\cite{Kofinas:2006hr}. The form of counter terms are related with the dimension of bulk spacetime. It is impossible to write down the counter term explicitly for general dimensions. Action (\ref{EinsteinAction}) is already regularized at $\rho_{\infty}$, we need to add counter terms to cancel divergence. For example the standard counter term for $AdS_3$ is \begin{eqnarray} I_{ct}^{d=2}=\frac{1}{8\pi G}\int_{\partial M}\sqrt{h}, \end{eqnarray} where $h$ is the determinant of the boundary $\rho=\rho_\infty$. The action is \begin{eqnarray} I(n;d=2)=-\frac{n V_{\Sigma}}{8G}(\rho_H^2+2), \end{eqnarray} The entanglement \ren entropy is \begin{eqnarray} S_n^{(d=2)}=\frac{V_\Sigma}{8G}(1+\frac{1}{n}), \end{eqnarray} the result is the same as in the field theory approach \cite{Holzhey:1994we}, as well as by the holographic method \cite{Hung:2011nu}. Another example is $AdS_5$, the action is \begin{eqnarray}\label{Counterterm4} I(n;d=4)=\frac{nV_{\Sigma}}{8G}\Big(-\frac{3}{4}+\rho_H^2+\rho_H^4\Big). \end{eqnarray} Then we get the entanglement \ren entropy \begin{eqnarray} S_n^{(d=4)}=\frac{n\pi V_{\Sigma}}{8G(n-1)}\Big(2-\rho_H^2-\rho_H^4\Big), \end{eqnarray} In appendix B we also calculate the cases $d=3$ and $d=5$, the general form for arbitrary dimension is \begin{eqnarray} S_n= \frac{n V_\Sigma L^{d-1}}{8(n-1)G}(2-\rho_H^d-\rho_H^{d-2}). \end{eqnarray} We will prove this in next subsection. It is interesting to take the limit $n\to 1$, we recover the entanglement entropy \begin{eqnarray} S_A=\lim_{n\to 1}S_n=\frac{V_{\Sigma}}{4G}, \end{eqnarray} where we have used solution to eq.(\ref{QuadraticEquation}). \subsection{Lovelock gravity for arbitrary dimension} The Lovelock action in $d+1$ dimensions is \cite{Lovelock:1971yv}. \begin{eqnarray} \label{Lovelock} I=I_{bulk}+I_{surface}=\frac{1}{16 \text{$\pi $G}}\int _Md^{d+1}x\sqrt{-g}\sum _{p=0}^{\left[\frac{d}{2}\right]} L_{2 p} \alpha _p+c_d\int _{\partial M}d^d \text{xB}_d \end{eqnarray} where \begin{eqnarray} L_{2p}\equiv \frac{1}{2^p}\delta^{[\nu_1 \nu_2...\nu_{2p-1}\nu_{2p}]}_{[\mu_1 \mu_2...\mu_{2p-1}\mu_{2p}]}R^{\mu _1 \mu _2}{}_{\nu _1 \nu _2}\cdot \cdot \cdot R^{\mu _{2 p-1} \mu _{2 p}}{}_{\nu _{2 p-1} \nu _{2 p}}, \end{eqnarray} $B_d$ denotes the boundary term and $c_d$ is the suitable constant, $\alpha_0=\frac{d(d+1)}{L^2}$, $\alpha_1=1$ and others are arbitrary. For later convenience we also define \begin{eqnarray} \lambda_p=(-1)^p\frac{(d-2)!}{(d-2p)!} \frac{\alpha_p}{L^{2p-2}}, \end{eqnarray} with $\lambda_0=1$ and $\lambda_1=-1$. Lovelock theory admits a pure $AdS_{d+1}$ solution with the curvature scale $\tilde L=L/\sqrt{f_{\infty}}$, where $f_{\infty}$ satisfies the following expression \begin{eqnarray}\label{ConstraintOflambda} \sum _{p=0}^{\left[\frac{d}{2}\right]} \lambda _p f_{\infty }^p=0, \end{eqnarray} we choose the root $f_\infty\to 1$ in the limit $\lambda_p\to 0$. By using the equation of motion of Lovelock gravity $f(\rho)$ should satisfy the following relation \begin{eqnarray}\label{ConstraintLovelock} \sum_{p=0}^{[\frac{d}{2}]}\lambda_p[1+f(\rho)]^p\rho^{d-2p}L^{2p-2}=\mu, \end{eqnarray} where $\mu$ is the integral constant. At the horizon $\rho=\rho_H$, $f(\rho)$ is vanishing, we get \begin{eqnarray}\label{mu} \mu=\sum_{p=0}^{[\frac{d}{2}]}\lambda_p\rho_{H}^{d-2p}L^{2p-2} \end{eqnarray} and \begin{eqnarray}\label{Lovelockderivative} f'(\rho_H)\sum_{p=0}^{[\frac{d}{2}]}p\lambda_p \rho_{H}^{d-2p}L^{2p-2}+\sum_{p=0}^{[\frac{d}{2}]}(d-2p)\lambda_p \rho_{H}^{d-2p-1}L^{2p-2}=0. \end{eqnarray} From (\ref{ConstraintLovelock}) we obtain the asymptotically behavior of $f(\rho)$ at large $\rho$, \begin{eqnarray}\label{Expandingoffrho} f(\rho)=\frac{\rho^2}{\tilde L^2}-1+\mu \Big(\sum_{p=1}^{[\frac{d}{2}]}p\lambda_p f_\infty^{p-1} \Big)^{-1}\frac{1}{\rho^{d-2}}+..., \end{eqnarray} the constant $N$ in (\ref{BulkGeometry}) should be $\tilde L$. Plugging (\ref{KeyFormula}) into (\ref{Lovelockderivative}) we get the solution of the bulk geometry with the period $\tau\sim \tau +2n\pi$, with the root of the following algebraic equation \begin{eqnarray}\label{Alegbraequation} \frac{2}{n\tilde L}\sum_{p=0}^{[\frac{d}{2}]}p\lambda_p \rho_{H}^{d-2p}L^{2p-2}+\sum_{p=0}^{[\frac{d}{2}]}(d-2p)\lambda_p \rho_{H}^{d-2p-1}L^{2p-2}=0. \end{eqnarray} It is obvious that the solution for $n=1$ is $\rho_{H}=\tilde L$ by using eq.(\ref{ConstraintOflambda}), which is the location of the horizon of the pure AdS Lovelock solution in the hyperbolic foliation.\\ The second term in (\ref{Lovelock}) is the boundary term at the infinity $\rho=\rho_{\infty}$, and contains the curvature term $\mathcal R$ and extrinsic curvature term $K$ at the boundary. This would make the variation work well and renormalize the action. These terms, also referred to as the Kounterterms, are discussed in \cite{Olea:2006vd}\cite{Kofinas:2006hr}\cite{Kofinas:2007ns}, see appendix B for more details. The action is different for even and odd dimensions, we should discuss them separately. We just quote the renormalized action for even dimension (see appendix B), \begin{eqnarray}\label{ActionofLovelock} I(n)=\frac{V_\Sigma}{16\pi G}\Big[&-&4\pi\sum_{p=1}^{[\frac{d}{2}]}\frac{d-1}{d-2p+1}p\lambda_p\rho_H^{d-2p+1}L^{2p-2}\nonumber \\ &-&2n\pi \tilde L (d-1)\mu +\frac{4\pi (d-1)!!^2}{d}\sum_{p=1}^{[\frac{d}{2}]}\frac{d-1}{d-2p+1}p \lambda_p \tilde L^{d-2p+1}L^{2p-2}\Big], \end{eqnarray} where $\mu$ is given by (\ref{mu}) and $\rho_{H}$ is the root of (\ref{Alegbraequation}). Taking $\rho_H=\tilde L$ and $n=1$ we get action $I(1)$. The entanglement \ren entropy (\ref{RenyiFormula}) is \begin{eqnarray} S_n=\frac{n V_\Sigma}{4(n-1)G}\Big[&&\frac{1}{n}\sum_{p=1}^{[\frac{d}{2}]}\frac{d-1}{d-2p+1}p\lambda_p\rho_H^{d-2p+1}L^{2p-2}\nonumber \\ &&-\tilde L^{d-1}\sum_{p=1}^{[\frac{d}{2}]}p\lambda_p f_\infty^{2p-2}+\frac{1}{2}(d-1)\mu\Big]. \end{eqnarray} For the odd dimensions the renormalized action does not contain the third term, which is related to the vacuum energy, in the square bracket of (\ref{ActionofLovelock}), see appendix B. This term does not contribute to the entanglement \ren entropy, so the result in odd dimensions is the same as in even dimensions.\\ The Gauss-Bonnet gravity ($d\ge 4$) is a special case of Lovelock gravity theory with $\lambda_0=1$, $\lambda_1=-1$, $\lambda_2\equiv \lambda$ and other coefficients being zero. The entanglement \ren entropy is \begin{eqnarray} S_n&=&\frac{n V_\Sigma \tilde L^{d-1}}{8(n-1)G}\Big[\frac{\left(\frac{d \rho _H^4}{f_{\infty }}+(d-4) \lambda f_{\infty }-(d-2) \rho _H^2\right) \left(\frac{2 (d-1) \lambda f_{\infty } \rho _H^{d-4}}{d-3}-\rho _H^{d-2}\right)}{\rho _H^2-2 \lambda f_{\infty }}\nonumber\\ &&-\frac{4 (d-1) \lambda f_{\infty }}{d-3}+2 +(d-1) \left(\lambda f_{\infty } \rho _H^{d-4}+\frac{\rho _H^d}{f_{\infty }}-\rho _H^{d-2}\right)\Big], \end{eqnarray} the result is the same as in \cite{Hung:2011nu}. In the limit $\lambda\to 0$ we get the entanglement \ren entropy for Einstein gravity in arbitrary dimensions \begin{eqnarray} S_n= \frac{n V_\Sigma L^{d-1}}{8(n-1)G}(2-\rho_H^d-\rho_H^{d-2}). \end{eqnarray} \subsection{Relation between generalized entropy method and previous approach} In \cite{Casini:2011kv} the authors use a different approach and finally get the holographic entanglement entropy formula (\ref{HolographicFormula}). The construction of \cite{Casini:2011kv} relies on conformal transformation which maps the causal development $\mathcal D $ of the ball to the new geometry $R\times H^{d-1}$. Using the algebraic approach to quantum field theory they obtain the result that the reduced density matrix of the ball is related to the thermal density matrix on the geometry $R\times H^{d-1}$. The entropy of thermal density state can be calculated by AdS/CFT correspondence, which relates this thermal state to the bulk topological black hole. The generalized gravitational entropy is similar as the above statement. In section \uppercase\expandafter{\romannumeral2} transformation (\ref{Transformation}) is the same as the one in \cite{Casini:2011kv} in the Euclidean version. Actually the generalized gravitational entropy method does not need to prove that the density matrix on $S^1\times H^{d-1}$ is thermal. A conformal transformation maps our spatial subregion $A$ and its complementary $\bar A$ into the surface $\tau=0$ and $\tau=\pi$ in the new spacetime respectively. The DOFs on $\tau=0$ and $\tau=\pi$ represent the whole system. Coordinate $\tau$ in the new spacetime is periodic and does not shrink to zero at the entangling surface. So the ``replica trick'' would produce the smooth geometry\footnote{The idea of generalized gravitational entropy is analogous to the paper \cite{Holzhey:1994we}, in which the authors find the conformal transformation mapping the 2-dimensional Euclidean spacetime to strip, and the internal and its complementary part are mapped to two boundary of the strip.}. By using AdS/CFT correspondence we can calculate the density matrix. Actually this property also appears in \cite{Barvinsky:1994jca}\cite{Maldacena:2001kr}, in which the authors discuss the entanglement of the eternal black hole. \\ The wave function of the system on $S^1\times H^{d-1}$ can be written as the path integral of fields in the interval $0\le\tau\le \pi$ with the boundary conditions $\phi(\tau=0)=\phi_1$ and $\phi(\tau=\pi)=\phi_2$. After tracing over the DOFs of the subregion $\bar A$, i.e., the surface $\tau=\pi$ in the new geometry, the reduced density matrix $\rho_A$ is equal to the partition function at temperature $T=\frac{1}{2\pi}$ derived in \cite{Casini:2011kv}. The n-th power of $\rho_A$ is still a thermal state with temperature $T/n$. This just tells the fact that the thermal density matrix can arise from entanglement, see \cite{Maldacena:2001kr}. The two approaches for spheric entangling surface can be viewed as the same phenomenon in different perspectives. But the latter is easily generalized to arbitrary entangling surface. Our calculation of the entanglement \ren entropy in this paper is similar to \cite{Hung:2011nu}, as expected. The approach in \cite{Casini:2011kv} can be seen as evidence for the generalized gravitational entropy method. \section{Entanglement Entropy beyond Simple case} Our discussion above is about a sphere and an infinite plane, in both cases the dual bulk geometry can be found. For a general entangling surface it is expected that the geometry wll depend on Euclidean time $\tau$, it seems impossible to construct the analytical solution of the bulk geometry. In this section we use the argument of generalized entropy and try to get some results for an entangling surface beyond the simple cases. \\ Let us first consider the entangling surface to be an infinite stripe in d-dimensional flat spacetime. The subregion $A$ is defined as $x_0=0$, $0<x_1<l$, $-l_0<x_i<l_0$$(i=2,...,d-1)$, where $l_0$ is infinite. There are two separated entangling surfaces located at $x_1=0$ and $x_1=l$. Unlike the infinite plane that we discussed in section \uppercase\expandafter{\romannumeral2} it seems impossible to map the entangling area into a region and keep the metric to be independent of the new time coordinate $\tau$. The dependence on time $\tau$ is expected for a general configuration when we try to make a conformal transformation, see \cite{Lewkowycz:2013nqa}. We expect that the boundary defines the bulk geometry and the $\tau$ circle shrinks to zero in the bulk. There are surfaces where without knowing the exact form of the bulk geometry we can also get the leading contribution.\\ The metric of the d-dimensional spacetime can be written as \begin{eqnarray}\label{MetricStripe} ds^2=dzd\bar z+\sum_{i=2}^{d-1}dx_i^2, \end{eqnarray} where $z=ix_0+x_1$. We consider the following transformation, \begin{eqnarray}\label{TransfromationStripe} \frac{z-l}{z}=e^{-\omega}, \end{eqnarray} and define the new coordinate $\omega=i\tau+u$. The metric (\ref{MetricStripe}) becomes \begin{eqnarray}\label{StripeMetric} ds^2=\Omega^2\Big(d\tau^2+du^2+4\frac{(\cosh u-\cos\tau)^2}{l^2}\sum_{i=2}^{d-1}dx_i^2\Big), \end{eqnarray} where $\Omega^2=l^{-2}z\bar z (z-l)(\bar z-l)$. The subregion $A$ is mapped to $\tau=\pi$ and $-\infty<u<+\infty$, $u=-\infty (+\infty)$ corresponds to the entangling surface $x_1=0(l)$. Its complementary part $\bar A$ is mapped to $\tau=0$. This is what we expect. Actually we may find many different transformations to get the desired result. Another argument for the above transformation (\ref{TransfromationStripe}) is that we would obtain the transformation for the infinite plane (\ref{TransformationPlane}) in the limit $l\to\infty$. In fact in the limit $l\to \infty$, $\Omega^2 \to z\bar z\equiv x_0^2+x_1^2$, $\Omega^2 du^2\to (d\sqrt{z\bar z})^2$. The limit $l\to \infty$ means we only have one entangling plane located at $x_1=0$. \\ Although unable to construct the dual bulk geometry with a ``time'' dependent boundary, we can analyse its asymptotical behavior. In the limit $|u|\to \infty$, metric (\ref{StripeMetric}) is asymptotically conformal to \begin{eqnarray}\label{AsymptoticalMetric} ds^2=d\tau^2+du^2+\frac{e^{\pm 2u}}{l^2}\sum_{i=2}^{d-1}dx_i^2, \end{eqnarray} where $\pm$ depends on limit $u\to -\infty(+\infty)$. Near the region $|u|\sim \infty$ we define the new coordinate $r=le^{\pm u}$, metric (\ref{AsymptoticalMetric}) changes to \begin{eqnarray}\label{PlaneSimilar} ds^2=d\tau^2+\frac{dr^2+\sum_{i=2}^{d-1}dx_i^2}{r^2}, \end{eqnarray} This is just metric (\ref{TransformationPlane}) in section 2.1. We expect the dual geometry of (\ref{StripeMetric}) is asymptotic to (\ref{DualGeoofPlane}), which is the dual geometry of boundary metric (\ref{PlaneSimilar}). The leading contribution to the entanglement entropy is \begin{eqnarray}\label{AreaLawStripe} S_A\sim \int dr dx_2...dx_{d-1} \frac{1}{r^{d-1}}\sim \frac{A_{ent}}{\delta^{d-2}}, \end{eqnarray} where $A_{ent}\equiv l_0^{d-2}$ denotes the area of the entangling surface, $\delta$ is the cut off to regularize the result. This is required by the area law of entanglement entropy. Notice that we have two entangling surfaces here, their leading contributions both have the form of (\ref{AreaLawStripe}). The contribution is from the region $|u|\sim \infty$, near the entangling surfaces. This result confirms the fact that the leading contribution of the entanglement entropy is due to the short distance coupling near the entangling surface. From the transformation we observe that $z\to 0$, the entangling surface, in the limit $u\to \infty$. The new coordinate $r=le^{-u} \to \sqrt{z\bar z}=\sqrt{x_0^2+x_1^2}$, consistent with the coordinate $r$ that is defined in section 2.1. This implies that transformation (\ref{TransfromationStripe}) reduces to simple transformation of the infinite plane (\ref{TransformationPlane}). This is natural because the entangling surfaces locally look like an infinite plane.\\ For general entangling surface the metric can be written in Gauss normal coordinates \begin{eqnarray}\label{MetricArbitrary} ds^2=dt^2+dy^2+g_{ij}(y,x)dx^idx^j,\quad i,j=1,...,d-2, \end{eqnarray} where $y$ is the coordinate normal to the entangling surface, $x_i$ are the ones tangent to the surface. The metric $g_{ij}(y,x)=g_{ij}(0,x)+2yK_{ij}+O(y^2)$ near the surface. The entangling surface is defined by $t=0$ and $y=0$. The subregion $A$ $(y<0)$ and its complementary part $\bar A$ are mapped to a $\tau=0$ and $\tau=\pi$ by a suitable transformation $U$, where $\tau$ is periodic and encircles the entangling surface. The new geometry is generally time dependent. Our discussion in the above example can be generalized to an arbitrary entangling surface. Metric (\ref{MetricArbitrary}) near the entangling surface is \begin{eqnarray} ds^2=dt^2+dy^2+\delta_{ij}d\xi^id\xi^j+..., \quad i,j=1,...,d-2, \end{eqnarray} where new coordinates $\xi_i$ satisfy $\delta_{ij}d\xi^id\xi^j=g(0,x)dx^idx^j$, ``...'' are terms are of higher orders. As we have mentioned the local property of the entangling surface is same as an infinite plane. We expect the metric (\ref{MetricArbitrary}) after transformation $U$ would reduce to \begin{eqnarray}\label{InfintePlaneTrans} ds^2=\Omega^2(d\tau^2+\frac{dr^2+\delta_{ij}d\xi^id\xi^j}{r^2}), \end{eqnarray} near the entangling surface, where $y=r\cos\tau$, $t=r\sin\tau$, and $\Omega=r^2$, see the above example. And the dual bulk geometry would also have an asymptotical behavior as the dual geometry of (\ref{InfintePlaneTrans}) locally. We can glue the different patches smoothly and choose a cut-off $r\sim \delta$, the leading contribution of the entanglement entropy would be \begin{eqnarray}\label{AreaLaw} S_{leading}&\sim& \int_{r\sim 0}drd\xi_1d\xi_2...d\xi_{d-2}\frac{1}{r^{d-1}} \nonumber \\ &=&\int_{r\sim 0}dr dx_1...dx_{d-2}\frac{1}{r^{d-1}}\sqrt{Det g_{ij}(0,x)}\sim\frac{A_{ent}}{\delta^{d-2}}, \end{eqnarray} where $r\sim 0$ means the integral region is near the entangling surface and along the entangling surface, and we use $|\frac{\partial \xi_i}{\partial x_j}|=\sqrt{Det g_{ij}(0,x)}$. (\ref{AreaLaw}) is the area law of the entanglement entropy\cite{Srednicki:1993im}\cite{Eisert:2008ur}. \section{Discussion and Conclusion} Our above discussion is mainly about the entanglement property of a spatial subregion $A$ on the boundary of the bulk geometry. The generalized gravitational entropy is seen as the the entanglement entropy of more general system, the holographic entanglement entropy is only an example by the $AdS/CFT$ correspondence. For gravity theory with a dual boundary theory the density matrix $\rho$ in the bulk has a well defined map in the boundary field theory. This motivates a similar explanation for gravity theory without knowing the dual boundary field theory. For a general quantum theory we should know the wave function of the system to define the entanglement property of subregion in it. We follow the proposal of \cite{Barvinsky:1994jca}, where the authors consider the eternal version of a Schwarzchild black hole with mass $M$ in 4 dimension. The wave function of the black hole is the Euclidean path integral over field configurations $\phi$ on the region $-2\pi M<\tau<2\pi M$, where $\tau$ is the Euclidean time with the period $8\pi M$. $\phi(\tau=0)=\phi_1$ and $\phi(\tau=4\pi M)=\phi_2$ are the boundary conditions of the wave function. The boundary is the Cauchy surface with the topology $R\times S^2$. As was shown in \cite{Barvinsky:1994jca}, boundaries $\tau=0$ and $\tau=4\pi M$ represent the inside and the outside DOFs respectively. The wave function can be written as a path integral \begin{eqnarray}\label{WaveFuncofBlackhole} \Phi[\phi_1,\phi_2]=\int_{\phi(\tau=0)=\phi_1,\phi(\tau=4\pi M)=\phi_2}D\phi e^{-I_E}, \end{eqnarray} where $I_E$ is the Euclidean action including gravity and matter. With the wave function we can define the reduced density matrix $\rho$ by tracing over DOFs inside the horizon, a path integral on the $\tau$ circle with a cut at $\tau=8\pi M$. The n-th power of $\rho$ is given by the path integral on another spacetime with $\tau \sim \tau +8n\pi M$, similar to the generalized gravitational entropy. But in the new spacetime there is a conical singularity on the surface where the $\tau$ circle shrinks to zero \cite{Solodukhin S N}. This seems to contradict to the argument of the generalized gravitational entropy. But in fact we may consider the orbifold of the smooth bulk solution $M_n$ with $\tau\sim\tau+2n\pi$ $M_n/Z_n$, then there is a conical defect $8\pi M(1-1/n)\equiv 8\pi M\epsilon$ at the horizon. We can calculate the on-shell action $\hat I$ without considering the effect of the conical defect, then $I(n)=n\hat I$. Actually $\partial_\epsilon \hat I$ is equal to the variance of the action with an opening angle $8\pi M(1+\epsilon)$ by $\epsilon$, see the paper\cite{Lewkowycz:2013nqa} \cite{Dong:2013qoa}. We can use both methods to calculate entropy. But the generalized gravitational entropy offers a physical explanation. The calculation is simple in the semiclassical approximation . The smooth spacetime with period $\tau\sim\tau+8n\pi M$ is the Schwarzchild solution with parameter $M$ replaced by $n M$. The renormalized on-shell action is $I(n)=-\frac{4\pi nM^2}{G} $\footnote{The renormalization procedure is the same as in \cite{Gibbons}}, so the entanglement \ren entropy is $S_n=\frac{4nM^2}{G}$. We find the entanglement entropy $S=\lim_{n\to 1}S_n=\frac{A}{4G}$. Actually the on-shell action of the smooth spacetime and the one with a conical singularity are equal in the limit $n\to 1$. The entanglement \ren entropy due to the generalized gravitational entropy also has a simple and nice relation with the thermodynamical ensemble of the black hole, we can slightly modify the \ren entropy formula (\ref{RenyiCalculation}) as in \cite{Baez} \begin{eqnarray} S_n=\frac{TI(n)-T_0I(1)}{T-T_0}=-\frac{F(T)-F(T_0)}{T-T_0}, \end{eqnarray} where $T_0$ is the temperature of the black hole with parameter $M$, $F(T)=-TI(n)$ is the free energy of the system at temperature $T$, we define the new temperature $T=T_0/n$. We observe that $T$ is just the black hole temperature with the period $\tau\sim \tau+8n\pi M$. In the limit $n\to 1$, $S_n\to S=-\frac{\partial F}{\partial T}$, which is the thermal relation between the free energy and the entropy of the system. Our discussion here is valid only in the semiclassical approximation, it is worth to study whether the on-shell generalized gravitational entropy method gives the same answer as in the off-shell approach at the quantum level when quantum matter appears.\\ Another interesting thing concerning the entanglement entropy is the the variance of entanglement entropy for excited state that is discussed in many paper recently, e.g., \cite{Bhattacharya:2012mi}-\cite{He:2013rsa}. It's interesting to consider how to understand the variance of the entanglement entropy by the generalized gravitational entropy.\\ In this paper we use the generalized gravitational entropy method to study the \ren entropy and entanglement entropy for a subregion in conformal field theory. For the simple sphere case we find the transformation mapping the region to desired geometry with a bulk dual. By AdS/CFT correspondence we calculate the entanglement \ren entropy of the sphere for Einstein gravity. We find that the approach has a close relation with the previous one \cite{Casini:2011kv} and confirm the relation between the reduced and the thermal states for the field theory. We also study the entangling surface beyond the simplest case. Although unable to construct the dual bulk geometry for a general case, one can get some information about the entanglement entropy from the local property near the entangling surface. And we derive the famous area law of entanglement entropy by this method. \section{Acknowledgement} We wish to thank Song He and Rong-xin Miao for useful discussions on this topic.
\section{Introduction} The human brain neural network has an enormous complexity containing about $10^{11}$ neurons and $10^{14}$ synapses linking various neurons \cite{neuronumber}. Such a complex network can only be compared with the World Wide Web (WWWW) which indexed size is estimated to be of about $10^{10}$ pages \cite{wwwsize}. This comparison gives an idea that the methods of computer science, developed for WWW analysis, can be suitable for the investigations of neural networks. Among these methods the PageRank algorithm of the Google matrix of WWW \cite{brin} clearly demonstrated its efficiency being at the heart of Google search engine \cite{googlebook}. Thus we can expect that the Google matrix analysis can find useful applications for the neural networks. This approach has been tested in \cite{szbrain} on a reduced brain model of mammalian thalamocortical systems studied in \cite{izhikevich}. However, it is more interesting to perform the Google matrix analysis for real neural networks. In this Letter we apply this analysis to characterize the properties of neural network of {\it C.elegans} (worm). The full connectivity of this directed network is known and documented at \cite{wormatlas}. The number of linked neurons (nodes) is $N=279$ with the number of synaptic connections and gap junctions (links) between them being $N_{\ell}=2990$. Recently, there is a growing interest to the complex network approach for investigation of brain neural networks \cite{arenas,bullmore},\cite{kaiser,chklovskii},\cite{towlson}. Generally these networks are directional but it is difficult to determine directionality of links by physical and physiological measurements. Thus, at present, the worm network is practically the only example of neural network where the directionality of all links is established \cite{wormatlas}. The analysis of certain properties this directed network has been reported recently in \cite{chklovskii,towlson}, however, the approach based on the Google matrix has not been used yet. Thus we think that this study will allow to highlight the features of worm network using recent advancements of computer science. \section{Google matrix construction} The Google matrix $G$ of {\it C.elegans} is constructed using the connectivity matrix elements $S_{ij}=S_{syn,ij}+S_{gap,ij}$, where $S_{syn}$ is an asymmetric matrix of synaptic links whose elements are $1$ if neuron $j$ connects to neuron $i$ through a chemical synaptic connection and $0$ otherwise. The matrix part $S_{gap}$ is a symmetric matrix describing gap junctions between pairs of cells, $S_{gap,ij}=S_{gap,ji}=1$ if neurons $i$ and $j$ are connected through a gap junction and $0$ otherwise. Following the standard rule \cite{brin,googlebook}, the matrix elements $S_{ij}$ are renormalized ($S_{ij} = S_{ij}/\sum_i S_{ij}$) for each column with non-zero elements; the columns with all zero elements are replaced by columns with all elements $1/N$. Thus the sum of elements in each column is equal to unity and the Google matrix takes the form \begin{equation} G_{ij}=\alpha S_{ij} + (1-\alpha)/N \; \; . \label{eq1} \end{equation} Here $\alpha$ is the damping factor introduced in \cite{brin}. In the context of the WWW, the last term of the equation describes a probability for a random surfer to jump on any node of the network \cite{googlebook}. We use the usual value $\alpha=0.85$ \cite{googlebook}. All matrix elements $S_{syn,ij}, S_{gap,ij}, S_{ij}$ are given at \cite{wormgooglematrix} \begin{figure}[!ht] \begin{center} \includegraphics[width=7.7cm]{fig1.eps} \caption {(Colour on-line) Google matrix $G$ (left) and $G^*$ (right) for the neural network of {\it C.elegans} for $N=279$ connected neurons. Matrix elements $G_{KK'}$ are shown in the basis of PageRank index $K$ (and $K'$) and elements $G^*_{K^*,K^{*'}}$ are shown in the basis of CheiRank index $K^*$ (and $K^{*'}$) at $\alpha=0.85$. Here, $x$ and $y$ axes show $1 \leq K, K' \leq N$ and $1 \leq K^*, K^{*'} \leq N$; the elements $G_{11}, {G^*}_{11}$ are placed at the top left corner; color is proportional to the square root of matrix elements which are changing from black at minimum value $(1-\alpha)/N$ to light yellow at maximum.} \label{fig1} \end{center} \end{figure} The eigenspectrum $\lambda_i$ and right eigenvectors $\psi_i(j)$ of $G$ satisfy the equation \begin{equation} \sum_{j'} G_{jj'} \psi_i(j') = \lambda_i \psi_i(j) \; . \label{eq2} \end{equation} The eigenvector at $\lambda=1$ is known as the PageRank vector. According to the Perron-Frobenius theorem \cite{googlebook} its elements $P(j) \sim \psi_1(j)$ are positive and their sum is normalized to unity. Thus $P(j)$ gives a probability to find a random surfer on a node $j$. All nodes can be ordered in a decreasing order of probability $P(K_j)$ with highest probability at top values of PageRank index $K_j=1, 2, ...$. For large matrices $P(j)$ can be found numerically by the iteration method \cite{googlebook} but for {\it C.elegans} case it can be obtained by a direct matrix diagonalization. It is also useful to consider the Google matrix obtained from the network with inverted directions of links (see e.g. \cite{alik,zzs},\cite{2dmotor}). The matrix $G^*$ for this network with inverted direction of links is constructed following the same definition (\ref{eq1}). The PageRank vector of this matrix $G^*$ is called the CheiRank vector with probability $P^*(K^*_j)$ and CheiRank index $K^*$. According to the known results \cite{brin,googlebook} the top nodes of PageRank are the most popular pages, while the top nodes of CheiRank are the most communicative nodes \cite{zzs,2dmotor}. \begin{figure}[!ht] \begin{center} \includegraphics[trim = 0mm 0mm 0mm 0mm, clip, width=7.7cm]{fig2.eps} \caption {(Colour on-line) Top panel: spectrum of eigenvalues $\lambda$ for the Google matrices $G$ and $G^*$ at $\alpha=0.85$ (black and red symbols). Bottom panel: IPR $\xi$ of eigenvectors as a function of corresponding $Re \lambda$ (same colors).} \label{fig2} \end{center} \end{figure} The structure of the matrix elements of $G$, presented in the PageRank ordering of nodes, and $G^*$, presented in the CheiRank ordering of nodes, is shown in Fig.~\ref{fig1}. The number of nonzero elements $N_G$ with indexes less than $K$ is respectively $N_G/K \approx 1.2, 10$ at $K=10, 100$. These values correspond approximately to those of WWW networks of Universities of Cambridge and Oxford being significantly smaller than the values of Twitter network characterized by a strong connectivity between top PageRank nodes with $N_g/K \approx 100$ for $K=100$ (see Fig.2 in \cite{physrev}). We note that the average number of links per neuron is $\eta = N_{\ell}/N =10.71$ being approximately the same as for WWW of Universities of Cambridge and Oxford in 2006 \cite{2dmotor}. The global matrix structure is asymmetric. This leads to a complex spectrum of eigenvalues of $G$ and $G^*$ as shown in top panel of Fig.~\ref{fig2}. The imaginary part of eigenvalues is distributed in a range $-0.2 < Im \lambda <0.2$ which is more narrow than for the networks of Wikipedia and UK universities \cite{wikispectrum}. This is related to a significant number of symmetric links. On the other side the networks of Le Monde or Python have comparable width for $Im \lambda$ \cite{wikispectrum}. We find that the second by modulus eigenvalues are $\lambda_2= 0.8214$ for $G$ and $\lambda_2= 0.8608$ for $G^*$. Thus the network relaxation time $\tau = 1/|\ln \lambda_2|$ is approximately $5, 6.7$ iterations of $G, G^*$. The properties of eigenstates $\psi_i$ can be characterized by the Inverse Participation Ratio (IPR) $\xi_i=\left( \sum_j |\psi_i(j)|^2 \right)^2/\sum_j |\psi_i(j)|^4$, which is broadly used in analysis of electron conductivity in disordered systems (see e.g. \cite{wikispectrum,physrev}). This quantity effectively determines the number of nodes on which is located an eigenstate $\psi_i$. We see that some eigenstates have rather large $\xi \approx N/3$ while others have $\xi$ located only on about ten nodes. We will return to the discussion of properties of eigenstates later. \section{CheiRank versus PageRank} The dependence of probabilities of PageRank and CheiRank vectors on their indexes $K$ and $K^*$ is shown in Fig.~\ref{fig3}. A formal fit for a power law dependence $P \propto 1/K^{\nu}, P^* \propto 1/K^{* \nu}$ in the range $1 \leq K, K^* \leq 200$ gives $\nu = 0.33 \pm 0.03$ for PageRank and $\nu = 0.50 \pm 0.03$ for CheiRank. Of course, the number of nodes is small compared to the WWW or Wikipedia networks but on average we can say that a power law provides a satisfactory description of data. We note that the values of $\nu$ are notably smaller than the usual exponent value $\nu \approx 0.9$ (in $K$), $0.6$ (in $K^*$) found for the WWW or Wikipedia networks (see e.g. \cite{googlebook,zzs}). Also in our neural network we find that the exponent in $K$ is smaller then in $K^*$ while usually one finds the opposite situation. Also we have IPR $\xi \approx 85$ for $P$ and $\xi \approx 23$ for $P^*$ so that PageRank is distributed over a larger number of neurons. It is possible that such an inversion is related to a significant importance of outgoing links in neural systems: in a sense such links transfer orders, while ingoing links bring instructions to a given neuron from other neurons. We note that somewhat similar situation appears for networks of Business Process Management (BMP) where {\it Principals} of a company are located at the top CheiRank position while the top PageRank positions belong to company {\it Contacts} \cite{abel}. \begin{figure}[!ht] \begin{center} \includegraphics[trim = 0mm 0mm 0mm 0mm, clip,width=7.5cm]{fig3.eps} \caption {(Colour on-line) \emph{Left panel:} dependence of PageRank (CheiRank) probability $P(K)$ ($P^*(K^*)$) on its index $K$ ($K^*$) shown by black (red) curve. \emph{Right panel:} dependence of ImpactRank probability $P$ ($P^*$) on its index $K$ ($K^*$), obtained via propagator of $G$ ($G^*$) at $\alpha=0.85$ and $\gamma= 0.7$ for the initial probability located on neuron RMGL (see text).} \label{fig3} \end{center} \end{figure} The correlations between PageRank and CheiRank vectors is convenient to characterize by the correlator $\kappa=N\sum_i P(i) P^*(i) -1=0.125$. For C.elegans network the value of correlator is relatively small compared to those found for Wikipedia ($\kappa \approx 4$) and WWW of UK universities ($\kappa \sim 3$) \cite{2dmotor}. In a sense for C.elegans neural network the situation if more similar to the networks of Linux Kernel ($\kappa \approx 0$) \cite{alik} and BMP ($\kappa =0.164$) \cite{abel}. Thus, the C.elegans network has practically no correlations between ingoing and outgoing links. It is argued in \cite{alik,2dmotor} that such a network structure allows to perform a control of information flow in a more efficient way. Namely, it allows to reduce the propagation of errors in software codes. It seems that the neural networks also adopt such a structure. \begin{figure}[!ht] \begin{center} \includegraphics[width=7.5cm]{fig4.eps} \caption {(Colour on-line) PageRank - CheiRank plane $(K,K^*)$ showing distribution of neurons according to their ranking. \emph{Left panel :} soma region coloration - head (red), middle (green), tail (blue). \emph{Right panel : } neuron type coloration - sensory (red), motor (green), interneuron (blue), polymodal (purple) and unknown (black). The classifications and colors are given according to WormAtlas \cite{wormatlas}.} \label{fig4} \end{center} \end{figure} Each neuron $i$ belongs to two ranks $K_i$ and $K^*_i$ and it is convenient to represent the distribution of neurons on the two-dimensional plane (2D) of PageRank-CheiRank indexes $(K,K^*)$ shown in Fig.~\ref{fig4}. The plot confirms that there are little correlations between both ranks since the points are scattered over the whole plane. Neurons ranked at top $K$ positions of PageRank have their soma located mainly in both extremities of the worm (head and tail) showing that neurons in those regions have important connections coming from many other neurons which control head and tail movements. This tendency is even more visible for neurons at top $K^*$ positions of CheiRank but with a preference for head and middle regions. In general, neurons, that have their soma in the middle region of the worm, are quite highly ranked in CheiRank but not in PageRank. The neurons located at the head region have top positions in CheiRank and also PageRank, while the middle region has some top CheiRank indexes but rather large indexes of PagRank (Fig.~\ref{fig4} left panel). The neuron type coloration (Fig.~\ref{fig4} right panel) also reveals that sensory neurons are at top PageRank positions but at rather large CheiRank indexes, whereas in general motor neurons are in the opposite situation. \begin{table}[!ht] \caption{Top twenty neurons of PageRank (PR), CheiRank (CR); ImpactRank of $G$ (IMPR) and $G^*$ (IMCR) at initial state RMGL at $\gamma=0.7$; following \cite{wormatlas}, the colors mark: interneurons (blue \emph{bu}), motor neurons (green \emph{gn}), sensory neurons (red \emph{rd}), polymodal neurons (purple \emph{pu}).} \begin{center} \resizebox{0.9\columnwidth}{!}{ \begin{tabular}{|c|c|c||c|c|} \hline & PR & CR & IMPR & IMCR \\ \hline \hline 1 & AVAR \emph{(bu)} & AVAL \emph{(bu)} & RMGL \emph{(bu)} & RMGL \emph{(bu)}\\ 2 & AVAL \emph{(bu)} & AVAR \emph{(bu)} & URXL \emph{(bu)} & AVAL \emph{(bu)}\\ 3 & PVCR \emph{(bu)} & AVBR \emph{(bu)} & ADEL \emph{(rd)} & ASHL \emph{(rd)}\\ 4 & RIH \emph{(bu)} & AVBL \emph{(bu)} & AIAL \emph{(bu)} & AVBR \emph{(bu)}\\ 5 & AIAL \emph{(bu)} & DD02 \emph{(gn)}& IL2L \emph{(rd)} & URXL \emph{(bu)}\\ 6 & PHAL \emph{(rd)} & VD02 \emph{(gn)}& ADLL \emph{(rd)} & AVEL \emph{(bu)}\\ 7 & PHAR \emph{(rd)} & DD01 \emph{(gn)}& PVQL \emph{(bu)} & RIBL \emph{(bu)}\\ 8 & ADEL \emph{(rd)} & RIBL \emph{(bu)} & ALML \emph{(rd)} & RMDR \emph{(pu)}\\ 9 & HSNR \emph{(gn)}& RIBR \emph{(bu)} & ASKL \emph{(rd)} & RMDL \emph{(pu)}\\ 10 & RMGR\emph{(bu)} & VD04 \emph{(gn)}& CEPDL\emph{(rd)} & RMDVL \emph{(pu)}\\ 11 & VC03\emph{(gn)}& VD03 \emph{(gn)}& ASHL \emph{(rd)} & AVAR \emph{(bu)}\\ 12 & AIAR\emph{(bu)} & VD01 \emph{(gn)}& AWBL \emph{(rd)} & SIBVR\emph{(bu)}\\ 13 & AVBL\emph{(bu)} & AVER \emph{(bu)} & SAADR\emph{(bu)} & AIBR \emph{(bu)}\\ 14 & PVPL\emph{(bu)} & RMEV \emph{(gn)}& RMHR \emph{(gn)} & ADAL \emph{(bu)}\\ 15 & AVM \emph{(rd)} & RMDVR\emph{(pu)} & RMHL \emph{(gn)} & RMHL \emph{(gn)}\\ 16 & AVKL\emph{(bu)} & AVEL \emph{(bu)} & RIH \emph{(bu)} & AVBL \emph{(bu)}\\ 17 & HSNL\emph{(gn)}& VD05 \emph{(gn)}& OLQVL\emph{(pu)}& SIBVL\emph{(bu)}\\ 18 & RMGL\emph{(bu)} & SMDDR\emph{(pu)} & AIML \emph{(bu)} & ASKL \emph{(rd)}\\ 19 & AVHR\emph{(bu)} & DD03 \emph{(gn)}& HSNL \emph{(gn)} & RID \emph{(bu)}\\ 20 & AVFL\emph{(bu)} & VA02 \emph{(gn)}& SDQR \emph{(bu)} & SMBVL\emph{(pu)}\\ \hline \end{tabular}} \end{center} \label{table1} \end{table} The top $20$ neurons of PageRank and CheiRank vectors are given in the first two columns of Table~\ref{table1}. We note that both rankings favor important signal relaying neurons such as $AVA$ and $AVB$ that integrate signals from crucial nodes and in turn pilot other crucial nodes. Neurons $AVAL,AVAR$, $AVBL,AVBR$ and $AVEL,AVER$ are considered to belong to the rich club analyzed in\cite{towlson}. The right panel in Fig.~\ref{fig3} and second two columns of Table~\ref{table1} represent ImpactRank which is discussed below. \begin{figure}[!ht] \begin{center} \includegraphics[width=7.5cm]{fig5.eps} \caption {(Colour on-line) Distribution of neurons in the plane $(K,K^*)$ of equal opportunity ranks (see text); colors are the same as in Fig.~\ref{fig4}. } \label{fig5} \end{center} \end{figure} We can also use 2DRank index $K_2$, discussed in \cite{zzs}, which counts nodes in order of their appearance on ribs of squares in $(K,K^*)$ plane with the square size growing from $K=1$ to $K=N$. The top neurons in $K_2$ are AVAL, AVAR, AVBL, AVBR, PVCR. Thus at the top $K_2$ values we find dominance of interneurons. More detailed listings are available at \cite{wormgooglematrix}. It may be also useful to consider renormalized equal opportunity rank recently discussed in \cite{banky}. In this approach PageRank probability of node $i$ is replaced by $P(i)/d(i)$ where $d(i)$ is in-degree of node $i$. For the Google matrix this recipe should be replaced by $P(i) \rightarrow P(i)/\sum_j G_{ij}$ and respectively for CheiRank by $P^*(i) \rightarrow P^*(i)/\sum_j G^*_{ij}$. The corresponding rank indexes $K,K^*$ rank the neurons in the decreasing order of these renormalized probabilities. The distribution of nodes in the plane $(K,K^*)$ is shown in Fig.~\ref{fig5}. In this ranking the top $K$ nodes correspond to important sensory neurons rather than information relaying centers, whereas the top nodes of $K^*$ are composed mainly by motor neurons. Thus such an approach allows to highlight additional features of C.elegans network being complementary to PageRank and CheiRank properties discussed above. Tables for neuron renormalized ranking are available at \cite{wormgooglematrix}. \section{ImpactRank} In certain cases it is useful to determine an influence or impact of a given neuron on other neurons. A recent proposal of ImpactRank, described in \cite{physrev}, is based on the probability distribution of a vector $v_f =(1-\gamma) (1- \gamma G)^{-1} v_0$, $v^*_f =(1-\gamma) (1- \gamma G^*)^{-1} v_0$, where $v_0$ is initially populated neuron. The vector $v_f$ can be viewed as a Green function propagator. The computation of $v_f$ is obtained numerically by a summation of geometrical expansion series which are convergent within approximately first $200$ terms at $\gamma \sim 0.7$ (see also \cite{physrev}). The distributions of probabilities of ImpactRank $P(i)=v_f(i)$, $P^*(i)=v^*_f(i)$ versus the corresponding ImpactRank indexes $K,K^*$ are shown in Fig.~\ref{fig3} (right panel) for the initial state neuron $RMGL$. The corresponding top $20$ ImpactRank neurons influenced by $RMGL$ are given in columns $IMPR$, $IMCR$ of Table~\ref{table1}. The analysis of neurons linked to $RMGL$ shows that indeed, ImpactRank correctly selects neurons influenced by $RMGL$. The neurons in the top list of $P(i)$ are those pointed by outgoing links of $RMGL$ while those in the top list of $P^*(i)$ are those that have ingoing links to $RMGL$. Such a method can be easily applied to other initial neuron states of interest showing a contamination propagation over the neural network starting from initial neuron $RMGL$. \section{Properties of Eigenstates} The Google matrix analysis of the Wikipedia hyperlink network \cite{wikispectrum} demonstrated that the eigenstates with large values of $|\lambda|$ select well defined communities. Thus we can expect that other eigenstates of matrices $G$ and $G^*$ with $|\lambda| <1$ correspond to certain physiological functions of worm neural network. It is convenient to order index of eigenstates in a decreasing order of $|\lambda_i|$ with $\lambda_1=1$. \begin{table}[!ht] \caption{Top ten neurons of the eigenvectors of $G$ (left panel) and $G^*$ (right panel) corresponding to the 10th largest eigenvalues $|\lambda|$; IPR are respectively $\xi \approx 5$ and $\xi \approx 4$.} \begin{center} \resizebox{7.5cm}{!}{ \begin{tabular}{|c|c|c|} \hline & $\lambda_{10}=-0.49446$ & $|\psi_i|$ \\ \hline \hline 1 & AIAR & 0.11986 \\ 2 & AIAL & 0.11159 \\ 3 & ASIL & 0.096475 \\ 4 & ASIR & 0.096236 \\ 5 & AWAR & 0.024228 \\ 6 & ASHR & 0.022241 \\ 7 & RMGR & 0.018502 \\ 8 & AIMR & 0.018387 \\ 9 & ADLL & 0.01837 \\ 10 & PVQL & 0.017547 \\ \hline \end{tabular} \begin{tabular}{|c|c|c|} \hline & $\lambda_{10}=-0.45784$ & $|\psi^*_i|$ \\ \hline \hline 1 & AVAL & 0.10651 \\ 2 & AVAR & 0.079403 \\ 3 & AVBR & 0.036779 \\ 4 & VD05 & 0.025086 \\ 5 & VA09 & 0.02438 \\ 6 & VD06 & 0.020977 \\ 7 & VA08 & 0.020242 \\ 8 & AVBL & 0.019225 \\ 9 & DD02 & 0.018684 \\ 10 & PDB & 0.016485 \\ \hline \end{tabular}} \end{center} \label{table2} \end{table} \begin{table}[!ht] \caption{Same as in Table~\ref{table2} for $48th$ largest eigenvalue modulus $|\lambda|$; IPR are respectively $\xi \approx 54$ and $\xi \approx 47$.} \begin{center} \resizebox{7.5cm}{!}{ \begin{tabular}{|c|c|c|} \hline & $\lambda_{48}=-0.30615-0.07037i$ & $|\psi_i|$ \\ \hline \hline 1 & RIH & 0.017854 \\ 2 & BDUR & 0.017737 \\ 3 & OLLR & 0.016701 \\ 4 & CEPDR & 0.016463 \\ 5 & RMGR & 0.016357 \\ 6 & AIAL & 0.016072 \\ 7 & ASHR & 0.015585 \\ 8 & VC04 & 0.015265 \\ 9 & ASKR & 0.014 \\ 10 & IL2R & 0.013978 \\ \hline \end{tabular} \begin{tabular}{|c|c|c|} \hline & $\lambda_{48}=0.26353-0.095716i$ & $|\psi^*_i|$ \\ \hline \hline 1 & RMEV & 0.026461 \\ 2 & RIBR & 0.013343 \\ 3 & OLQDR & 0.013145 \\ 4 & IL1DL & 0.012932 \\ 5 & IL1DR & 0.012911 \\ 6 & RIAR & 0.012896 \\ 7 & RICR & 0.012728 \\ 8 & OLQDL & 0.012586 \\ 9 & RIGR & 0.012256 \\ 10 & SMDDR & 0.011958 \\ \hline \end{tabular} } \end{center} \label{table3} \end{table} The top ten neurons in eigenfunction amplitude for four specific eigenstates of $G$ and $G^*$ are given in Table~\ref{table2}, Table~\ref{table3}. In Table~\ref{table2} we have eigenstates with low value of IPR so that the corresponding wavefunctions are localized essentially on only about 4 neurons being $AIAR$, $AIAL$, $ASISL$, $ASIR$ and $AVAL$, $AVAR$, $AVBR$ for $\lambda_{10}$ of $G$ and $G^*$ respectively. In Table~\ref{table3} the values of IPR are rather large and these eigenstates are spread over many neurons. \begin{figure}[!ht] \begin{center} \includegraphics[width=7.5cm]{fig6.eps} \caption {(Colour on-line) Dependence of amplitude of eigenstates $|\psi_i(K)|$ of $G$ and $|\psi^*_i(K^*)|$ of $G^*$ on PageRank index $K$ (left panel) and CheiRank index $K^*$ (right panel); here $x$-axis shows values of $K$ and $K^*$, while $y$-axis shows index $i$ of eigenstates being ordered in a decreasing order of $|\lambda_i|$ (see text). The whole index range $1 \leq K, K^* \leq 279$ is shown with PageRank (CheiRank) vector being at the bottom line of each panel. The color is proportional to $|\psi_i(j)|$ changing from minimum blue value to maximum value in red.} \label{fig6} \end{center} \end{figure} To determine if some eigenvectors are localized on a certain group of neurons, we plot in Fig.~\ref{fig6} the amplitude of each eigenstate horizontally in the basis of neurons ordered by indexes of $K$ and $K^*$ of PageRank and CheiRank vectors. The eigenstates of $G$ matrix show four distinct vertical stripes at $K=149, K=165, K=185, K=261$ which correspond respectively to neurons $PVDR$, $IL2DR$, $IL2DL$, $PLNR$. The same plot for $G^*$ matrix shows a larger number of stripes which have less pronounced amplitudes. These stripes of $G^*$ are located on the following neurons $K^*=116 ( RIPL)$, $K^*=123 (RIPR)$, $K^*=120 (AS07)$, $K^*=122 (AS10)$, $K^*=135 (DB06)$, $K^*=137 (DB05)$, $K^*=215 (DA07)$, $K^*=162 (VA10)$, $K^*=172 (SIADL)$, $K^*=181 (SIAVL)$, $K^*=199 (SIAVR)$, $K^*=221 (SIADR)$. We think that an identification of eigenstates with certain physiological functions of worm can be an interesting task which however requires further more detailed studies in collaboration with physiologists. The tables of top 20 nodes of eigenstates with 50 largest $|\lambda_i|$ values are available at \cite{wormgooglematrix}. \section{Discussion} In this Letter, we analyzed the neural network of C.Elegans using Google matrix approach to directed networks which proved its efficiency for the WWW. We classify worm neurons using PageRank and CheiRank probabilities corresponding to the principal vectors of the Google matrix with direct and inverted links. Thus neurons in the head region take top positions in PageRank, CheiRank and combined 2DRank. Also interneurons occupy top ranking positions. We show that influences and interdependency between neurons can be studied using the ImpactRank propagator. We argue that the eigenvectors with large modulus of eigenvalues of the Google matrix may select specific physiological functions. This conjecture still need to be investigated in more detailed studies. Our research shows that the Google matrix analysis represents a useful and powerful method of neural network analysis. We thank Emma K.\ Towlson and Petra E.\ V\'ertes for useful discussions and for providing us the links between neurons available from C.elegans neural network data set at \cite{wormatlas}. This work is supported in part by EC FET Open project "New tools and algorithms for directed network analysis" (NADINE No 288956).
\section{Introduction} Compressive sensing is a recently emerged technique of signal sampling and reconstruction, the main purpose of which is to recover sparse signals from much fewer linear measurements \cite{ChenDonoho98,CandesTao05,Donoho06} \begin{align} \boldsymbol{y}=\boldsymbol{Ax} \end{align} where $\boldsymbol{A}\in\mathbb{R}^{m\times n}$ is the sampling matrix with $m\ll n$, and $\boldsymbol{x}$ denotes the $n$-dimensional sparse signal with only $K$ nonzero coefficients. Such a problem has been extensively studied and a variety of algorithms that provide consistent recovery performance guarantee were proposed, e.g. \cite{ChenDonoho98,CandesTao05,Donoho06,TroppGilbert07,Wainwright09,DaiMilenkovic09}. In practice, sparse signals usually have additional structures that can be exploited to enhance the recovery performance. For example, the atomic decomposition of multi-band signals \cite{Mishali09} or audio signals \cite{GribonvalBacry03} usually results in a block-sparse structure in which the nonzero coefficients occur in clusters. In addition, a discrete wavelet transform of an image naturally yields a tree structure of the wavelet coefficients, with each wavelet coefficient serving as a ``parent'' for a few ``children'' coefficients \cite{HeCarin09}. A number of algorithms, e.g., block-OMP \cite{EldarKuppinger10}, mixed $\ell_2/\ell_1$ norm-minimization \cite{EldarMishali09}, group LASSO \cite{YuanLin06}, StructOMP \cite{HuangZhang11}, and model-based CoSaMP \cite{BaraniukCevher10} were proposed for recovery of block-sparse signals, and their recovery behaviors were analyzed in terms of the model-based restricted isometry property (RIP) \cite{EldarMishali09,BaraniukCevher10} and the mutual coherence \cite{EldarKuppinger10}. Analyses suggested that exploiting the inherent structure of sparse signals helps improve the recovery performance considerably. These algorithms, albeit effective, require the knowledge of the block structure (such as locations and sizes of blocks) of sparse signals \emph{a priori}. In practice, however, the prior information about the block structure of sparse signals is often unavailable. For example, we know that images have structured sparse representations but the exact tree structure of the coefficients is unknown to us. To address this difficulty, a hierarchical Bayesian ``spike-and-slab'' prior model is introduced in \cite{HeCarin09,YuSun12} to encourage the sparseness and promote the cluster patterns simultaneously. Nevertheless, for both works \cite{HeCarin09,YuSun12}, the posterior distribution cannot be derived analytically, and a Markov chain Monte Carlo (MCMC) sampling method has to be employed for Bayesian inference. In \cite{PelegEldar12,DremeauHerzet12}, a graphical prior, also referred to as the ``Boltzmann machine'', was used to model the statistical dependencies between atoms. Specifically, the Boltzmann machine is employed as a prior on the support of a sparse representation. However, the maximum a posterior (MAP) estimator with such a prior involves an exhaustive search over all possible sparsity patterns. To overcome the intractability of the combinatorial search, a greedy method \cite{PelegEldar12} and a variational mean-field approximation method \cite{DremeauHerzet12} were proposed to approximate the MAP. Recently, a sparse Bayesian learning method was proposed in \cite{ZhangRao13} to address the sparse signal recovery problem when the block structure is unknown. In \cite{ZhangRao13}, the components of the signal are partitioned into a number of overlapping blocks and each block is assigned a Gaussian prior. An expanded model is then used to convert the overlapping structure into a block diagonal structure so that the conventional block sparse Bayesian learning algorithm can be readily applied. In this paper, we develop a new Bayesian method for block-sparse signal recovery when the block-sparse patterns are entirely unknown. Similar to the conventional sparse Bayesian learning approach \cite{Tipping01,JiXue08}, a Bayesian hierarchical Gaussian framework is employed to model the sparse prior, in which a set of hyperparameters are introduced to characterize the Gaussian prior and control the sparsity of the signal components. Conventional sparse learning approaches, however, assume independence between the elements of the sparse signal. Specifically, each individual hyperparameter is associated independently with each coefficient of the sparse signal. To model the block-sparse patterns, in this paper, we propose a coupled hierarchical Gaussian framework in which the sparsity of each coefficient is controlled not only by its own hyperparameter, but also by the hyperparameters of its immediate neighbors. Such a prior encourages clustered patterns and suppresses ``isolated coefficients'' whose pattern is different from that of its neighboring coefficients. An expectation-maximization (EM) algorithm is developed to learn the hyperparameters characterizing the coupled hierarchical model and to estimate the block-sparse signal. Our proposed algorithm not only admits a simple iterative procedure for Bayesian inference. It also demonstrates superiority over other existing methods for block-sparse signal recovery. The rest of the paper is organized as follows. In Section \ref{sec:model}, we introduce a new coupled hierarchical Gaussian framework to model the sparse prior and the dependencies among the signal components. An expectation-maximization (EM) algorithm is developed in Section \ref{sec:inference} to learn the hyperparameters characterizing the coupled hierarchical model and to estimate the block-sparse signal. Section \ref{sec:inference-un} extends the proposed Bayesian inference method to the scenario where the observation noise variance is unknown. Relation of our work to other existing works are discussed in \ref{sec:discussions}, and an iterative reweighted algorithm is proposed for the recovery of block-sparse signals. Simulation results are provided in Section \ref{sec:simulation}, followed by concluding remarks in Section \ref{sec:conclusion}. \section{Hierarchical Prior Model} \label{sec:model} We consider the problem of recovering a block-sparse signal $\boldsymbol{x}\in\mathbb{R}^{n}$ from noise-corrupted measurements \begin{align} \boldsymbol{y}=\boldsymbol{A}\boldsymbol{x}+\boldsymbol{w} \end{align} where $\boldsymbol{A}\in\mathbb{R}^{m\times n}$ ($m<n$) is the measurement matrix, and $\boldsymbol{w}$ is the additive multivariate Gaussian noise with zero mean and covariance matrix $\sigma^2\boldsymbol{I}$. The signal $\boldsymbol{x}$ has a block-sparse structure but the exact block pattern such as the location and size of each block is unavailable to us. In the conventional sparse Bayesian learning framework, to encourage the sparsity of the estimated signal, $\boldsymbol{x}$ is assigned a Gaussian prior distribution \begin{align} p(\boldsymbol{x}|\boldsymbol{\alpha})=\prod_{i=1}^n p(x_i|\alpha_i) \end{align} where $p(x_i|\alpha_i)=\mathcal{N}(x_i|0,\alpha_i^{-1})$, and $\boldsymbol{\alpha}\triangleq\{\alpha_i\}$ are non-negative hyperparameters controlling the sparsity of the signal $\boldsymbol{x}$. Clearly, when $\alpha_i$ approaches infinity, the corresponding coefficient $x_i$ becomes zero. By placing hyperpriors over $\{\alpha_i\}$, the hyperparameters $\{\alpha_i\}$ can be learned by maximizing their posterior probability. We see that in the above conventional hierarchical Bayesian model, each hyperparameter is associated independently with each coefficient. The prior model assumes independence among coefficients and has no potential to encourage clustered sparse solutions. To exploit the statistical dependencies among coefficients, we propose a new hierarchical Bayesian model in which the prior for each coefficient not only involves its own hyperparameter, but also the hyperparameters of its immediate neighbors. Specifically, a prior over $\boldsymbol{x}$ is given by \begin{align} p(\boldsymbol{x}|\boldsymbol{\alpha})=&\prod_{i=1}^n p(x_i|\alpha_i,\alpha_{i+1},\alpha_{i-1}) \label{eq3} \end{align} where \begin{align} p(x_i|\alpha_i,\alpha_{i+1},\alpha_{i-1})= \mathcal{N}(x_i|0,(\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1})^{-1}) \label{prior} \end{align} and we assume $\alpha_0=0$ and $\alpha_{n+1}=0$ for the end points $x_1$ and $x_n$, $0\leq \beta\leq 1$ is a parameter indicating the relevance between the coefficient $x_i$ and its neighboring coefficients $\{x_{i+1},x_{i-1}\}$. To better understand this prior model, we can rewrite (\ref{prior}) as \begin{align} p(x_i|\alpha_i,\alpha_{i+1},\alpha_{i-1})\propto p(x_i|\alpha_i)[p(x_i|\alpha_{i+1})]^{\beta}[p(x_i|\alpha_{i-1})]^{\beta} \label{eq2} \end{align} where $p(x_i|\alpha_j)=\mathcal{N}(x_i|0,\alpha_j^{-1})$ for $j=i,i+1,i-1$. We see that the prior for $x_i$ is proportional to a product of three Gaussian distributions, with the coefficient $x_i$ associated with one of the three hyperparameters $\{\alpha_i,\alpha_{i+1},\alpha_{i-1}\}$ for each distribution. When $\beta=0$, the prior distribution (\ref{eq2}) reduces to the prior for the conventional sparse Bayesian learning. When $\beta>0$, the sparsity of $x_i$ not only depends on the hyperparameter $\alpha_i$, but also on the neighboring hyperparameters $\{\alpha_{i+1},\alpha_{i-1}\}$. Hence it can be expected that the sparsity patterns of neighboring coefficients are related to each other. Also, such a prior does not require the knowledge of the block-sparse structure of the sparse signal. It naturally has the tendency to suppress isolated non-zero coefficients and encourage structured-sparse solutions. Following the conventional sparse Bayesian learning framework, we use Gamma distributions as hyperpriors over the hyperparameters $\{\alpha_i\}$, i.e. \begin{align} p(\boldsymbol{\alpha})=\prod_{i=1}^n\text{Gamma}(\alpha_i|a,b)=\prod_{i=1}^n\Gamma(a)^{-1}b^a\alpha^{a}e^{-b\alpha} \label{alpha-prior} \end{align} where $\Gamma(a)=\int_{0}^{\infty}t^{a-1}e^{-t}dt$ is the Gamma function. The choice of the Gamma hyperprior results in a learning process which tends to switch off most of the coefficients that are deemed to be irrelevant, and only keep very few relevant coefficients to explain the data. This mechanism is also called as ``automatic relevance determination''. In the conventional sparse Bayesian framework, to make the Gamma prior non-informative, very small values, e.g. $10^{-4}$, are assigned to the two parameters $a$ and $b$. Nevertheless, in this paper, we use a more favorable prior which sets a larger $a$ (say, $a=1$) in order to achieve the desired ``pruning'' effect for our proposed hierarchical Bayesian model. Clearly, the Gamma prior with a larger $a$ encourages large values of the hyperparameters, and therefore promotes the sparseness of the solution since the larger the hyperparameter, the smaller the variance of the corresponding coefficient. \section{Proposed Bayesian Inference Algorithm} \label{sec:inference} We now proceed to develop a sparse Bayesian learning method for block-sparse signal recovery. For ease of exposition, we assume that the noise variance $\sigma^2$ is known \emph{a priori}. Extension of the Bayesian inference to the case of unknown noise variance will be discussed in the next section. Based on the above hierarchical model, the posterior distribution of $\boldsymbol{x}$ can be computed as \begin{align} p(\boldsymbol{x}|\boldsymbol{\alpha},\boldsymbol{y})\propto& p(\boldsymbol{x}|\boldsymbol{\alpha})p(\boldsymbol{y}|\boldsymbol{x}) \label{x-posterior} \end{align} where $\boldsymbol{\alpha}\triangleq\{\alpha_i\}$, $p(\boldsymbol{x}|\boldsymbol{\alpha})$ is given by (\ref{eq3}), and \begin{align} p(\boldsymbol{y}|\boldsymbol{x})=& \frac{1}{(\sqrt{2\pi\sigma^2})^{m}}\exp\bigg(-\frac{\|\boldsymbol{y}-\boldsymbol{A}\boldsymbol{x}\|_2^2}{2\sigma^2}\bigg) \end{align} It can be readily verified that the posterior $p(\boldsymbol{x}|\boldsymbol{\alpha},\boldsymbol{y})$ follows a Gaussian distribution with its mean and covariance given respectively by \begin{align} \boldsymbol{\mu}=&\sigma^{-2}\boldsymbol{\Phi}\boldsymbol{A}^T\boldsymbol{y} \nonumber\\ \boldsymbol{\Phi}=&(\sigma^{-2}\boldsymbol{A}^T\boldsymbol{A}+\boldsymbol{D})^{-1} \label{eq4} \end{align} where $\boldsymbol{D}$ is a diagonal matrix with its $i$th diagonal element equal to $(\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1})$, i.e. \begin{align} \boldsymbol{D}\triangleq\text{diag}(\alpha_1+\beta\alpha_{2}+\beta\alpha_{0},\ldots, \alpha_{n}+\beta\alpha_{n-1}+\beta\alpha_{n+1}) \label{D-definition} \end{align} Given a set of estimated hyperparameters $\{\alpha_i\}$, the maximum a posterior (MAP) estimate of $\boldsymbol{x}$ is the mean of its posterior distribution, i.e. \begin{align} \boldsymbol{\hat{x}}_{\text{MAP}}=\boldsymbol{\mu}= (\boldsymbol{A}^T\boldsymbol{A}+\sigma^{2}\boldsymbol{D})^{-1} \boldsymbol{A}^T\boldsymbol{y} \label{posterior-mean} \end{align} Our problem therefore reduces to estimating the set of hyperparameters $\{\alpha_i\}$. With hyperpriors placed over $\alpha_{i}$, learning the hyperparameters becomes a search for their posterior mode, i.e. maximization of the posterior probability $p(\boldsymbol{\alpha}|\boldsymbol{y})$. A strategy to maximize the posterior probability is to exploit the expectation-maximization (EM) formulation which treats the signal $\boldsymbol{x}$ as the hidden variables and maximizes the expected value of the complete log-posterior of $\boldsymbol{\alpha}$, i.e. $E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}}[\log p(\boldsymbol{\alpha}|\boldsymbol{x})]$, where the operator $E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}}[\cdot]$ denotes the expectation with respect to the distribution $p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha})$. Specifically, the EM algorithm produces a sequence of estimates $\boldsymbol{\alpha}^{(t)}$, $t=1,2,3,\ldots$, by applying two alternating steps, namely, the E-step and the M-step \cite{FangJi08}. \textbf{E-Step:} Given the current estimates of the hyperparameters $\boldsymbol{\alpha}^{(t)}$ and the observed data $\boldsymbol{y}$, the E-step requires computing the expected value (with respect to the missing variables $\boldsymbol{x}$) of the complete log-posterior of $\boldsymbol{\alpha}$, which is also referred to as the Q-function; we have \begin{align} Q(\boldsymbol{\alpha}|\boldsymbol{\alpha}^{(t)})=& E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)}}[\log p(\boldsymbol{\alpha}|\boldsymbol{x})] \nonumber\\ =&\int p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)})\log p(\boldsymbol{\alpha}|\boldsymbol{x}) d\boldsymbol{x} \nonumber\\ =&\int p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)})\log [p(\boldsymbol{\alpha})p(\boldsymbol{x}|\boldsymbol{\alpha})]d\boldsymbol{x}+ c \label{Q-function-org} \end{align} where $c$ is a constant independent of $\boldsymbol{\alpha}$. Ignoring the term independent of $\boldsymbol{\alpha}$, and recalling (\ref{eq3}), the Q-function can be re-expressed as \begin{align} Q(\boldsymbol{\alpha}|\boldsymbol{\alpha}^{(t)}) =& \log p(\boldsymbol{\alpha})+\frac{1}{2}\sum_{i=1}^n\bigg(\log (\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1}) \nonumber\\ &-(\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1})\int p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)})x_i^2 d\boldsymbol{x}\bigg) \end{align} Since the posterior $p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)})$ is a multivariate Gaussian distribution with its mean and covariance matrix given by (\ref{eq4}), we have \begin{align} \int p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)})x_i^2 d\boldsymbol{x}=E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)}}\left[x_i^2\right]=\hat{\mu}_i^2+\hat{\phi}_{i,i} \end{align} where $\hat{\mu}_i$ denotes the $i$th entry of $\boldsymbol{\hat{\mu}}$, $\hat{\phi}_{i,i}$ denotes the $i$th diagonal element of the covariance matrix $\boldsymbol{\hat{\Phi}}$, $\boldsymbol{\hat{\mu}}$ and $\boldsymbol{\hat{\Phi}}$ are computed according to (\ref{eq4}), with $\boldsymbol{\alpha}$ replaced by the current estimate $\boldsymbol{\alpha}^{(t)}$. With the specified prior (\ref{alpha-prior}), the Q-function can eventually be written as \begin{align} &Q(\boldsymbol{\alpha}|\boldsymbol{\alpha}^{(t)})\nonumber\\ =&\sum_{i=1}^n\bigg(a\log\alpha_i-b\alpha_i+\frac{1}{2}\log (\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1}) \nonumber\\ &-\frac{1}{2}(\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1})(\hat{\mu}_i^2+\hat{\phi}_{i,i})\bigg) \label{Q-function} \end{align} \textbf{M-Step:} In the M-step of the EM algorithm, a new estimate of $\boldsymbol{\alpha}$ is obtained by maximizing the Q-function, i.e. \begin{align} \boldsymbol{\alpha}^{(t+1)} = \arg\max_{\boldsymbol{\alpha}}Q(\boldsymbol{\alpha}|\boldsymbol{\alpha}^{(t)}) \label{M-opt} \end{align} For the conventional sparse Bayesian learning, maximization of the Q-function can be decoupled into a number of separate optimizations in which each hyperparameter $\alpha_i$ is updated independently. This, however, is not the case for the problem being considered here. We see that the hyperparameters in the Q-function (\ref{Q-function}) are entangled with each other due to the logarithm term $\log(\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1})$. In this case, an analytical solution to the optimization (\ref{M-opt}) is difficult to obtain. Gradient descend methods can certainly be used to search for the optimal solution. Nevertheless, such a gradient-based search method, albeit effective, does not provide any insight into the learning process. Also, gradient-based methods involve higher computational complexity as compared with an analytical update rule. To overcome the drawbacks of gradient-based methods, we consider an alternative strategy which aims at finding a simple, analytical sub-optimal solution of (\ref{M-opt}). Such an analytical sub-optimal solution can be obtained by examining the optimality condition of (\ref{M-opt}). Suppose $\boldsymbol{\alpha}^{\ast}$ is the optimal solution of (\ref{M-opt}), then the first derivative of the Q-function with respect to $\boldsymbol{\alpha}$ equals to zero at the optimal point, i.e. \begin{align} \frac{\partial Q(\boldsymbol{\alpha}|\boldsymbol{\alpha}^{(t)})}{\partial\boldsymbol{\alpha}} \bigg |_{\boldsymbol{\alpha}=\boldsymbol{\alpha}{\ast}}=\boldsymbol{0} \end{align} To examine this optimality condition more thoroughly, we compute the first derivative of the Q-function with respect to each individual hyperparameter: \begin{align} \frac{\partial Q(\boldsymbol{\alpha}|\boldsymbol{\alpha}^{(t)})}{\partial\alpha_i} =\frac{a}{\alpha_i}-b-\frac{1}{2}\omega_i +\frac{1}{2}(\nu_i+&\beta\nu_{i+1}+\beta\nu_{i-1})\nonumber\\ & \quad \forall i=1,\ldots,n \end{align} where $\nu_0=0$, $\nu_{n+1}=0$, and for $i=1,\ldots,n$, we have \begin{align} \omega_i\triangleq&(\hat{\mu}_i^2+\hat{\phi}_{i,i})+\beta(\hat{\mu}_{i+1}^2+\hat{\phi}_{i+1,i+1}) +\beta(\hat{\mu}_{i-1}^2+\hat{\phi}_{i-1,i-1}) \label{omega} \\ \nu_i\triangleq&\frac{1}{\alpha_i+\beta\alpha_{i+1}+\beta\alpha_{i-1}} \end{align} Note that for notational convenience, we allow the subscript indices of the notations $\hat{\mu}_i$ and $\hat{\phi}_{i,i}$ in (\ref{omega}) equal to $0$ and $n+1$. Although these notations $\{\hat{\mu}_0, \hat{\phi}_{0,0},\hat{\mu}_{n+1}, \hat{\phi}_{n+1,n+1}\}$ do not have any meaning, they can be used to simplify our expression. Clearly, they should all be set equal to zero, i.e. $\hat{\mu}_0=\hat{\mu}_{n+1}=\hat{\phi}_{0,0}=\hat{\phi}_{n+1,n+1}=0$. Recalling the optimality condition, we therefore have \begin{align} \frac{a}{\alpha_i^{\ast}}+ \frac{1}{2}(\nu_i^{\ast}+\beta\nu_{i+1}^{\ast}+\beta\nu_{i-1}^{\ast}) =b+\frac{1}{2}\omega_i \quad \forall i=1,\ldots,n\label{eq5} \end{align} where $\nu_0^{\ast}=0$, $\nu_{n+1}^{\ast}=0$, and \begin{align} \nu_i^{\ast}\triangleq&\frac{1}{\alpha_i^{\ast}+\beta\alpha_{i+1}^{\ast}+\beta\alpha_{i-1}^{\ast}}\qquad \forall i=1,\ldots,n \nonumber \end{align} Since all hyperparameters $\{\alpha_i\}$ and $\beta$ are non-negative, we have \begin{align} \frac{1}{\alpha_i^{\ast}}>&\nu_i^{\ast}>0 \qquad \forall i=1,\ldots,n \nonumber\\ \frac{1}{\beta\alpha_{i+1}^{\ast}}>&\nu_i^{\ast}>0 \qquad \forall i=1,\ldots,n-1 \nonumber\\ \frac{1}{\beta\alpha_{i-1}^{\ast}}>&\nu_i^{\ast}>0 \qquad \forall i=2,\ldots,n \nonumber \end{align} Hence the term on the left-hand side of (\ref{eq5}) is lower and upper bounded respectively by \begin{align} \frac{a+c_0}{\alpha_i^{\ast}}\geq\frac{a}{\alpha_i^{\ast}}+ \frac{1}{2}(\nu_i^{\ast}+\beta\nu_{i+1}^{\ast}+\beta\nu_{i-1}^{\ast})>\frac{a}{\alpha_i^{\ast}} \label{eq6} \end{align} where $c_0=1.5$ for $i=2,\ldots,n-1$, and $c_0=1$ for $i=\{1,n\}$. Combining (\ref{eq5})--(\ref{eq6}), we arrive at \begin{align} \alpha_i^{\ast}\in \left[\frac{a}{0.5\omega_i+b}, \frac{a+c_0}{0.5\omega_i+b}\right]\quad \forall i=1,\ldots,n \label{eq12} \end{align} With $a=1$, and $b=10^{-4}$, a sub-optimal solution to (\ref{M-opt}) can be obtained as \begin{align} \hat{\alpha}_i=\frac{\kappa}{0.5\omega_i+10^{-4}} \quad \forall i=1,\ldots,n \label{hp-update} \end{align} for some $\kappa$ within the range $1+c_0\geq\kappa\geq 1$. We see that the solution (\ref{hp-update}) provides a simple rule for the hyperparameter update. Also, notice that the update rule (\ref{hp-update}) resembles that of the conventional sparse Bayesian learning work \cite{Tipping01,JiXue08} except that the parameter $\omega_i$ is equal to $\hat{\mu}_i^2+\hat{\phi}_{i,i}$ for the conventional sparse Bayesian learning method, while for our case, $\omega_i$ is a weighted summation of $\hat{\mu}_j^2+\hat{\phi}_{j,j}$ for $j=i-1,i,i+1$. For clarity, we now summarize the EM algorithm as follows. \begin{enumerate} \item At iteration $t$ ($t=0,1,\ldots$): Given a set of hyperparameters $\boldsymbol{\alpha}^{(t)}=\{\alpha_i^{(t)}\}$, compute the mean $\boldsymbol{\hat{\mu}}$ and covariance matrix $\boldsymbol{\hat{\Phi}}$ of the posterior distribution $p(\boldsymbol{x}|\boldsymbol{\alpha}^{(t)},\boldsymbol{y})$ according to (\ref{eq4}), and compute the MAP estimate $\boldsymbol{\hat{x}}^{(t)}$ according to (\ref{posterior-mean}). \item Update the hyperparameters $\boldsymbol{\alpha}^{(t+1)}$ according to (\ref{hp-update}), where $\omega_i$ is given by (\ref{omega}). \item Continue the above iteration until $\|\boldsymbol{\hat{x}}^{(t+1)}-\boldsymbol{\hat{x}}^{(t)}\|_2\leq\epsilon$, where $\epsilon$ is a prescribed tolerance value. \end{enumerate} \emph{Remarks:} Although the above algorithm employs a sub-optimal solution (\ref{hp-update}) to update the hyperparameters in the M-step, numerical results show that the sub-optimal update rule is quite effective and presents similar recovery performance as using a gradient-based search method. This is because the sub-optimal solution (\ref{hp-update}) provides a reasonable estimate of the optimal solution when the parameter $a$ is set away from zero, say, $a=1$. Numerical results also suggest that the proposed algorithm is insensitive to the choice of the parameter $\kappa$ in (\ref{hp-update}) as long as $\kappa$ is within the range $[a, a+c_0]$ for a properly chosen $a$. We simply set $\kappa=a$ in our following simulations. The update rule (\ref{hp-update}) not only admits a simple analytical form which is computationally efficient, it also provides an insight into the EM algorithm. The Bayesian Occam's razor which contributes to the success of the conventional sparse Bayesian learning method also works here to automatically select an appropriate simple model. To see this, note that in the E-step, when computing the posterior mean and covariance matrix, a large hyperparameter $\alpha_i$ tends to suppress the values of the corresponding components $\{\mu_j,\phi_j\}$ for $j=i-1,i,i+1$ (c.f. (\ref{eq4})). As a result, the value of $\omega_i$ becomes small, which in turn leads to a larger hyperparameter $\alpha_i$ (c.f. (\ref{hp-update})). This negative feedback mechanism keeps decreasing most of the entries in $\boldsymbol{\hat{x}}$ until they reach machine precision and become zeros, while leaving only a few prominent nonzero entries survived to explain the data. Meanwhile, we see that each hyperparameter $\alpha_i$ not only controls the sparseness of its own corresponding coefficient $x_i$, but also has an impact on the sparseness of the neighboring coefficients $\{x_{i+1},x_{i-1}\}$. Therefore the proposed EM algorithm has the tendency to suppress isolated non-zero coefficients and encourage structured-sparse solutions. \section{Bayesian Inference: Unknown Noise Variance} \label{sec:inference-un} In the previous section, for simplicity of exposition, we assume that the noise variance $\sigma^2$ is known \emph{a priori}. This assumption, however, may not hold valid in practice. In this section, we discuss how to extend our previously developed Bayesian inference method to the scenario where the noise variance $\sigma^2$ is unknown. For notational convenience, define \begin{align} \gamma\triangleq\sigma^{-2} \nonumber \end{align} Following the conventional sparse Bayesian learning framework \cite{Tipping01}, we place a Gamma hyperprior over $\gamma$: \begin{align} p(\gamma)=\text{Gamma}(\gamma|c,d)=\Gamma(c)^{-1}d^c\gamma^{c}e^{-d\gamma} \label{gamma-prior} \end{align} where the parameters $c$ and $d$ are set to small values, e.g. $c=d=10^{-4}$. As we already derived in the previous section, given the hyperparameters $\boldsymbol{\alpha}$ and the noise variance $\sigma^2$, the posterior $p(\boldsymbol{x}|\boldsymbol{\alpha},\gamma,\boldsymbol{y})$ follows a Gaussian distribution with its mean and covariance matrix given by (\ref{eq4}). The MAP estimate of $\boldsymbol{x}$ is equivalent to the posterior mean. Our problem therefore becomes jointly estimating the hyperparameters $\boldsymbol{\alpha}$ and the noise variance $\sigma^2$ (or equivalently $\gamma$). Again, the EM algorithm can be used to learn these parameters via maximizing their posterior probability $p(\boldsymbol{\alpha},\gamma|\boldsymbol{y})$. The alternating EM steps are briefly discussed below. \textbf{E-Step}: In the E-step, given the current estimates of the parameters $\{\boldsymbol{\alpha}^{(t)},\gamma^{(t)}\}$ and the observed data $\boldsymbol{y}$, we compute the expected value (with respect to the missing variables $\boldsymbol{x}$) of the complete log-posterior of $\{\boldsymbol{\alpha},\gamma\}$, that is, $E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\log p(\boldsymbol{\alpha},\gamma|\boldsymbol{x},\boldsymbol{y})]$, where the operator $E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\cdot]$ denotes the expectation with respect to the distribution $p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)})$. Since \begin{align} p(\boldsymbol{\alpha},\gamma|\boldsymbol{x},\boldsymbol{y})\propto p(\boldsymbol{\alpha})p(\boldsymbol{x}|\boldsymbol{\alpha})p(\gamma)p(\boldsymbol{y}|\boldsymbol{x},\gamma) \end{align} the Q-function can be expressed as a summation of two terms \begin{align} Q(\boldsymbol{\alpha},\gamma|\boldsymbol{\alpha}^{(t)},\gamma^{(t)})=& E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\log p(\boldsymbol{\alpha})p(\boldsymbol{x}|\boldsymbol{\alpha})] \nonumber\\ &+ E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\log p(\gamma)p(\boldsymbol{y}|\boldsymbol{x},\gamma)] \label{eq7} \end{align} where the first term has exactly the same form as the Q-function (\ref{Q-function-org}) obtained in the previous section, except with the known noise variance $\sigma^2$ replaced by the current estimate $(\sigma^{(t)})^2=1/\gamma^{(t)}$, and the second term is a function of the variable $\gamma$. \textbf{M-Step}: We observe that in the Q-function (\ref{eq7}), the parameters $\boldsymbol{\alpha}$ and $\gamma$ to be learned are separated from each other. This allows the estimation of $\boldsymbol{\alpha}$ and $\gamma$ to be decoupled into the following two independent problems: \begin{align} \boldsymbol{\alpha}^{(t+1)} =& \arg\max_{\boldsymbol{\alpha}}E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\log p(\boldsymbol{\alpha})p(\boldsymbol{x}|\boldsymbol{\alpha})] \label{opt-2} \\ \gamma^{(t+1)} =& \arg\max_{\gamma} E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\log p(\gamma)p(\boldsymbol{y}|\boldsymbol{x},\gamma)] \label{opt-3} \end{align} The first optimization problem (\ref{opt-2}) has been thoroughly studied in the previous section, where we provided a simple analytical form (\ref{hp-update}) for the hyperparameter update. We now discuss the estimation of the parameter $\gamma$. Recalling (\ref{gamma-prior}), we have \begin{align} &E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}}[\log p(\gamma)p(\boldsymbol{y}|\boldsymbol{x},\gamma)] \nonumber\\ =&\frac{m}{2}\log\gamma-\frac{\gamma}{2}E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}} \left[\|\boldsymbol{y}-\boldsymbol{A}\boldsymbol{x}\|_2^2\right]+c\log\gamma-d\gamma \label{eq13} \end{align} Computing the first derivative of (\ref{eq13}) with respect to $\gamma$ and setting it equal to zero, we get \begin{align} \frac{1}{\gamma}=\frac{\chi+2d}{m+2c} \label{eq11} \end{align} where \begin{align} \chi\triangleq E_{\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)}} \left[\|\boldsymbol{y}-\boldsymbol{A}\boldsymbol{x}\|_2^2\right] \nonumber \end{align} Note that the posterior $p(\boldsymbol{x}|\boldsymbol{y},\boldsymbol{\alpha}^{(t)},\gamma^{(t)})$ follows a Gaussian distribution with mean $\boldsymbol{\hat{\mu}}$ and covariance matrix $\boldsymbol{\hat{\Phi}}$, where $\boldsymbol{\hat{\mu}}$ and $\boldsymbol{\hat{\Phi}}$ are computed via (\ref{eq4}) with $\gamma$ (i.e. $\sigma^2$) and $\boldsymbol{\alpha}$ replaced by the current estimates $\{\gamma^{(t)},\boldsymbol{\alpha}^{(t)}\}$. Hence $\chi$ can be computed as \begin{align} \chi=&\boldsymbol{y}^T\boldsymbol{y}-2 E[\boldsymbol{x}^T\boldsymbol{A}^T\boldsymbol{y}]+E[\boldsymbol{x}^T\boldsymbol{A}^T \boldsymbol{A}\boldsymbol{x}] \nonumber\\ =&\boldsymbol{y}^T\boldsymbol{y}-2\boldsymbol{\hat{\mu}}^T\boldsymbol{A}^T\boldsymbol{y}+ \boldsymbol{\hat{\mu}}^T\boldsymbol{A}^T \boldsymbol{A}\boldsymbol{\hat{\mu}}+\text{tr}\left(\boldsymbol{\hat{\Phi}}\boldsymbol{A}^T\boldsymbol{A}\right) \nonumber\\ \stackrel{(a)}{=}& \|\boldsymbol{y}-\boldsymbol{A}\boldsymbol{\hat{\mu}}\|_2^2+(\gamma^{(t)})^{-1}\sum_{i=1}^n\rho_i \label{eq10} \end{align} where the last equality $(a)$ follows from \begin{align} \text{tr}\left(\boldsymbol{\hat{\Phi}}\boldsymbol{A}^T\boldsymbol{A}\right)=& \text{tr}\left(\boldsymbol{\hat{\Phi}}\boldsymbol{A}^T\boldsymbol{A}+(\gamma^{(t)})^{-1} \boldsymbol{\hat{\Phi}}\boldsymbol{\hat{D}}-(\gamma^{(t)})^{-1} \boldsymbol{\hat{\Phi}}\boldsymbol{\hat{D}}\right) \nonumber\\ =&(\gamma^{(t)})^{-1}\text{tr}\left(\boldsymbol{\hat{\Phi}}(\gamma^{(t)}\boldsymbol{A}^T\boldsymbol{A}+\boldsymbol{\hat{D}}) - \boldsymbol{\hat{\Phi}}\boldsymbol{\hat{D}}\right) \nonumber\\ =&(\gamma^{(t)})^{-1}\text{tr}\left(\boldsymbol{I}-\boldsymbol{\hat{\Phi}}\boldsymbol{\hat{D}}\right) \nonumber\\ =&(\gamma^{(t)})^{-1}\sum_{i=1}^n\rho_i \end{align} in which $\boldsymbol{\hat{D}}$ is given by (\ref{D-definition}) with $\boldsymbol{\alpha}$ replaced by the current estimate $\boldsymbol{\alpha}^{(t)}$, and \begin{align} \rho_i\triangleq 1-\hat{\phi}_{i,i}(\alpha_{i}^{(t)}+\beta\alpha_{i-1}^{(t)}+\beta\alpha_{i+1}^{(t)}) \qquad \forall i \label{rho} \end{align} Note that $\alpha_{0}^{(t)}$ and $\alpha_{n+1}^{(t)}$ are set to zero when computing $\rho_1$ and $\rho_n$. Substituting (\ref{eq10}) back into (\ref{eq11}), a new estimate of $\gamma$, i.e. the optimal solution to (\ref{opt-3}), is given by \begin{align} \frac{1}{\gamma^{(t+1)}}=\frac{\|\boldsymbol{y}-\boldsymbol{A}\boldsymbol{\hat{\mu}}\|_2^2 +(\gamma^{(t)})^{-1}\sum_{i}\rho_i+2d}{m+2c} \label{gamma-update} \end{align} The above update formula has a similar form as that for the conventional sparse Bayesian learning (c.f. \cite[Equation (50)]{Tipping01}). The only difference lies in that $\{\rho_i\}$ are computed differently: for the conventional sparse Bayesian learning method, $\rho_i$ is computed as $\rho_i=1-\hat{\phi}_{i,i}\alpha_{i}^{(t)}$, while $\rho_i$ is given by (\ref{rho}) for our algorithm. The sparse Bayesian learning algorithm with unknown noise variance is now summarized as follows. \begin{enumerate} \item At iteration $t$ ($t=0,1,\ldots$): given the current estimates of $\boldsymbol{\alpha}^{(t)}$ and $\gamma^{(t)}$, compute the mean $\boldsymbol{\hat{\mu}}$ and the covariance matrix $\boldsymbol{\hat{\Phi}}$ of the posterior distribution $p(\boldsymbol{x}|\boldsymbol{\alpha}^{(t)},\gamma^{(t)},\boldsymbol{y})$ via (\ref{eq4}), and calculate the MAP estimate $\boldsymbol{\hat{x}}^{(t)}$ according to (\ref{posterior-mean}). \item Compute a new estimate of $\boldsymbol{\alpha}$, denoted as $\boldsymbol{\alpha}^{(t+1)}$, according to (\ref{hp-update}), where $\omega_i$ is given by (\ref{omega}); update $\gamma$ via (\ref{gamma-update}), which yields a new estimate of $\gamma$, denoted as $\gamma^{(t+1)}$. \item Continue the above iteration until $\|\boldsymbol{\hat{x}}^{(t+1)}-\boldsymbol{\hat{x}}^{(t)}\|_2\leq\epsilon$, where $\epsilon$ is a prescribed tolerance value. \end{enumerate} \section{Discussions} \label{sec:discussions} \subsection{Related Work} Sparse Bayesian learning is a powerful approach for regression, classification, and sparse representation. It was firstly introduced by Tipping in his pioneering work \cite{Tipping01}, where the regression and classification problem was addressed and a sparse Bayesian learning approach was developed to automatically remove irrelevant basis vectors and retain only a few `relevant' vectors for prediction. Such an automatic relevance determination mechanism and the resulting sparse solution not only effectively avoid the overfitting problem, but also render superior regression and classification accuracy. Later on in \cite{Wipf06,JiXue08}, sparse Bayesian learning was introduced to solve the sparse recovery problem. In a series of experiments, sparse Bayesian learning demonstrated superior stability for sparse signal recovery, and presents uniform superiority over other methods. In \cite{WipfRao07}, sparse Bayesian learning was generalized to solve the simultaneous (block) sparse recovery problem, in which a group of coefficients sharing the same sparsity pattern are assigned a multivariate Gaussian prior parameterized by a common hyperparameter that controls the sparsity of this group of coefficients. Specifically, we have \begin{align} p(\boldsymbol{x}_i|\alpha_i)=\mathcal{N}(0,\alpha_i^{-1}\boldsymbol{I}) \end{align} where $\boldsymbol{x}_i$ denotes the group of coefficients that share a same sparsity pattern, $\alpha_i$ is the hyperparameter controlling the sparsity of $\boldsymbol{x}_i$. In \cite{ZhangRao11}, the above model was further improved to accommodate temporally correlated sources \begin{align} p(\boldsymbol{x}_i|\alpha_i)=\mathcal{N}(0,\alpha_i^{-1}\boldsymbol{B}_i) \end{align} in which $\boldsymbol{B}_i$ is a positive definite matrix that captures the correlation structure of $\boldsymbol{x}_i$. We see that, in both models \cite{WipfRao07,ZhangRao11}, each coefficient is associated with only one sparseness-controlling hyperparameter. This explicit assignment of each coefficient to a certain hyperparameter requires to know the exact block sparsity pattern \emph{a priori}. In contrast, for our hierarchical Bayesian model, each coefficient is associated with multiple hyperparameters, and the hyperparameters are somehow related to each other through their commonly connected coefficients. Such a coupled hierarchical model has the potential to encourage block-sparse patterns, while without imposing any stringent or pre-specified constraints on the structure of the recovered signals. This property enables the proposed algorithm to learn the block-sparse structure in an automatic manner. Recently, Zhang and Rao extended the block sparse Bayesian learning framework to address the sparse signal recovery problem when the block structure is unknown \cite{ZhangRao13}. In their work \cite{ZhangRao13}, the signal $\boldsymbol{x}$ is partitioned into a number of overlapping blocks $\{\boldsymbol{x}_i\}$ with identical block sizes, and each block $\boldsymbol{x}_i$ is assigned a Gaussian prior $p(\boldsymbol{x}_i|\alpha_i)=\mathcal{N}(0,\alpha_i^{-1}\boldsymbol{B}_i)$. To address the overlapping issue, the original data model is converted into an expanded model which removes the overlapping structure by adding redundant columns to the original measurement matrix $\boldsymbol{A}$ and stacking all blocks $\{\boldsymbol{x}_i\}$ to form an augmented vector. In doing this way, the prior for the new augmented vector has a block diagonal form similar as that for the conventional block sparse Bayesian learning. Thus conventional block sparse Bayesian learning algorithms such as \cite{ZhangRao11} can be applied to the expanded model. This overlapping structure provides flexibility in defining a block-sparse pattern. Hence it works well even when the block structure is unknown. A critical difference between our work and \cite{ZhangRao11} is that for our method, a prior is directly placed on the signal $\boldsymbol{x}$, while for the method proposed in \cite{ZhangRao11}, a rigorous formulation of the prior for $\boldsymbol{x}$ is not available, instead, a prior is assigned to the augmented new signal which is constructed by stacking a number of overlapping blocks $\{\boldsymbol{x}_i\}$. \subsection{A Proposed Iterative Reweighted Algorithm} Sparse Bayesian learning algorithms have a close connection with the reweighted $\ell_1$ or $\ell_2$ methods. In fact, a dual-form analysis \cite{WipfNagaranjan10} reveals that sparse Bayesian learning can be considered as a non-separable reweighted strategy solving a non-separable penalty function. Inspired by this insight, we here propose a reweighted $\ell_1$ method for the recovery of block-sparse signals when the block structure of the sparse signal is unknown. Conventional reweighted $\ell_1$ methods iteratively minimize the following weighted $\ell_1$ function (for simplicity, we consider the noise-free case): \begin{align} \min_{\boldsymbol{x}}\quad& \sum_{i=1}^n w_i^{(t)}|x_i|\nonumber\\ \text{s.t.}\quad& \phantom{0}\boldsymbol{Ax}=\boldsymbol{y}\label{opt-2} \end{align} where the weighting parameters are given by $w_i^{(t)}=1/(|x_i^{(t-1)}|+\epsilon),\forall i$, and $\epsilon$ is a pre-specified positive parameter. In a series of experiments \cite{CandesWakin08}, the above iterative reweighted algorithm outperforms the conventional $\ell_1$-minimization method by a considerable margin. The fascinating idea of the iterative reweighted algorithm is that the weights are updated based on the previous estimate of the solution, with a large weight assigned to the coefficient whose estimate is already small and vice versa. As a result, the value of the coefficient which is assigned a large weight tends to be smaller (until become negligible) in the next estimate. This explains why iterative reweighted algorithms usually yield sparser solutions than the conventional $\ell_1$-minimization method. As discussed in our previous section, the basic idea of our proposed sparse Bayesian learning method is to establish a coupling mechanism such that the sparseness of neighboring coefficients are somehow related to each other. With this in mind, we slightly modify the weight update rule of the reweighted $\ell_1$ algorithm as follows \begin{align} w_i^{(t)}=\frac{1}{|x_i^{(t-1)}|+\beta|x_{i+1}^{(t-1)}|+\beta|x_{i-1}^{(t-1)}|+\epsilon} \qquad \forall i \end{align} We see that unlike the conventional update rule, the weight $w_i^{(t)}$ is not only a function of its corresponding coefficient $x_i^{(t-1)}$, but also dependent on the neighboring coefficients $\{x_{i+1}^{(t-1)},x_{i-1}^{(t-1)}\}$. In doing this way, a coupling effect between the sparsity patterns of neighboring coefficients is established. Hence the modified reweighted $\ell_1$-minimization algorithm has the potential to encourage block-sparse solutions. Experiments show that the proposed modified reweighted $\ell_1$ method yields considerably improved results over the conventional reweighted $\ell_1$ method in recovering block-sparse signals. It also serves as a good reference method for comparison with the proposed Bayesian sparse learning approach. \section{Simulation Results} \label{sec:simulation} We now carry out experiments to illustrate the performance of our proposed algorithm, also referred to as the pattern-coupled sparse Bayesian learning (PC-SBL) algorithm, and its comparison with other existing methods. The performance of the proposed algorithm will be examined using both synthetic and real data\footnote{Matlab codes for our algorithm are available at http://www.junfang-uestc.net/}. The parameters $a$ and $b$ for our proposed algorithm are set equal to $a=0.5$ and $b=10^{-4}$ throughout our experiments. \subsection{Synthetic Data} Let us first consider the synthetic data case. In our simulations, we generate the block-sparse signal in a similar way to \cite{ZhangRao13}. Suppose the $n$-dimensional sparse signal contains $K$ nonzero coefficients which are partitioned into $L$ blocks with random sizes and random locations. Specifically, the block sizes $\{B_l\}_{l=1}^L$ can be determined as follows: we randomly generate $L$ positive random variables $\{r_l\}_{l=1}^L$ with their sum equal to one, then we can simply set $B_l=\lceil K r_l\rceil$ for the first $L-1$ blocks and $B_L=K-\sum_{l=1}^{L-1}B_l$ for the last block, where $\lceil x\rceil$ denotes the ceiling operator that gives the smallest integer no smaller than $x$. Similarly, we can partition the $n$-dimensional vector into $L$ super-blocks using the same set of values $\{r_l\}_{l=1}^L$, and place each of the $L$ nonzero blocks into each super-block with a randomly generated starting position (the starting position, however, is selected such that the nonzero block will not go beyond the super-block). Also, in our experiments, the nonzero coefficients of the sparse signal $\boldsymbol{x}$ and the measurement matrix $\boldsymbol{A}\in\mathbb{R}^{m\times n}$ are randomly generated with each entry independently drawn from a normal distribution, and then the sparse signal $\boldsymbol{x}$ and columns of $\boldsymbol{A}$ are normalized to unit norm. Two metrics are used to evaluate the recovery performance of respective algorithms, namely, the normalized mean squared error (NMSE) and the success rate. The NMSE is defined as $\|\boldsymbol{x}-\boldsymbol{\hat{x}}\|_2^2/\|\boldsymbol{x}\|_2^2$, where $\boldsymbol{\hat{x}}$ denotes the estimate of the true signal $\boldsymbol{x}$. The success rate is computed as the ratio of the number of successful trials to the total number of independent runs. A trial is considered successful if the NMSE is no greater than $10^{-4}$. In our simulations, the success rate is used to measure the recovery performance for the noiseless case, while the NMSE is employed to measure the recovery accuracy when the measurements are corrupted by additive noise. We first examine the recovery performance of our proposed algorithm (PC-SBL) under different choices of $\beta$. As indicated earlier in our paper, $\beta$ ($0\leq\beta\leq 1$) is a parameter quantifying the dependencies among neighboring coefficients. Fig. \ref{fig1} depicts the success rates vs. the ratio $m/n$ for different choices of $\beta$, where we set $n=100$, $K=25$, and $L=4$. Results (in Fig. \ref{fig1} and the following figures) are averaged over 1000 independent runs, with the measurement matrix and the sparse signal randomly generated for each run. The performance of the conventional sparse Bayesian learning method (denoted as ``SBL'') \cite{Tipping01} and the basis pursuit method (denoted as ``BP'') \cite{ChenDonoho98,CandesTao05} is also included for our comparison. We see that when $\beta=0$, our proposed algorithm performs the same as the SBL. This is an expected result since in the case of $\beta=0$, our proposed algorithm is simplified as the SBL. Nevertheless, when $\beta>0$, our proposed algorithm achieves a significant performance improvement (as compared with the SBL and BP) through exploiting the underlying block-sparse structure, even without knowing the exact locations and sizes of the non-zero blocks. We also observe that our proposed algorithm is not very sensitive to the choice of $\beta$ as long as $\beta>0$: it achieves similar success rates for different positive values of $\beta$. For simplicity, we set $\beta=1$ throughout our following experiments. \begin{figure}[!t] \centering \includegraphics[width=9cm]{suc-rates-ratio-beta} \caption{Success rates of the proposed algorithm vs. the ratio $m/n$ for different choices of $\beta$.} \label{fig1} \end{figure} Next, we compare our proposed algorithm with some other recently developed algorithms for block-sparse signal recovery, namely, the expanded block sparse Bayesian learning method (EBSBL) \cite{ZhangRao13}, the Boltzman machine-based greedy pursuit algorithm (BM-MAP-OMP) \cite{PelegEldar12}, and the cluster-structured MCMC algorithm (CluSS-MCMC) \cite{YuSun12}. The modified iterative reweighted $\ell_1$ method (denoted as MRL1) proposed in Section \ref{sec:discussions} is also examined in our simulations. Note that all these algorithms were developed without the knowledge of the block-sparse structure. The block sparse Bayesian learning method (denoted as BSBL) developed in \cite{ZhangRao13} is included as well. Although the BSBL algorithm requires the knowledge of the block-sparse structure, it still provides decent performance if the presumed block size, denoted by $h$, is properly selected. Fig. \ref{fig2} plots the success rates of respective algorithms as a function of the ratio $m/n$ and the sparsity level $K$, respectively. Simulation results show that our proposed algorithm achieves highest success rates among all algorithms and outperforms other methods by a considerable margin. We also noticed that the modified reweighted $\ell_1$ method (MRL1), although not as good as the proposed PD-SBL, still delivers acceptable performance which is comparable to the BSBL and better than the BM-MAP-OMP and the CluSS-MCMC. \begin{figure*}[!t] \centering \subfigure[Success rates vs. $m/n$, $n=100$, $K=25$, and $L=4$]{\includegraphics[width=9cm]{suc-rates-ratio}} \hfil \subfigure[Success rates vs. the sparsity level $K$, $m=40$, $n=100$, and $L=3$]{\includegraphics[width=9cm]{suc-rates-sparsity}} \caption{Success rates of respective algorithms.} \label{fig2} \end{figure*} We now consider the noisy case where the measurements are contaminated by additive noise. The observation noise is assumed multivariate Gaussian with zero mean and covariance matrix $\sigma^2\boldsymbol{I}$. Also, in our simulations, the noise variance is assumed unknown (except for the BM-MAP-OMP). The NMSEs of respective algorithms as a function of the ratio $m/n$ and the sparsity level $K$ are plotted in Fig. \ref{fig3}, where the white Gaussian noise is added such that the signal-to-noise ratio (SNR), which is defined as $\text{SNR(dB)}\triangleq 20\log_{10}(\|\boldsymbol{Ax}\|_2/\|\boldsymbol{w}\|_2)$, is equal to $15$dB for each iteration. We see that our proposed algorithm yields a lower estimation error than other methods in the presence of additive Gaussian noise. \begin{figure*}[!t] \centering \subfigure[Normalized MSEs vs. $m/n$, $n=100$, $K=25$, and $L=4$]{\includegraphics[width=9cm]{NMSE-ratio}} \hfil \subfigure[Normalized MSEs vs. the sparsity level $K$, $m=40$, $n=100$, and $L=3$]{\includegraphics[width=9cm]{NMSE-sparsity}} \caption{Normalized MSEs of respective algorithms.} \label{fig3} \end{figure*} \subsection{Real Data} In this subsection, we carry out experiments on real world images. As it is well-known, images have sparse (or approximately sparse) structures in certain over-complete basis, such as wavelet or discrete cosine transform (DCT) basis. Moreover, the sparse representations usually demonstrate clustered structures whose significant coefficients tend to be located together (see Fig. \ref{fig6}). Therefore images are suitable data sets for evaluating the effectiveness of a variety of block-sparse signal recovery algorithms. We consider two famous pictures `Lena' and `Pirate' in our simulations. In our experiments, the image is processed in a columnwise manner: we sample each column of the $128\times 128$ image using a randomly generated measurement matrix $\boldsymbol{A}\in \mathbb{R}^{m\times 128}$, recover each column from the $m$ measurements, and reconstruct the image based on the $128$ estimated columns. Fig. \ref{fig4} and \ref{fig5} show the original images `Lena' and `Pirate' and the reconstructed images using respective algorithms, where we set $m=64$ and $m=80$ respectively. We see that our proposed algorithm presents the finest image quality among all methods. The result, again, demonstrates its superiority over other existing methods. The reconstruction accuracy of respective algorithms can also be observed from the reconstructed wavelet coefficients. We provide the true wavelet coefficients of one randomly selected column from the image `Lena', and the wavelet coefficients reconstructed by respective algorithms. Results are depicted in Fig. \ref{fig6}. It can be seen that our proposed algorithm provides reconstructed coefficients that are closest to the groundtruth. \begin{figure*}[!t] \centering \begin{tabular}{cccccc} \hspace*{-3ex} \includegraphics[width=5cm]{original-lena}& \hspace*{-10ex} \includegraphics[width=5cm]{SC-SBL-lena} & \hspace*{-10ex} \includegraphics[width=5cm]{SBL-lena} & \hspace*{-10ex} \includegraphics[width=5cm]{BP-lena} \\ \hspace*{-3ex}\includegraphics[width=5cm]{CluSS-lena} & \hspace*{-10ex} \includegraphics[width=5cm]{EBSBL-lena} & \hspace*{-10ex} \includegraphics[width=5cm]{BSBL-lena} & \hspace*{-10ex} \includegraphics[width=5cm]{BM-lena} \end{tabular} \caption{The original image `Lena' and the reconstructed images using respective algorithms.} \label{fig4} \end{figure*} \begin{figure*}[!t] \centering \begin{tabular}{cccccc} \hspace*{-3ex} \includegraphics[width=5cm]{original-pirate}& \hspace*{-10ex} \includegraphics[width=5cm]{SC-SBL-pirate} & \hspace*{-10ex} \includegraphics[width=5cm]{SBL-pirate} & \hspace*{-10ex} \includegraphics[width=5cm]{BP-pirate} \\ \hspace*{-3ex}\includegraphics[width=5cm]{CluSS-pirate} & \hspace*{-10ex} \includegraphics[width=5cm]{EBSBL-pirate} & \hspace*{-10ex} \includegraphics[width=5cm]{BSBL-pirate} & \hspace*{-10ex} \includegraphics[width=5cm]{BM-pirate} \end{tabular} \caption{The original image `Pirate' and the reconstructed images using respective algorithms.} \label{fig5} \end{figure*} \begin{figure*}[!t] \centering \begin{tabular}{cccc} \hspace*{-3ex} \includegraphics[width=4.5cm]{original-DCT}& \hspace*{-5ex} \includegraphics[width=4.5cm]{SC-SBL-DCT} & \hspace*{-5ex} \includegraphics[width=4.5cm]{SBL-DCT} & \hspace*{-5ex} \includegraphics[width=4.5cm]{BP-DCT} \\ \hspace*{-3ex}\includegraphics[width=4.5cm]{CluSS-DCT} & \hspace*{-5ex} \includegraphics[width=4.5cm]{EBSBL-DCT} & \hspace*{-5ex} \includegraphics[width=4.5cm]{BSBL-DCT} & \hspace*{-5ex} \includegraphics[width=4.5cm]{BM-DCT} \end{tabular} \caption{The true wavelet coefficients and the coefficients reconstructed by respective algorithms.} \label{fig6} \end{figure*} \section{Conclusions} \label{sec:conclusion} We developed a new Bayesian method for recovery of block-sparse signals whose block-sparse structures are entirely unknown. A pattern-coupled hierarchical Gaussian prior model was introduced to characterize both the sparseness of the coefficients and the statistical dependencies between neighboring coefficients of the signal. The prior model, similar to the conventional sparse Bayesian learning model, employs a set of hyperparameters to control the sparsity of the signal coefficients. Nevertheless, in our framework, the sparsity of each coefficient not only depends on its corresponding hyperparameter, but also depends on the neighboring hyperparameters. Such a prior has the potential to encourage clustered patterns and suppress isolated coefficients whose patterns are different from their respective neighbors. The hyperparameters, along with the sparse signal, can be estimated by maximizing their posterior probability via the expectation-maximization (EM) algorithm. Numerical results show that our proposed algorithm achieves a significant performance improvement as compared with the conventional sparse Bayesian learning method through exploiting the underlying block-sparse structure, even without knowing the exact locations and sizes of the non-zero blocks. It also demonstrates its superiority over other existing methods and provides state-of-the-art performance for block-sparse signal recovery.
\section{Introduction} Currently, the question of pulsation modulation of magnetic field in stars, both with convective and radiative envelopes, is still opened. \textbf{RR Lyr.} \cite[Babcock (1958)]{bb58} reported a detection of a magnetic field in RR~Lyr. The longitudinal component of the field was found to be variable from $-$1580 to +540 G, but showed no correlation with the pulsation cycle of the star. \cite[Romanov et al.~(1987, 1994)]{romanov87} also registered significant magnetic field in RR Lyr, and found the field to be variable with an amplitude of up to 1.5 kG over the pulsation cycle. On the other hand, \cite[Preston (1967)]{prest67} and \cite[Chadid et al.~(2004)]{chadid04} detected no convincing evidence of a photospheric magnetic field in the star in the years 1963--1964 and 1999--2002, respectively. \textbf{$\eta$ Aql.} Photoelectric magnetometer observations of $\eta$ Aql, performed by \cite[Borra et al.~(1981, 1984)]{borra81,borra84}, detected no magnetic field in this star. \cite[Plachinda (2000)]{[plach00} was the first who detected magnetic field on $\eta$ Aql and reported pulsation modulation of the longitudinal component from $-$100 to +50 G. \cite[Wade et al.~(2002)]{wade02} detected no statistically significant longitudinal magnetic field in $\eta$ Aql during 3 nights in 2001 and concluded that $\eta$ Aql is a non-magnetic star, at least at a level of 10 G. \cite[Grunhut et al.~(2010)]{grunhut04} registered clear Zeeman signatures in Stokes V parameter for $\eta$~Aql and eight other supergiants. \textbf{$\gamma$ Peg.} \cite[Butkovskaya \& Plachinda (2007)]{butk07} reported the modulation of the longitudinal magnetic field in $\beta$ Cephei-type star $\gamma$ Peg (B2 IV)~with the amplitude of about 7 G over the 0.15-day pulsation period of the star. In order to shed some light on the problem of pulsational modulation of stellar magnetic fields, we continued spectropolarimetric monitoring of $\eta$ Aql during 60 nights between 2002 and 2012 using Coud\'e spectrograph at the 2.6-m Shajn telescope of the Crimean Astrophysical Observatory (Ukraine). \section{Results} The technique of Zeeman splitting, used for the measurement of the longitudinal magnetic field, is described in detail by \cite[Butkovskaya \& Plachinda (2007)]{butk07}. We folded all values of the measured longitudinal magnetic field with the pulsational period according to the pulsation ephemeris: JD $= 2450100.861+7.176726E$ \cite[(Kiss \& Vink\'o 2000)]{kiss00}, where $E$ is the number of pulsation cycles. As an example, the pulsation modulation of the longitudinal magnetic field of $\eta$ Aql in 2002 and 2004 is illustrated in Fig. \ref{fig1}. We found that the magnetic field sinusoidally varies in phase of the radial pulsation of $\eta$ Aql. This confirms the previous conclusions of \cite[Plachinda (2000)]{plach00}. However, the amplitude $B$, mean field $B_0$, and phases of maximum and minimum field are changing from year to year (see Table 1 where the F-test indicates the statistical reliability of the detected variability for each year). The possible reason for those variations is stellar axial rotation or dynamo mechanisms. \begin{table} \begin{center} \caption{Parameters of the variability of the longitudinal magnetic field of $\eta$~Aql.} \label{tab1} {\scriptsize \begin{tabular}{|c|c|c|r|r|r|}\hline {\bf Year} & {\bf Phase max} & {\bf Phase min} & {\bf Amplitude $B$, G} & {\bf Mean $B_{0}$, G} & {F-test}\\ \hline 2002 & 0.45 & 0.95& 12.7 $\pm$ 2.2 & 5.5 $\pm$ 1.5 & 0.99 \\ 2004 & 0.78 & 0.28& 13.9 $\pm$ 2.4 & 4.7 $\pm$ 1.5 & 0.99 \\ 2010 & 0.38 & 0.88& 4.3 $\pm$ 1.1 &$-$3.1 $\pm$ 1.1 & 0.98 \\ 2012 & 0.60 & 0.10& 4.2 $\pm$ 2.3 &$-$0.7 $\pm$ 1.5 & 0.69 \\ \hline \end{tabular} } \end{center} \end{table} \begin{figure}[!ht] \begin{center} \includegraphics[width=4.0in]{butkovskaya_fig1c.eps} \caption{Longitudinal magnetic field of $\eta$ Aql folded with the 7.176726-day pulsation period: a) our data obtained in 2002 (\textit{closed and open circles}), data by \cite[Wade et al.~(2002)]{wade02} (\textit{black squares}), and \cite[Grunhut et al.~(2010)]{grunhut04} (\textit{open triangles}); b) our data obtained in 2004 (\textit{closed and open circles}), data by \cite[Borra et al.~(1981, 1984)]{borra81,borra84} (\textit{black squares}). Fitting sinusoids are shown by strong lines. Open circles represent our data that have not been taken into account in the fits.} \label{fig1} \end{center} \end{figure}
\section{Introduction} Valiant's celebrated {\em probably approximately correct} (=PAC) model~\citep{Valiant84} of machine learning led to an extensive research that yielded a whole scientific community devoted to computational learning theory. In the PAC learning model, a learner is given an oracle access to randomly generated samples $(X,Y)\in {\cal X}\times\{0,1\}$ where $X$ is sampled from some {\em unknown} distribution ${\cal D}$ on ${\cal X}$ and $Y=h^{*}(X)$ for some {\em unknown} function $h^{*} : {\cal X} \to \{0,1\}$. Furthermore, it is assumed that $h^*$ comes from a predefined \emph{hypothesis class} ${\cal H}$, consisting of $0,1$ valued functions on ${\cal X}$. The learning problem defined by ${\cal H}$ is to find a function $h:{\cal X}\to\{0,1\}$ that minimizes $\Err_{{\cal D}}(h):=\Pr_{X\sim{\cal D}}(h(X)\not=h^*(X))$. For concreteness' sake we take ${\cal X}=\{\pm 1\}^n$, and we consider the learning problem tractable if there is an algorithm that on input $\epsilon$, runs in time $\poly(n,1/\epsilon)$ and outputs, w.h.p., a hypothesis $h$ with $\Err(h)\le\epsilon$. Assuming $\mathbf{P}\ne\mathbf{NP}$, the status of most basic {\em computational} problems is fairly well understood. In a sharp contrast, almost $30$ years after Valiant's paper, the status of most basic {\em learning} problems is still wide open -- there is a huge gap between the performance of the best known algorithms and hardness results: \begin{itemize} \item No known algorithms can learn depth $2$ circuits, i.e., $\mathrm{DNF}$ formulas. In contrast, we can only rule out learning of circuits of depth $d$, for some unspecified constant $d$ \citep{Kharitonov93}. This result is based on a relatively strong assumption (a certain subexponential lower bound on factoring Blum integers). Under more standard assumptions (RSA in secure), the best we can do is rule out learning of depth $\log n$ circuits \citep{KearnsVa89}. \item It is possible to agnostically learn halfspaces (see section \ref{sec:learning_background} for a definition of agnostic learning) with an approximation ratio of $O\left(\frac{n}{\log n}\right)$. On the other hand, the best known lower bound only rules out exact agnostic learning (\cite{FeldmanGoKhPo06}, based on \cite{KlivansSh06}, under the assumption that the $\tilde{O}\left(n^{1.5}\right)$ unique shortest vector problem is hard). \item No known algorithm can learn intersections of $2$ halfspaces, whereas Klivans and Sherstov~\citep{KlivansSh06} only rule out learning intersections of polynomially many halfspaces (again assuming that $\tilde{O}\left(n^{1.5}\right)$-uSVP is hard). \end{itemize} The crux of the matter, leading to this state of affairs, has to do with the learner's freedom to return {\em any} hypothesis. A learner who may return hypotheses outside the class ${\cal H}$ is called an {\em improper learner}. This additional freedom makes such algorithms potentially more powerful than proper learners. On the other hand, this added flexibility makes it difficult to apply standard reductions from $\mathbf{NP}$-hard problems. Indeed, there was no success so far in proving intractability of a learning problem based on $\mathbf{NP}$-hardness. Moreover, as Applebaum, Barak and Xiao \cite{ApplebaumBaXi08} showed, many standard ways to do so are doomed to fail, unless the polynomial hierarchy collapses. The vast majority of existing lower bounds on learning utilize the crypto-based argument, suggested in \cite{KearnsVa89}. Roughly speaking, to prove that a certain learning problem is hard, one starts with a certain collection of functions, that by assumption are one-way trapdoor permutations. This immediately yields some hard (usually artificial) learning problem. The final step is to reduce this artificial problem to some natural learning problem. Unlike the difficulty in establishing lower bounds for improper learning, the situation in \emph{proper} learning is much better understood. Usually, hardness of proper learning is proved by showing that it is $\mathbf{NP}$-hard to distinguish a realizable sample from an unrealizable sample. I.e., it is hard to tell whether there is some hypothesis in ${\cal H}$ which has zero error on a given sample. This, however, does not suffice for the purpose of proving lower bounds on improper learning, because it might be the case that the learner finds a hypothesis (not from ${\cal H}$) that does not err on the sample even though no $h\in{\cal H}$ can accomplish this. In this paper we present a new methodology for proving hardness of improper learning. Loosely speaking, we show that improper learning is impossible provided that it is hard to distinguish a realizable sample from a \emph{randomly generated} unrealizable sample. Feige \cite{Feige02} conjectured that random 3-$\mathrm{SAT}$ formulas are hard to refute. He derived from this assumption certain hardness of approximation results, which are not known to follow from $\mathbf{P}\ne\mathbf{NP}$. We put forward a (fairly strong) assumption, generalizing Feige's assumption to certain predicates other that 3-$\mathrm{SAT}$. Under this assumption, we show: \begin{enumerate} \item\label{echad} Learning $\mathrm{DNF}$'s is hard. \item\label{shtaim} Agnostically learning halfspaces with a constant approximation ratio is hard, even over the boolean cube. \item\label{shalosh} Learning intersection of $\omega(1)$ halfspaces is hard, even over the boolean cube. \item\label{arba} Learning finite automata is hard. \item\label{chamesh} Learning parity is hard. \end{enumerate} We note that result~\ref{arba} can be established using the cryptographic technique \citep{KearnsVa89}. Result~\ref{chamesh} is often taken as a hardness {\em assumption}. We also conjecture that under our generalization of Feige's assumption it is hard to learn intersections of even constant number of halfspaces. We present a possible approach to the case of four halfspaces. To the best of our knowledge, these results easily imply most existing lower bounds for improper learning. \subsection{Comparison to the cryptographic technique} There is a crucial reversal of order that works in our favour. To lower bound improper learning, we actually need much less than what is needed in cryptography, where a problem and a distribution on instances are appropriate if they fool {\em every algorithm}. In contrast, here we are presented with {\em a concrete} learning algorithms and we devise a problem and a distribution on instances that fail it. Second, cryptographic assumptions are often about the hardness of number theoretic problems. In contrast, the average case assumptions presented here are about $\mathrm{CSP}$ problems. The proximity between $\mathrm{CSP}$ problems and learning problems is crucial for our purposes: Since distributions are very sensitive to gadgets, reductions between average case problems are much more limited than reductions between worst case problems. \subsection{On the role of average case complexity} A key question underlying the present study and several additional recent papers is what can be deduced from the average case hardness of specific problems. Hardness on average is crucial for cryptography, and the security of almost all modern cryptographic systems hinges on the average hardness of certain problems, often from number theory. As shown by Kearns and Valiant \citep{KearnsVa89}, the very same hardness on average assumptions can be used to prove hardness of improper $\mathrm{PAC}$ learning of some hypothesis classes. Beyond these classic results, several recent works, starting from Feige's seminal work \citep{Feige02}, show that average case hardness assumptions lead to dramatic consequences in complexity theory. The main idea of \citep{Feige02} is to consider two possible avenues for progress beyond the classic uses of average hardness: (i) Derive hardness in additional domains, (ii) Investigate the implications of hardness-on-average of other problems. For example, what are the implications of average hardness of $3$-$\mathrm{SAT}$? What about other $\mathrm{CSP}$ problems? Feige \citep{Feige02} and then \citep{Alekhnovich03,BarakKiSt13} show that average case hardness of $\mathrm{CSP}$ problems have surprising implications in hardness of approximation, much beyond the consequences of standard complexity assumptions, or even cryptographic assumptions. Recently, \citep{berthet2013computational} and \cite{daniely2013more} show that hardness on average of planted clique and $3$-$\mathrm{SAT}$ have implications in learning theory, in the specific context of computational-sample tradeoffs. In particular, they show that in certain learning tasks (sparse $\mathrm{PCA}$ and learning halfspaces over sparse vectors) more data can be leveraged to speed up computation. As we show here, average case hardness of $\mathrm{CSP}$ problems has implications even on the hardness of very fundamental tasks in learning theory. Namely, determining the tractability of $\mathrm{PAC}$ learning problems, most of which are presently otherwise inaccessible. \section{Preliminaries} \subsection{Learning Theory}\label{sec:learning_background} A {\em hypothesis class}, ${\cal H}$, is a series of collections of functions ${\cal H}_n\subset \{0,1\}^{{\cal X}_n},\;n=1,2,\ldots$. We often abuse notation and identify ${\cal H}$ with ${\cal H}_n$. The instance space, ${\cal X}_n$, that we consider is either ${\cal X}_n=\{\pm 1\}^n$, ${\cal X}_n=\{0, 1\}^n$ or ${\cal X}_n=\{-1,1,0\}^n$. Concrete hypothesis classes, such as halfspaces, DNF's etc., are denoted $\mathrm{HALFSPACES},\mathrm{DNF}$ etc. Also ${\cal Z}_n:={\cal X}_n\times\{0,1\}$. Distributions on ${\cal Z}_n$ (resp. ${\cal Z}_n^m$) are denoted ${\cal D}_n$ (resp. ${\cal D}_n^m$). {\em Ensembles} of distributions are denoted by ${\cal D}$. That is, ${\cal D}=\{{\cal D}_n^{m(n)}\}_{n=1}^\infty$ where ${\cal D}_n^{m(n)}$ is a distributions on ${\cal Z}_n^{m(n)}$. We say that ${\cal D}$ is a {\em polynomial ensemble} if $m(n)$ is upper bounded by some polynomial in $n$. The error of a hypothesis $h:{\cal X}_n\to\{0,1\}$ w.r.t. ${\cal D}_n$ on ${\cal Z}_n$ is defined as $\Err_{{\cal D}_n}(h)=\Pr_{(x,y)\sim{\cal D}_n }\left(h(x)\ne y\right)$. For a hypothesis class ${\cal H}_n$, we define $\Err_{{\cal D}_n}({\cal H}_n)=\min_{h\in{\cal H}_n}\Err_{{\cal D}_n}(h)$. We say that a distribution ${\cal D}_n$ is \emph{realizable} by $h$ (resp. ${\cal H}_n$) if $\Err_{{\cal D}_n}(h)=0$ (resp. $\Err_{{\cal D}_n}({\cal H}_n)=0$). Similarly, we say that ${\cal D}_n$ is {\em $\epsilon$-almost realizable} by $h$ (resp. ${\cal H}_n$) if $\Err_{{\cal D}_n}(h)\le\epsilon$ (resp. $\Err_{{\cal D}_n}({\cal H}_n)\le\epsilon$). A {\em sample} is a sequence $S=\{(x_1,y_1),\ldots (x_m,y_m)\}\in{\cal Z}^m_n$. The {\em empirical error} of a hypothesis $h:{\cal X}_n\to\{0,1\}$ w.r.t. sample $S$ is $\Err_{S}(h)=\frac{1}{m}\sum_{i=1}^m1(h(x_i)=y_i)$. The {\em empirical error} of a hypothesis class ${\cal H}_n$ w.r.t. $S$ is $\Err_{S}({\cal H}_n)=\min_{h\in{\cal H}_n}\Err_S(h)$. We say that a sample $S$ is \emph{realizable} by $h$ if $\Err_S(h)=0$. The sample $S$ is {\em realizable} by ${\cal H}_n$ if $\Err_S({\cal H}_n)=0$. Similarly, we define the notion of {\em $\epsilon$-almost realizable} sample (by either a hypothesis $h:{\cal X}_n\to\{0,1\}$ or a class ${\cal H}_n$). A {\em learning algorithm}, denoted ${\cal L}$, obtains an error parameter $0<\epsilon<1$, a confidence parameter $0<\delta<1$, a complexity parameter $n$, and an access to an oracle that produces samples according to unknown distribution ${\cal D}_n$ on ${\cal Z}_n$. It should output a (description of) hypothesis $h:{\cal X}_n\to\{0,1\}$. We say that the algorithm ${\cal L}$ {\em (PAC) learns} the hypothesis class ${\cal H}$ if, for every realizable distribution ${\cal D}_n$, with probability $\ge 1-\delta$, ${\cal L}$ outputs a hypothesis with error $\le \epsilon$. We say that an algorithm ${\cal L}$ {\em agnostically learns} ${\cal H}$ if, for every distribution ${\cal D}_n$, with probability $\ge 1-\delta$, ${\cal L}$ outputs a hypothesis with error $\le \Err_{{\cal D}_n}({\cal H})+\epsilon$. We say that an algorithm ${\cal L}$ {\em approximately agnostically learns} ${\cal H}$ with approximation ratio $\alpha=\alpha(n)\ge 1$ if, for every distribution ${\cal D}_n$, with probability $\ge 1-\delta$, ${\cal L}$ outputs a hypothesis with error $\le \alpha\cdot \Err_{{\cal D}_n}({\cal H})+\epsilon$. We say that ${\cal L}$ is {\em efficient} if it runs in time polynomial in $n,1/\epsilon$ and $1/\delta$, and outputs a hypothesis that can be evaluated in time polynomial in $n,1/\epsilon$ and $1/\delta$. We say that ${\cal L}$ is {\em proper} (with respect to ${\cal H}$) if it always outputs a hypothesis in ${\cal H}$. Otherwise, we say that ${\cal L}$ is {\em improper}. Let ${\cal H}=\{{\cal H}_n\subset\{0,1\}^{{\cal X}_n}\mid n=1,2\ldots\}$ and ${\cal H}'=\{{\cal H}'_n\subset\{0,1\}^{{\cal X}'_n}\mid n=1,2\ldots\}$ be two hypothesis classes. We say the ${\cal H}$ is {\em realized} by ${\cal H}'$ if there are functions $g:\mathbb N\to \mathbb N$ and $f_n:{\cal X}_n\to{\cal X}'_{g(n)},\;n=1,2,\ldots$ such that for every $n$, ${\cal H}_n\subset \{h'\circ f_n\mid h'\in{\cal H}'_n\}$. We say that ${\cal H}$ is {\em efficiently realized} by ${\cal H}'$ if, in addition, $f_n$ can be computed it time polynomial in $n$. Note that if ${\cal H}'$ is efficiently learnable (respectively, agnostically learnable, or approximately agnostically learnable) and ${\cal H}$ is efficiently realized by ${\cal H}'$, then ${\cal H}$ is efficiently learnable (respectively, agnostically learnable, or approximately agnostically learnable) as well. \subsection{Constraints Satisfaction Problems} Let $P:\{\pm 1\}^K\to \{0,1\}$ be some boolean predicate (that is, $P$ is any non-constant function from $\{\pm 1\}^K$ to $\{0,1\}$). A {\em $P$-constraint} with $n$ variables is a function $C:\{\pm 1\}^n\to\{0,1\}$ of the form $C(x)=P(j_1x_{i_1},\ldots,j_Kx_{i_K})$ for $j_l\in \{\pm 1\}$ and $K$ distinct $i_l\in [n]$. The {\em CSP problem,} $\mathrm{CSP}(P)$, is the following. An instance to the problem is a collection $J=\{C_1,\ldots,C_m\}$ of $P$-constraints and the objective is to find an assignment $x\in \{\pm 1\}^n$ that maximizes the fraction of satisfied constraints (i.e., constraints with $C_i(x)=1$). The {\em value} of the instance $J$, denoted $\mathrm{VAL}(J)$, is the maximal fraction of constraints that can be simultaneously satisfied. If $\mathrm{VAL}(J)=1$, we say that $J$ is satisfiable. For $1\ge\alpha>\beta>0$, the problem $\mathrm{CSP}^{\alpha,\beta}(P)$ is the decision promise problem of distinguishing between instances to $\mathrm{CSP}(P)$ with value $\ge \alpha$ and instances with value $\le \beta$. Denote $\Lval(P)=\E_{x\sim \mathrm{Uni}(\{\pm 1\}^K)}P(x)$. We note that for every instance $J$ to $\mathrm{CSP}(P)$, $\mathrm{VAL}(J)\ge \Lval(P)$ (since a random assignment $\psi\in\{\pm 1\}^n$ satisfies in expectation $\Lval(P)$ fraction of the constraints). Therefore, the problem $\mathrm{CSP}^{\alpha,\beta}(P)$ is non-trivial only if $\beta\ge \Lval(P)$. We say that $P$ is {\em approximation resistant} if, for every $\epsilon>0$, the problem $\mathrm{CSP}^{1-\epsilon,\Lval(P)+\epsilon}(P)$ is $\mathbf{NP}$-hard. Note that in this case, unless $\mathbf{P}=\mathbf{NP}$, no algorithm for $\mathrm{CSP}(P)$ achieves better approximation ratio than the naive algorithm that simply chooses a random assignment. We will use even stronger notions of approximation resistance: We say that $P$ is {\em approximation resistant on satisfiable instances} if, for every $\epsilon>0$, the problem $\mathrm{CSP}^{1,\Lval(P)+\epsilon}(P)$ is $\mathbf{NP}$-hard. Note that in this case, unless $\mathbf{P}=\mathbf{NP}$, no algorithm for $\mathrm{CSP}(P)$ achieves better approximation ratio than a random assignment, even if the instance is guaranteed to be satisfiable. We say that $P$ is {\em heredity approximation resistant on satisfiable instances} if every predicate that is implied by $P$ (i.e., every predicate $P':\{\pm 1\}^K\to\{0,1\}$ that satisfies $\forall x,\;P(x)\Rightarrow P'(x)$) is approximation resistant on satisfiable instances. Similarly, we define the notion of {\em heredity approximation resistance}. We will consider average case variant of the problem $\mathrm{CSP}^{\alpha,\beta}(P)$. Fix $1\ge \alpha> \Lval(P)$. By a simple counting argument, for sufficiently large constant $C>0$, the value of a random instance with $\ge C\cdot n$ constraints is about $\Lval(P)$, in particular, the probability that a (uniformly) random instance to $\mathrm{CSP}(P)$ with $n$ variables and $\ge Cn$ constraints will have value $\ge \alpha$ is exponentially small. Therefore, the problem of distinguishing between instances with value $\ge\alpha$ and random instances with $m(n)$ constraints can be thought as an average case analogue of $\mathrm{CSP}^{\alpha,\Lval(P)+\epsilon}$. We denote this problem by $\mathrm{CSP}^{\alpha,\mathrm{rand}}_{m(n)}(P)$. Precisely, we say that the problem $\mathrm{CSP}^{\alpha,\mathrm{rand}}_{m(n)}(P)$ is easy, if there exists an efficient randomized algorithm, ${\cal A}$, with the following properties: \begin{itemize} \item If $J$ is an instance to $\mathrm{CSP}(P)$ with $n$ variables, $m(n)$ constraints, and value $\ge \alpha$, then \[ \Pr_{\text{coins of }{\cal A}}\left({\cal A}(J)=``\mathrm{VAL}(J)\ge\alpha"\right)\ge\frac{3}{4} \] \item If $J$ is a random instance to $\mathrm{CSP}(P)$ with $n$ variables and $m(n)$ constraints then, with probability $1-o_n(1)$ over the choice of $J$, \[ \Pr_{\text{coins of }{\cal A}}\left({\cal A}(J)=``J\text{ is random}"\right)\ge \frac{3}{4}~. \] \end{itemize} The problem $\mathrm{CSP}^{\alpha,\mathrm{rand}}_{m(n)}(P)$ will play a central role. In particular, the case $\alpha=1$, that is, the problem of distinguishing between satisfiable instances and random instances. This problem is also known as the problem of refuting random instances to $\mathrm{CSP}(P)$. A simple observation is that the problem $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ becomes easier as $m$ grows: If $m'\ge m$, we can reduce instances of $\mathrm{CSP}^{1,\mathrm{rand}}_{m'(n)}(P)$ to instances of $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ by simply drop the last $m'(n)-m(n)$ clauses. Note that if the original instance was either random or satisfiable, the new instance has the same property as well. Therefore, a natural metric to evaluate a refutation algorithm is the number of random constraints that are required to guarantee that the algorithm will refute the instance with high probability. Another simple observation is that if a predicate $P':\{\pm 1\}^K\to \{0,1\}$ is implied by $P$ then the problem $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P')$ is harder than $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P')$. Indeed, given an instance to $\mathrm{CSP}(P)$, we can create an instance to $\mathrm{CSP}(P')$ by replacing each constraint $C(x)=P(j_1x_{i_1},\ldots,j_Kx_{j_K})$ with the constraint $C'(x)=P'(j_1x_{i_1},\ldots,j_Kx_{j_K})$. We note that this reduction preserves both satisfiability and randomness, and therefore establishes a valid reduction from $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ to $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P')$. \subsection{Resolution refutation and Davis Putnam algorithms}\label{sec:preliminaries_res} A clause is a disjunction of literals, each of which correspond to a distinct variable. Given two clauses of the form $x_i\vee C$ and $\neg x_i\vee D$ for some clauses $C,D$, the {\em resolution rule} infer the clause $C\vee D$. Fix a predicate $P:\{\pm 1\}^K\to \{0,1\}$. A {\em resolution refutation} for an instance $J=\{C_1,\ldots,C_m\}$ to $\mathrm{CSP}(P)$ is a sequence of clauses $\tau=\{T_1,\ldots, T_r\}$ such that $T_r$ is the empty clause, and for every $1\le i\le r$, $T_i$ is either implied by some $C_j$ or resulted from the resolution rule applied on $T_{i_1}$ and $T_{i_2}$ for some $i_1,i_2<i$. We note that every un-satisfiable instance to $\mathrm{CSP}(P)$ has a resolution refutation (of possibly exponential length). We denote by $\mathrm{RES}(J)$ the length of the shortest resolution refutation of $J$. The length of resolution refutation of random $K$-$\mathrm{SAT}$ instances were extensively studied (e.g., \citep{BenWi99}, \citep{BeameKaPiSa98} and \citep{BeamePi96}). Two motivations for these study are the following. First, the famous result of \cite{CookRe79}, shows that $\mathbf{NP}\ne\mathbf{CoNP}$ if and only is there is no propositional proof system that can refute every instance $J$ to $K$-$\mathrm{SAT}$ in length polynomial in $|J|$. Therefore, lower bound on concrete proof systems might bring us closer to $\mathbf{NP}\ne\mathbf{CoNP}$. Also, such lower bounds might indicate that refuting such instances in general, is intractable. A second reason is that many popular algorithms implicitly produces a resolution refutation during their execution. Therefore, any lower bound on the size of the resolution refutation would lead to the same lower bound on the running time or the algorithm. A widely used and studied refutation algorithms of this kind are Davis-Putnam (DPLL) like algorithms \citep{davis1962machine}. A DPLL algorithm is a form of recursive search for a satisfying assignment which on $\mathrm{CSP}$ input $J$ operates as follows: If $J$ contains the constant predicate $0$, it terminates and outputs that the instance is un-satisfiable. Otherwise, a variable $x_i$ is chosen, according to some rule. Each assignment to $x_i$ simplifies the instance $J$, and the algorithm recurses on these simpler instances. \section{The methodology}\label{sec:methodology} We begin by discussing the methodology in the realm of realizable learning, and we later proceed to agnostic learning. Some of the ideas underling our methodology appeared, in a much more limited context, in \cite{daniely2013more}. To motivate the approach, recall how one usually proves that a class cannot be efficiently {\em properly} learnable. Given a hypothesis class ${\cal H}$, let $\Pi({\cal H})$ be the problem of distinguishing between an ${\cal H}$-realizable sample $S$ and one with $\Err_S({\cal H})\ge \frac{1}{4}$. If ${\cal H}$ is efficiently {\em properly} learnable then this problem is in\footnote{The reverse direction is almost true: If the search version of this problem can be solved in polynomial time, then ${\cal H}$ is efficiently learnable.} $\mathbf{RP}$: To solve $\Pi({\cal H})$, we simply invoke a proper learning algorithm ${\cal A}$ that efficiently learns ${\cal H}$, with examples drawn uniformly from $S$. Let $h$ be the output of ${\cal A}$. Since ${\cal A}$ properly learns ${\cal H}$, we have \begin{itemize} \item If $S$ is a realizable sample, then $\Err_S(h)$ is small. \item If $\Err_S({\cal H})\ge\frac{1}{4}$ then, {\em since $h\in{\cal H}$}, $\Err_S(h)\ge \frac{1}{4}$. \end{itemize} This gives an efficient way to decide whether $S$ is realizable. We conclude that if $\Pi({\cal H})$ is $\mathbf{NP}$-hard, then ${\cal H}$ is not efficiently learnable, unless $\mathbf{NP}=\mathbf{RP}$. However, this argument does not rule out the possibility that ${\cal H}$ is still learnable by an {\em improper} algorithm. Suppose now that ${\cal A}$ efficiently and improperly learns ${\cal H}$. If we try to use the above argument to prove that $\Pi({\cal H})$ can be efficiently solved, we get stuck -- suppose that $S$ is a sample and we invoke ${\cal A}$ on it, to get a hypothesis $h$. As before, if $S$ is realizable, $\Err_S(h)$ is small. However, if $S$ is not realizable, since $h$ not necessarily belongs to ${\cal H}$, it still might be the case that $\Err_S(h)$ is small. Therefore, the argument fails. We emphasize that this is not only a mere weakness of the argument -- there are classes for which $\Pi({\cal H})$ is $\mathbf{NP}$-hard, but yet, they are learnable by an improper algorithm\footnote{This is true, for example, for the class of $\mathrm{DNF}$ formulas with 3 $\mathrm{DNF}$ clauses.}. More generally, Applebaum et al \cite{ApplebaumBaXi08} indicate that it is unlikely that hardness of improper learning can be based on standard reductions from $\mathbf{NP}$-hard problems, as the one described here. We see that it is not clear how to establish hardness of improper learning based on the hardness of distinguishing between a realizable and an unrealizable sample. The core problem is that even if $S$ is not realizable, the algorithm might still return a good hypothesis. The crux of our new technique is the observation that if $S$ is {\em randomly generated} unrealizable sample then even improper algorithm cannot return a hypothesis with a small empirical error. The point is that the returned hypothesis is determined solely by the examples that ${\cal A}$ sees and its random bits. Therefore, if ${\cal A}$ is an efficient algorithm, the number of hypotheses it might return cannot be too large. Hence, if $S$ is ``random enough", it likely to be far from all these hypotheses, in which case the hypothesis returned by ${\cal A}$ would have a large error on $S$. We now formalize this idea. Let ${\cal D}=\{{\cal D}^{m(n)}_n\}_{n}$ be a polynomial ensemble of distributions, such that ${\cal D}^{m(n)}_n$ is a distribution on ${\cal Z}_n^{m(n)}$. Think of ${\cal D}^{m(n)}_n$ as a distribution that generates samples that are far from being realizable by ${\cal H}$. We say that it is hard to distinguish between a ${\cal D}$-random sample and a realizable sample if there is no efficient randomized algorithm ${\cal A}$ with the following properties: \begin{itemize} \item For every realizable sample $S\in {\cal Z}^{m(n)}_n$, \[\Pr_{\text{internal coins of }{\cal A}}\left({\cal A}(S)=``realizable"\right)\ge \frac{3}{4}~.\] \item If $S\sim {\cal D}_n^{m(n)}$, then with probability $1-o_n(1)$ over the choice of $S$, it holds that \[\Pr_{\text{internal coins of }{\cal A}}\left({\cal A}(S)=``unrelizable"\right)\ge \frac{3}{4}~.\] \end{itemize} For functions $p,\epsilon:\mathbb N\to(0,\infty)$, we say that ${\cal D}$ is {\em $(p(n),\epsilon(n))$-scattered} if, for large enough $n$, it holds that for every function $f:{\cal X}_n\to\{0,1\}$, \[ \Pr_{S\sim{\cal D}^{m(n)}_n}\left(\Err_{S}(f)\le \epsilon(n)\right)\le 2^{-p(n)}~. \] \begin{example} Let ${\cal D}^{m(n)}_n$ be the distribution over ${\cal Z}_n^{m(n)}$ defined by taking $m(n)$ independent uniformly chosen examples from ${\cal X}_n\times\{0,1\}$. For $f:{\cal X}_n\to\{0,1\}$, $\Pr_{S\sim{\cal D}^{m(n)}_n}\left(\Err_{S}(f)\le \frac{1}{4}\right)$ is the probability of getting at most $\frac{m(n)}{4}$ heads in $m(n)$ independent tosses of a fair coin. By Hoeffding's bound, this probability is $\le 2^{-\frac{1}{8} m(n)}$. Therefore, ${\cal D}=\{{\cal D}^{m(n)}_n\}_{n}$ is $\left(\frac{1}{8} m(n),1/4\right)$-scattered. \end{example} \begin{theorem}\label{thm:basic_realizable} Every hypothesis class that satisfies the following condition is not efficiently learnable. There exists $\beta>0$ such that for every $c>0$ there is an $(n^c,\beta)$-scattered ensemble ${\cal D}$ for which it is hard to distinguish between a ${\cal D}$-random sample and a realizable sample. \end{theorem} \begin{remark} The theorem and the proof below work verbatim if we replace $\beta$ by $\beta(n)$, provided that $\beta(n)> n^{-a}$ for some $a>0$. \end{remark} {\par\noindent {\bf Proof}\space\space} Let ${\cal H}$ be the hypothesis class in question and suppose toward a contradiction that algorithm ${\cal L}$ learns ${\cal H}$ efficiently. Let $M\left(n,1/\epsilon,1/\delta\right)$ be the maximal number of random bits used by ${\cal L}$ when run on the input $n,\epsilon,\delta$. This includes both the bits describing the examples produced by the oracle and ``standard" random bits. Since ${\cal L}$ is efficient, $M\left(n,1/\epsilon,1/\delta\right)< \poly(n,1/\epsilon, 1/\delta)$. Define \[ q(n)=M\left(n,1/\beta,4\right)+n~. \] By assumption, there is a $(q(n),\beta)$-scattered ensemble ${\cal D}$ for which it is hard to distinguish a ${\cal D}$-random sample from a realizable sample. Consider the algorithm ${\cal A}$ defined below. On input $S\in{\cal Z}_n^{m(n)}$, \begin{enumerate} \item Run ${\cal L}$ with parameters $n,\beta$ and $\frac{1}{4}$, such that the examples' oracle generates examples by choosing a random example from $S$. \item Let $h$ be the hypothesis that ${\cal L}$ returns. If $\Err_S(h)\le \beta$, output $\text{``realizable"}$. Otherwise, output $\text{``unrealizable"}$. \end{enumerate} Next, we derive a contradiction by showing that ${\cal A}$ distinguishes a realizable sample from a ${\cal D}$-random sample. Indeed, if the input $S$ is realizable, then ${\cal L}$ is guaranteed to return, with probability $\ge 1-\frac{1}{4}$, a hypothesis $h:{\cal X}_n\to\{0,1\}$ with $\Err_{S}(h)\le \beta$. Therefore, w.p. $\ge \frac{3}{4}$ ${\cal A}$ will output ``realizable". What if the input sample $S$ is drawn from ${\cal D}^{m(n)}_n$? Let ${\cal G}\subset \{0,1\}^{{\cal X}_n}$ be the collection of functions that ${\cal L}$ might return when run with parameters $n,\epsilon(n)$ and $\frac{1}{4}$. We note that $|{\cal G} |\le 2^{q(n)-n}$, since each hypothesis in ${\cal G}$ can be described by $q(n)-n$ bits. Namely, the random bits that ${\cal L}$ uses and the description of the examples sampled by the oracle. Now, since ${\cal D}$ is $(q(n),\beta)$-scattered, the probability that $\Err_{S}(h)\le \beta$ for some $h\in {\cal G}$ is at most $|{\cal G}|2^{-q(n)}\le 2^{-n}$. It follows that the probability that ${\cal A}$ responds ``realizable" is $\le 2^{-n}$. This leads to the desired contradiction and concludes our proof. \hfill $\Box$ Next, we discuss analogue theorem to theorem \ref{thm:basic_realizable} for (approximate) agnostic learning. Let ${\cal D}$ be a polynomial ensemble and $\epsilon:\mathbb N\to (0,1)$. We say that it is hard to distinguish between a ${\cal D}$-random sample and an $\epsilon$-almost realizable sample if there is no efficient randomized algorithm ${\cal A}$ with the following properties: \begin{itemize} \item For every sample $S\in {\cal Z}^{m(n)}_n$ that is $\epsilon(n)$-almost realizable, \[\Pr_{\text{internal coins of }{\cal A}}\left({\cal A}(S)=``almost \;realizable"\right)\ge 3/4~.\] \item If $S\sim {\cal D}_n^{m(n)}$, then with probability $1-o_n(1)$ over the choice of $S$, it holds that \[\Pr_{\text{internal coins of }{\cal A}}\left({\cal A}(S)=``unrelizable"\right)\ge \frac{3}{4}~.\] \end{itemize} \begin{theorem}\label{thm:basic_agnostic} Let $\alpha \ge 1$. Every hypothesis class that satisfies the following condition is not efficiently agnostically learnable with an approximation ratio of $\alpha$. For some $\beta$ and every $c>0$, there is a $(n^c,\alpha \beta+1/n)$-scattered ensemble ${\cal D}$ such that it is hard to distinguish between a ${\cal D}$-random sample and a $\beta$-almost realizable sample. \end{theorem} \begin{remark} As in theorem \ref{thm:basic_realizable}, the theorem and the proof below work verbatim if we replace $\alpha$ by $\alpha(n)$ and $\beta$ by $\beta(n)$, provided that $\beta(n)> n^{-a}$ for some $a>0$. \end{remark} {\par\noindent {\bf Proof}\space\space} Let ${\cal H}$ be the hypothesis class in question and suppose toward a contradiction that ${\cal L}$ efficiently agnostically learns ${\cal H}$ with approximation ratio of $\alpha$. Let $M\left(n,1/\epsilon,1/\delta\right)$ be the maximal number of random bits used by ${\cal L}$ when it runs on the input $n,\epsilon,\delta$. This includes both the bits describing the examples produced by the oracle and the ``standard" random bits. Since ${\cal L}$ is efficient, $M\left(n,1/\epsilon,1/\delta\right)< \poly(n,1/\epsilon, 1/\delta)$. Define, \[ q(n)=M\left(n,n,4\right)+n~. \] By the assumptions of the theorem, there is a $(q(n),\alpha\beta+1/n)$-scattered ensemble ${\cal D}$ such that it is hard to distinguish between a ${\cal D}$-random sample and a $\beta$-almost realizable sample. Consider the following efficient algorithm to distinguish between a ${\cal D}$-random sample and a $\beta$-almost realizable sample. On input $S\in{\cal Z}_n^{m(n)}$, \begin{enumerate} \item Run ${\cal L}$ with parameters $n,1/n$ and $\frac{1}{4}$, such that the examples are sampled uniformly from $S$. \item Let $h$ be the hypothesis returned by the algorithm ${\cal L}$. If $\Err_S(h)\le \alpha\beta+1/n$, return $\text{``almost realizable"}$. Otherwise, return $\text{``unrealizable"}$. \end{enumerate} Next, we derive a contradiction by showing that this algorithm, which we denote by ${\cal A}$, distinguishes between a realizable sample and a ${\cal D}$-random sample. Indeed, if the input $S$ is $\beta$-almost realizable, then ${\cal L}$ is guaranteed to return, with probability $\ge 1-\frac{1}{4}$, a hypothesis $h:{\cal X}_n\to\{0,1\}$ with $\Err_{S}(h)\le \alpha\beta+1/n$. Therefore, the algorithm ${\cal A}$ will return, w.p. $\ge \frac{3}{4}$, ``almost realizable". Suppose now that the input sample $S$ is drawn according to ${\cal D}_n$. Let ${\cal G}\subset \{0,1\}^{{\cal X}_n}$ be the collection of functions that the learning algorithm ${\cal L}$ might return when it runs with the parameters $n,1/n$ and $\frac{1}{4}$. Note that each hypothesis in ${\cal G}$ can be described by $q(n)-n$ bits, namely, the random bits used by ${\cal L}$ and the description of the examples sampled by the oracle. Therefore, $|{\cal G} |\le 2^{q(n)-n}$. Now, since ${\cal D}$ is $(q(n),\alpha\beta+1/n)$-scattered, the probability that some function in $h\in {\cal G}$ will have $\Err_{S}(h)\le \alpha\beta+1/n$ is at most $|{\cal G}|2^{-q(n)}\le 2^{-n}$. It follows that the probability that the algorithm ${\cal A}$ will return ``almost realizable" is $\le 2^{-n}$. \hfill $\Box$ \section{The strong random CSP assumption} In this section we put forward and discuss a new assumption that we call ``the strong random CSP assumption" or $\mathrm{SRCSP}$ for short. It generalizes Feige's assumption \citep{Feige02}, as well as the assumption of Barak, Kindler and Steurer \citep{BarakKiSt13}. This new assumption, together with the methodology described in section \ref{sec:methodology}, are used to establish lower bounds for improper learning. Admittedly, our assumption is strong, and an obvious quest, discussed in the end of this section is to find ways to derive similar conclusions from weaker assumptions. The $\mathrm{SRCSP}$ assumption claims that for certain predicates $P:\{\pm 1\}^K\to \{0,1\}, d>0$ and $\alpha>0$, the decision problem $\mathrm{CSP}^{\alpha,\mathrm{rand}}_{n^d}(P)$ is intractable. We first consider the case $\alpha=1$. To reach a plausible assumption, let us first discuss Feige's assumption, and the existing evidence for it. Denote by $\mathrm{SAT}_3:\{\pm 1\}^3\to \{0,1\}$ the $3$-$\mathrm{SAT}$ predicate $\mathrm{SAT}_3(x_1,x_2,x_3)=x_1\vee x_2\vee x_3$. \begin{assumption}[Feige]\label{hyp:feige} For every sufficiently large constant $C>0$, $\mathrm{CSP}^{1,\mathrm{rand}}_{C\cdot n}(\mathrm{SAT}_3)$ is intractable. \end{assumption} Let us briefly summarize the evidence for this assumption. \begin{itemize} \item {\bf Hardness of approximation.} Feige's conjecture can be viewed as a strengthening of Hastad's celebrated result \citep{haastad2001some} that $\mathrm{SAT}_3$ is approximation resistant on satisfiable instances. Hastad's result implies that under $\mathbf{P}\ne \mathbf{NP}$, it is hard to distinguish satisfiable instances to $\mathrm{CSP}(\mathrm{SAT}_3)$ from instances with value $\le \frac{7}{8}+\epsilon$. The collection of instances with value $\le \frac{7}{8}+\epsilon$ includes most random instances with $C\cdot n$ clauses for sufficiently large $C$. Feige's conjecture says that the problem remains intractable even when restricted to these random instances. We note that approximation resistance on satisfiable instances is a necessary condition for the validity of Feige's assumption. Indeed, for large enough $C>0$, with probability $1-o_n(1)$, the value of a random instance to $\mathrm{CSP}(\mathrm{SAT}_3)$ is $\le \frac{7}{8}+\epsilon$. Therefore, tractability of $\mathrm{CSP}^{1,\frac{7}{8}+\epsilon}(\mathrm{SAT}_3)$ would lead to tractability of $\mathrm{CSP}^{1,\mathrm{rand}}_{C\cdot n}(\mathrm{SAT}_3)$. \item {\bf Performance of known algorithms.} The problem of refuting random $3$-$\mathrm{SAT}$ formulas has been extensively studied and a many algorithms were studied. The best known algorithms \citep{feige2004easily} can refute random instances with $\Omega\left(n^{1.5}\right)$ random constraints. Moreover resolution lower bounds \citep{BenWi99} show that many algorithms run for exponential time when applied to random instances with $O\left(n^{1.5-\epsilon}\right)$ constraints. \end{itemize} We aim to generalize Feige's assumption in two aspects -- (i) To predicates other than $\mathrm{SAT}_3$, and (ii) To problems with super-linearly many constraints. Consider the problem $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ for some predicate $P:\{\pm 1\}^K\to\{0,1\}$. As above, the intractability of $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ strengthens the claim that $P$ is approximation resistant on satisfiable instances. Also, for $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ to be hard, it is necessary that $P$ is approximation resistant on satisfiable instances. In fact, as explained in section \ref{sec:preliminaries_res}, if $P':\{\pm 1\}^K\to\{0,1\}$ is implied by $P$, then the problem $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P)$ can be easily reduced to $\mathrm{CSP}^{1,\mathrm{rand}}_{m(n)}(P')$. Therefore, to preserve the argument of the first evidence of Feige's conjecture, it is natural to require that $P$ is {\em heredity} approximation resistant on satisfiable instances. Next, we discuss what existing algorithms can do. The best known algorithms for the predicate $\mathrm{SAT}_K(x_1,\ldots,x_K)=\vee_{i=1}^Kx_i$ can only refute random instances with $\Omega\left(n^{\lfloor\frac{K}{2}\rfloor}\right)$ constraints \citep{coja2010efficient}. This gives some evidence that it becomes harder to refute random instances of $\mathrm{CSP}(P)$ as the number of variables grows. Namely, that many random constraints are needed to efficiently refute random instances. Of course, some care is needed with counting the ``actual" number of variables. Clearly, only {\sl certain} predicates have been studied so far. Therefore, to reach a plausible assumption, we consider the {\em resolution refutation complexity} of random instances to $\mathrm{CSP}(P)$. And consequently, also the performance of a large class of algorithms, including Davis-Putnam style (DPLL) algorithms. Davis-Putnam algorithms have been subject to an extensive study, both theoretical and empirical. Due to the central place that they occupy, much work has been done since the late 80's, to prove lower bounds on their performance in refuting random $K$-$\mathrm{SAT}$ formulas. These works relied on the fact that these algorithms implicitly produce a resolution refutation during their execution. Therefore, to derive a lower bound on the run time of these algorithms, exponential lower bounds were established on the resolution complexity of random instances to $\mathrm{CSP}(\mathrm{SAT}_K)$. These lower bounds provide support to the belief that it is hard to refute not-too-dense random $K$-$\mathrm{SAT}$ instances. We define the {\em $0$-variability,} $\mathrm{VAR}_0(P)$, of a predicate $P$ as the smallest cardinality of a set of $P$'s variables such that there is an assignment to these variables for which $P(x)=0$, regardless of the values assigned to the other variables. By a simple probabilistic argument, a random $\mathrm{CSP}(P)$ instance with $\Omega\left(n^r\right)$ constraints, where $r=\mathrm{VAR}_0(P)$ is almost surely unsatisfiable with a resolution proof of constant size. Namely, w.p. $1-o_n(1)$, there are $2^r$ constraints that are inconsistent, since some set of $r$ variables appears in all $2^r$ possible ways in the different clauses. On the other hand, we show in section \ref{sec:res} that a random $\mathrm{CSP}(P)$ problem with $O\left(n^{c\cdot r}\right)$ constraints has w.h.p. exponential resolution complexity. Here $c>0$ is an absolute constant. Namely, \begin{theorem}\label{thm:res_main} There is a constant $C>0$ such that for every $d>0$ and every predicate $P$ with $\mathrm{VAR}_0(P)\ge C\cdot d$, the following holds. With probability $1-o_n(1)$, a random instance of $\mathrm{CSP}(P)$ with $n$ variables and $n^d$ constraints has resolution refutation length $\ge 2^{\Omega\left(\sqrt{n}\right)}$. \end{theorem} To summarize, we conclude that the parameter $\mathrm{VAR}_0(P)$ controls the resolution complexity of random instances to $\mathrm{CSP}(P)$. In light of the above discussion, we put forward the following assumption. \begin{assumption}[SRCSP -- part 1]\label{hyp:sat} There is a function $f:\mathbb N\to\mathbb N$ such that the following holds. Let $P$ be a predicate that is heredity approximation resistant on satisfiable instances with $\mathrm{VAR}_0(P)\ge f(d)$. Then, it is hard to distinguish between satisfiable instances of $\mathrm{CSP}(P)$ and random instances with $n^d$ constraints. \end{assumption} Next, we motivate a variant of the above assumption, that accommodates also predicates that are not heredity approximation resistant. A celebrated result of Raghavendra \citep{raghavendra2008optimal} shows that under the unique games conjecture \citep{khot2002power}, a certain $\mathrm{SDP}$-relaxation-based algorithm is (worst case) optimal for $\mathrm{CSP}(P)$, for every predicate $P$. Barak et al. \cite{BarakKiSt13} conjectured that this algorithm is optimal even on random instances. They considered the performance of this algorithm on random instances and purposed the following assumption, which they called the ``random $\mathrm{CSP}$ hypothesis". Define $\Uval(P)=\max_{{\cal D}}\E_{x\sim{\cal D}}P(x)$, where the maximum is taken over all pairwise uniform distributions\footnote{A distribution is {\em pairwise uniform} if, for every pair of coordinates, the distribution induced on these coordinates is uniform.} on $\{\pm 1\}^K$. \begin{assumption}[RSCP]\label{hyp:barak_kindler_steurer} For every $\epsilon>0$ and sufficiently large $C>0$, it is hard to distinguish instances with value $\ge \Uval(P)-\epsilon$ from random instances with $C\cdot n$ constraints. \end{assumption} Here we generalize the $\mathrm{RCSP}$ assumption to random instances with much more than $C\cdot n$ constraints. As in assumption \ref{hyp:sat}, the $0$-variability of $P$ serves to quantify the number of random constraints needed to efficiently show that a random instance has value $< \Uval(P)-\epsilon$. \begin{assumption}[SRSCP - part 2]\label{hyp:not_sat} There is a function $f:\mathbb N\to\mathbb N$ such that for every predicate $P$ with $\mathrm{VAR}_0(P)\ge f(d)$ and for every $\epsilon>0$, it is hard to distinguish between instances with value $\ge \Uval(P)-\epsilon$ and random instances with $n^d$ constraints. \end{assumption} Finally, we define the notion of a $\mathrm{SRCSP}$-hard problem. \begin{terminology} A computational problem is $\mathrm{SRCSP}$-hard if its tractability contradicts assumption \ref{hyp:sat} or \ref{hyp:not_sat}. \end{terminology} \subsection{Toward weaker assumptions}\label{sec:weaker} The $\mathrm{SRCSP}$ assumption is strong. It is highly desirable to arrive at similar conclusions from substantially weaker assumptions. A natural possibility that suggests itself is the $\mathrm{SRCSP}$ assumption, restricted to $\mathrm{SAT}$: \begin{assumption}\label{hyp:only_sat} There is a function $f:\mathbb N\to\mathbb N$ such that for every $K\ge f(d)$, it is hard to distinguish satisfiable instances of $\mathrm{CSP}(\mathrm{SAT}_K)$ from random instances with $n^d$ constraints. \end{assumption} We are quite optimistic regarding the success of this direction: The lower bounds we prove here use the $\mathrm{SRCSP}$-assumption only for certain predicates, and do not need the full power of the assumption. Moreover, for the hypothesis classes of $\mathrm{DNF}$'s, intersection of halfspaces, and finite automata, these predicates are somewhat arbitrary. In \citep{Feige02}, it is shown that for predicates of arity $3$, assumption \ref{hyp:not_sat} is implied by the same assumption restricted to the $\mathrm{SAT}$ predicate. This gives a hope to prove, based on assumption \ref{hyp:only_sat}, that the $\mathrm{SRCSP}$-assumption is true for predicates that are adequate to our needs. \section{Summary of results} \subsection{Learning $\mathrm{DNF}$'s} A {\em $\mathrm{DNF}$ clause} is a conjunction of literals. A {\em $\mathrm{DNF}$ formula} is a disjunction of $\mathrm{DNF}$ clauses. Each $\mathrm{DNF}$ formula over $n$ variables naturally induces a function on $\{\pm 1\}^n$. We define the size of a $\mathrm{DNF}$ clause as the number of its literals and the size of a $\mathrm{DNF}$ formula as the sum of the sizes of its clauses. As $\mathrm{DNF}$ formulas are very natural form of predictors, learning hypothesis classes consisting of $\mathrm{DNF}$'s formulas of polynomial size has been a major effort in computational learning theory. Already in Valiant's paper \citep{Valiant84}, it is shown that for every constant $q$, the hypothesis class of all $\mathrm{DNF}$-formulas with $\le q$ clauses is efficiently learnable. The running time of the algorithm is, however, exponential in $q$. We also note that Valiant's algorithm is improper. For general polynomial-size $\mathrm{DNF}$'s, the best known result \citep{klivans2001learning} shows learnability in time $\frac{1}{\epsilon}\cdot 2^{\tilde{O}\left(n^{\frac{1}{3}}\right)}$. Better running times (quasi-polynomial) are known under distributional assumptions \citep{LinialMaNi89,mansour1995nlog}. As for lower bounds, {\em properly} learning $\mathrm{DNF}$'s is known to be hard \citep{PittVa88}. However, proving hardness of improper learning of polynomial $\mathrm{DNF}$'s has remained a major open question in computational learning theory. Noting that $\mathrm{DNF}$ clauses coincide with depth $2$ circuits, a natural generalization of $\mathrm{DNF}$'s is circuits of small depth. For such classes, certain lower bounds can be obtained using the cryptographic technique. Kharitonov \cite{Kharitonov93} has shown that a certain subexponential lower bound on factoring Blum integers implies hardness of learning circuits of depth $d$, for some unspecified constant $d$. Under more standard assumptions (that the $\mathrm{RSA}$ cryptosystem is secure), best lower bounds \citep{KearnsVa89} only rule out learning of circuits of depth $\log(n)$. For a function $q:\mathbb N\to\mathbb N$, denote by $\mathrm{DNF}_{q(n)}$ the hypothesis class of functions over $\{\pm 1\}^n$ that can be realized by $\mathrm{DNF}$ formulas of size at most $q(n)$. Also, let $\mathrm{DNF}^{q(n)}$ be the hypothesis class of functions over $\{\pm 1\}^n$ that that can be realized by $\mathrm{DNF}$ formulas with at most $q(n)$ clauses. Since each clause is of size at most $n$, $\mathrm{DNF}^{q(n)}\subset \mathrm{DNF}_{nq(n)}$. As mentioned, for a constant $q$, the class $\mathrm{DNF}^{q}$ is efficiently learnable. We show that for every super constant $q(n)$, it is $\mathrm{SRCSP}$-hard to learn $\mathrm{DNF}^{q(n)}$: \begin{theorem}\label{thm:dnf_few_clauses} If $\lim_{n\to\infty}q(n)=\infty$ then learning $\mathrm{DNF}^{q(n)}$ is $\mathrm{SRCSP}$-hard. \end{theorem} Since $\mathrm{DNF}^{q(n)}\subset \mathrm{DNF}_{nq(n)}$, we immediately conclude that learning $\mathrm{DNF}$'s of size, say, $\le n\log(n)$, is $\mathrm{SRCSP}$-hard. By a simple scaling argument, we obtain an even stronger result: \begin{corollary}\label{cor:dnf_small} For every $\epsilon>0$, it is $\mathrm{SRCSP}$-hard to learn $\mathrm{DNF}_{n^\epsilon}$. \end{corollary} \begin{remark}\label{rem:boosting} Following the Boosting argument of Schapire \cite{Schapire89}, hardness of improper learning of a class ${\cal H}$ immediately implies that for every $\epsilon > 0$, there is no efficient algorithm that when running on a distribution that is realized by ${\cal H}$, guaranteed to output a hypothesis with error $\le \frac{1}{2}-\epsilon$. Therefore, hardness results of improper learning are very strong, in the sense that they imply that the algorithm that just makes a random guess for each example, is essentially optimal. \end{remark} \subsection{Agnostically learning halfspaces} Let $\mathrm{HALFSPACES}$ be the hypothesis class of halfspaces over $\{-1,1\}^n$. Namely, for every $w\in\mathbb R^n$ we define $h_w:\{\pm 1\}^n\to\{0,1\}$ by $h_w(x)=\sign\left(\inner{w,x}\right)$, and let \[ \mathrm{HALFSPACES}=\left\{h_w\mid w\in\mathbb R^n\right\}~. \] We note that usually halfspaces are defined over $\mathbb R^n$, but since we are interested in lower bounds, looking on this more restricted class just make the lower bounds stronger. The problem of learning halfspaces is as old as the field of machine learning, starting with the perceptron algorithm \citep{Rosenblatt58}, through the modern $\mathrm{SVM}$ \citep{Vapnik98}. As opposed to learning $\mathrm{DNF}$'s, learning halfspaces in the realizable case is tractable. However, in the agnostic PAC model, the best currently known algorithm for learning halfspaces runs in time exponential in $n$ and the best known approximation ratio of polynomial time algorithms is $O\left(\frac{n}{\log(n)}\right)$. Better running times (usually of the form $n^{\poly\left(\frac{1}{\epsilon}\right)}$) are known under distributional assumptions (e.g. \cite{kalai2008agnostically}). The problem of \emph{proper} agnostic learning of halfspaces was shown to be hard to approximate within a factor of $2^{\log^{1-\epsilon}(n)}$ \citep{arora1993hardness}. Using the cryptographic technique, improper learning of halfspaces is known to be hard, under a certain cryptographic assumption regarding the shortest vector problem (\cite{FeldmanGoKhPo06}, based on \cite{KlivansSh06}). No hardness results are known for \emph{approximately and improperly} learning halfspaces. Here, we show that: \begin{theorem}\label{thm:halfspaces} For every constant $\alpha\ge 1$, it is $\mathrm{SRCSP}$-hard to approximately agnostically learn $\mathrm{HALFSPACES}$ with an approximation ratio of $\alpha$. \end{theorem} \subsection{Learning intersection of halfspaces} For a function $q:\mathbb N\to\mathbb N$, we let $\mathrm{INTER}_{q(n)}$ be the hypothesis class of intersection of $\le q(n)$ halfspaces. That is, $\mathrm{INTER}_{q(n)}$ consists of all functions $f:\{\pm 1\}^n\to\{0,1\}$ for which there exist $w_1,\ldots w_k\in\mathbb R^{n}$ such that $f(x)=1$ if and only if $\forall i,\inner{w_i,x}>0$. Learning intersection of halfspaces has been a major challenge in machine learning. Beside being a natural generalization of learning halfspaces, its importance stems from {\em neural networks} \citep{bishop1995neural}. Learning neural networks was popular in the 80's, and enjoy a certain comeback nowadays. A neural network is composed of layers, each of which is composed of nodes. The first layer consists of $n$ nodes, containing the input values. The nodes in the rest of the layers calculates a value according to a halfspace (or a ``soft'' halfspace obtained by replacing the sign function with a sigmoidal function) applied on the values of the nodes in the previous layer. The final layer consists of a single node, which is the output of the whole network. Neural networks naturally induce several hypothesis classes (according to the structure of the network). The class of intersection of halfspaces is related to those classes, as it can be realized by very simple neural networks: the class $\mathrm{INTER}_{q(n)}$ can be realized by neural networks with only an input layer, a single hidden layer, and output layer, so that there are $q(n)$ nodes in the second layer. Therefore, lower bounds on improperly learning intersection of halfspaces implies lower bounds on improper learning of neural networks. Exact algorithms for learning $\mathrm{INTER}_{q(n)}$ run in time exponential in $n$. Better running times (usually of the form $n^{\poly\left(\frac{1}{\epsilon}\right)}$) are known under distributional assumptions (e.g. \cite{klivans2002learning}). It is known that properly learning intersection of even 2 halfspaces is hard \citep{khot2011hardness}. For improper learning, Klivans and Sherstov \cite{KlivansSh06} have shown that learning an intersection of polynomially many half spaces is hard, under a certain cryptographic assumption regarding the shortest vector problem. Noting that every $\mathrm{DNF}$ formula with $q(n)$ clauses is in fact the complement of an intersection of $q(n)$ halfspaces\footnote{In the definition of $\mathrm{INTER}$, we considered halfspaces with no threshold, while halfspaces corresponding to $\mathrm{DNF}$s do have a threshold. This can be standardly handled by padding the examples with a single coordinate of value $1$.}, we conclude from theorem \ref{thm:dnf_few_clauses} that intersection of every super constant number of halfsapces is hard. \begin{theorem}\label{thm:intersection} If $\lim_{n\to\infty}q(n)=\infty$ then learning $\mathrm{INTER}_{q(n)}$ is $\mathrm{SRCSP}$-hard. \end{theorem} In section \ref{sec:4half} we also describe a route that might lead to the result that learning $\mathrm{INTER}_{4}$ is $\mathrm{SRCSP}$-hard. \subsection{Additional results} In addition to the results mentioned above, we show that learning the class of finite automata of polynomial size is $\mathrm{SRCSP}$-hard. Hardness of this class can also be derived using the cryptographic technique, based on the assumption that the $\mathrm{RSA}$ cryptosystem is secure \citep{KearnsVa89}. Finally, we show that agnostically learning parity with any constant approximation ratio is $\mathrm{SRCSP}$-hard. Parity is not a very interesting class from the point of view of practical machine learning. However, learning this class is related to several other problems in complexity \citep{blum2003noise}. We note that hardness of agnostically learning parity, even in a more relaxed model than the agnostic PAC model (called the random classification noise model), is a well accepted hardness assumption. In section \ref{sec:res} we prove lower bounds on the size of a resolution refutation for random $\mathrm{CSP}$ instances. In section \ref{sec:base_on_np} we show that unless the polynomial hierarchy collapses, there is no ``standard reduction" from an $\mathbf{NP}$-hard problem (or a $\mathbf{CoNP}$-hard problem) to random $\mathrm{CSP}$ problems. \subsection{On the proofs} Below we outline the proof for $\mathrm{DNF}$s. The proof for halfspaces and parities is similar. For every $c>0$, we start with a predicate $P:\{\pm 1\}^K\to\{0,1\}$, for which the problem $\mathrm{CSP}^{1,\mathrm{rand}}_{n^c}(P)$ is hard according to the $\mathrm{SRCSP}$-assumption, and reduce it to the problem of distinguishing between a $(\Omega(n^c),\frac{1}{5})$-scattered sample and a realizable sample. Since $c$ is arbitrary, the theorem follows from theorem \ref{thm:basic_realizable}. The reduction is performed as follows. Consider the problem $\mathrm{CSP}(P)$. Each assignment naturally defines a function from the collection of $P$-constraints to $\{0,1\}$. Hence, if we think about the constraints as instances and about the assignments as hypotheses, the problem $\mathrm{CSP}(P)$ turns into some kind of a learning problem. However, in this interpretation, all the instances we see have positive labels (since we seek an assignment that satisfies as many instances as possible). Therefore, the problem $\mathrm{CSP}^{1,\mathrm{rand}}_{n^c}(P)$ results in ``samples" which are not scattered at all. To overcome this, we show that the analogous problem to $\mathrm{CSP}^{1,\mathrm{rand}}_{n^c}(P)$, where $(\neg P)$-constraints are also allowed, is hard as well (using the assumption on the hardness of $\mathrm{CSP}^{1,\mathrm{rand}}_{n^c}(P)$). The hardness of the modified problem can be shown by relying on the special predicate we work with. This predicate was defined in the recent work of Huang \citep{huang2013approximation}, and it has the property of being heredity approximation resistant, even though $|P^{-1}(1)|\le 2^{O\left(K^{1/3}\right)}$. At this point, we have an (artificial) hypothesis class which is $\mathrm{SRCSP}$-hard to learn by theorem \ref{thm:basic_realizable}. In the next and final step, we show that this class can be efficiently realized by $\mathrm{DNF}$s with $\omega(1)$ clauses. The reduction uses the fact that every boolean function can be expressed by a $\mathrm{DNF}$ formula (of possibly exponential size). Therefore, $P$ can be expressed by a $\mathrm{DNF}$ formula with $2^K$ clauses. Based on this, we show that each hypothesis in our artificial class can be realized by a $\mathrm{DNF}$ formula with $2^K$ clauses, which establishes the proof. The results about learning automata and learning intersection of $\omega(1)$ halfspaces follow from the result about $\mathrm{DNF}$s: We show that these classes can efficiently realize the class of $\mathrm{DNF}$s with $\omega(1)$ clauses. In section \ref{sec:4half} we suggest a route that might lead to the result that learning intersection of $4$ halfspaces is $\mathrm{SRCSP}$-hard: We show that assuming the unique games conjecture, a certain family of predicates are heredity approximation resistant. We show also that for these predicates, the problem $\mathrm{CSP}^{1,\alpha}(P)$ is $\mathbf{NP}$-hard for some $1>\alpha>0$. This leads to the conjecture that these predicates are in fact heredity approximation resistant. Conditioning on the correctness of this conjecture, we show that it is $\mathrm{SRCSP}$-hard to learn intersection of $4$-halfspaces. This is done using the strategy described for $\mathrm{DNF}$s. The proof of the resolution lower bounds (section \ref{sec:res}) relies on the strategy and the ideas introduced in \citep{haken1985intractability} and farther developed in \citep{BeamePi96,BeameKaPiSa98,BenWi99}. The proof that it is unlikely that the correctness of the $\mathrm{SRCSP}$-assumption can be based on $\mathbf{NP}$-hardness (section \ref{sec:base_on_np}) uses the idea introduced in \citep{ApplebaumBaXi08}: we show that if an $\mathbf{NP}$-hard problem (standardly) reduces to $\mathrm{CSP}^{\alpha,\mathrm{rand}}_{m(n)}(P)$, then the problem has a statistical zero knowledge proof. It follows that $\mathbf{NP}\subset \mathbf{SZKP}$, which collapses the polynomial hierarchy. \section{Future work}\label{sec:future} We elaborate below on some of the numerous open problems and research directions that the present paper suggests. \subsection{Weaker assumptions?} First and foremost, it is very desirable to draw similar conclusions from assumption substantially weaker than $\mathrm{SRCSP}$ (see section~\ref{sec:weaker}). Even more ambitiously, is it possible to reduce some $\mathbf{NP}$-hard problem to some of the problems that are deemed hard by the $\mathrm{SRCSP}$ assumption? In section \ref{sec:base_on_np}, we show that a pedestrian application of this approach is doomed to fail (unless the polynomial hierarchy collapses). This provides, perhaps, a moral justification for an ``assumption based" study of average case complexity. \subsection{The $\mathrm{SRCSP}$-assumption} We believe that the results presented here, together with \citep{Feige02,Alekhnovich03,daniely2013more,berthet2013computational} and \citep{BarakKiSt13}, make a compelling case that it is of fundamental importance for complexity theory to understand the hardness of random $\mathrm{CSP}$ problems. In this context, the $\mathrm{SRCSP}$ assumption is an interesting conjecture. There are, of course, many ways to try to refute it. On the other hand, current techniques in complexity theory seem too weak to prove it, or even to derive it from standard complexity assumptions. Yet, there are ways to provide more circumstantial evidence in favor of this assumption: \begin{itemize} \item As discussed in the previous section, one can try to derive it, even partially, from weaker assumptions. \item Analyse the performance of existing algorithms. In section \ref{sec:res} it is shown that no Davis-Putnam algorithm can refute the $\mathrm{SRCSP}$ assumption. Also, Barak et al \citep{BarakKiSt13} show that the basic $\mathbf{SDP}$ algorithm \citep{raghavendra2008optimal} cannot refute assumption \ref{hyp:not_sat}, and also \ref{hyp:sat} for certain predicates (those that contain a pairwise uniform distribution). Such results regarding additional classes of algorithms will lend more support to the assumption's correctness. \item Show lower bounds on the proof complexity of random $\mathrm{CSP}$ instances in refutation systems stronger than resolution. \end{itemize} For a further discussion, see \cite{BarakKiSt13}. Interest in the $\mathrm{SRCSP}$ assumption calls for a better understanding of heredity approximation resistance. For recent work in this direction, see \citep{huang2012approximation,huang2013approximation}. \subsection{More applications} We believe that the method presented here and the $\mathrm{SRCSP}$-assumption can yield additional results in learning and approximation. Here are several basic questions in learning theory that we are unable to resolve even under the $\mathrm{SRCSP}$-assumption. \begin{enumerate} \item Decision trees are very natural hypothesis class, that is not known to be efficiently learnable. Is it $\mathrm{SRCSP}$-hard to learn decision trees? \item What is the real approximation ratio of learning halfspaces? We showed that it is $\mathrm{SRCSP}$-hard to agnostically learn halfspaces with a constant approximation ratio. The best known algorithm only guarantees an approximation ratio of $\frac{n}{\log n}$. This is still a huge gap. See remark \ref{rem:halfsapces} for some speculations about this question. \item Likewise for learning large margin halfspaces (see remark \ref{rem:large_margin}) and for parity. \item Prove that it is $\mathrm{SRCSP}$-hard to learn intersections of a constantly many halfspaces. This might be true even for $2$ halfspaces. In section \ref{sec:4half}, we suggest a route to prove that intersection of $4$ halfspaces is $\mathrm{SRCSP}$-hard. \end{enumerate} Besides application to learning and approximation, it would be fascinating to see applications of the $\mathrm{SRCSP}$-assumption in other fields of complexity. It will be a poetic justice if we could apply it to cryptography. We refer the reader to \citep{BarakKiSt13} for a discussion. Finding implications in fields beyond cryptography, learning and approximation would be even more exciting. \section{Proofs of the lower bounds}\label{sec:proofs} Relying on our general methodology given in section \ref{sec:methodology}, to show that a learning problem is $\mathrm{SRCSP}$-hard, we need to find a scattered ensamble, ${\cal D}$, such that it is $\mathrm{SRCSP}$-hard to distinguish between a realizable sample and a ${\cal D}$-random sample. We will use the following simple criterion for an ensamble to be scattered. \begin{proposition}\label{prop:scat_crit} Let ${\cal D}$ be some distribution on a set ${\cal X}$. For even $m$, let $X_1,\ldots,X_m$ be independent random variables drawn according to ${\cal D}$. Consider the sample $S=\{(X_1,1),(X_2,0)\ldots,(X_{m-1},1),(X_{m},0)\}$. Then, for every $h:{\cal X}\to\{0,1\}$, \[ \Pr_{S}\left(\Err_S(h)\le \frac{1}{5}\right)\le 2^{-\frac{9}{100}m} \] \end{proposition} {\par\noindent {\bf Proof}\space\space} For $1\le i\le \frac{m}{2}$ let $T_i=1[h(X_{2i-1})\ne 1]+1[h(X_{2i})\ne 0]$. Note that $\Err_{S}(h)=\frac{1}{m}\sum_{i=1}^{\frac{m}{2}} T_i$. Also, the $T_i$'s are independent random variables with mean $1$ and values between $0$ and $2$. Therefore, by Hoeffding's bound, \[ \Pr_{S}\left(\Err_S(h)\le \frac{1}{5}\right)\le e^{-\frac{9}{100}m} \le 2^{-\frac{9}{100}m}~. \] \hfill $\Box$ \subsection{Learning DNFs} In this section we prove theorem \ref{thm:dnf_few_clauses} and corollary \ref{cor:dnf_small}. We will use the $\mathrm{SRCSP}$ assumption \ref{hyp:sat} with Huang's predicate \citep{huang2013approximation}. Let $k\ge 1$ and denote $K=k+\binom k3$. We index the first $k$ coordinates of vectors in $\{\pm 1\}^{K}$ by the numbers $1,2,\ldots,k$. The last $\binom{k}{3}$ coordinates are indexed by $\binom{[k]}{3}$. Let $H_k:\{\pm 1\}^{K}\to \{0,1\}$ be the predicate such that $H_k(x)=1$ if and only if there is a vector $y$ with hamming distance $\le k$ from $x$ such that, for every $A\in \binom{[k]}{3}$, $y_A=\prod_{i\in A}y_i$. The basic properties of $H_k$ are summarized in the following lemma due to \citep{huang2013approximation}. \begin{lemma}[\citep{huang2013approximation}]\label{lem_Huang} \ \begin{enumerate} \item $H_k$ is heredity approximation resistant on satisfiable instances. \item $|H_k^{-1}(1)|=\tilde{O}(K^{1/3})$. \item The $0$-variability of $H_k$ is $\ge k$. \item For every sufficiently large $k$, there exists $y^k\in\{\pm 1\}^{K}$ such that $H_k(x)=1\Rightarrow H_k(y^k\oplus x)=0$ \end{enumerate} \end{lemma} {\par\noindent {\bf Proof}\space\space} 1. and 2. were proved in \cite{huang2013approximation}. 3. is very easy. We proceed to 4. Choose $y^k\in \{\pm 1\}^K$ uniformly at random. By 2., for every $x\in \{\pm 1\}^K$, $\Pr\left(H_k(y^k\oplus x)=1\right)=2^{-K+\tilde{O}(K^{1/3})}$. Taking a union over all vectors $x\in H_k^{-1}(1)$, we conclude that the probability that one of them satisfies $H_k(y^k\oplus x)=1$ is $2^{-K+\tilde{O}(K^{1/3})+\tilde{O}(K^{1/3})}=2^{-K+\tilde{O}(K^{1/3})}$. For large enough $k$, this is less than $1$. Therefore, there exists a $y^k$ as claimed. \hfill $\Box$ {\par\noindent {\bf Proof}\space\space} (of theorem \ref{thm:dnf_few_clauses}) Let $d>0$ by assumption \ref{hyp:sat} and lemma \ref{lem_Huang}, for large enough $k$, it is $\mathrm{SRCSP}$-hard to distinguish between satisfiable instances to $\mathrm{CSP}(H_k)$ and random instances with $m=2n^d$ constraints. We will reduce this problem to the problem of distinguishing between a realizable sample to $\mathrm{DNF}^{q(n)}$ and a random sample drawn from a $\left(\frac{9}{50}n^d,1/5\right)$-scattered ensamble ${\cal D}$. Since $d$ is arbitrary, the theorem follows from theorem \ref{thm:basic_realizable}. The reduction works as follows. Let $y^k$ be the vector from lemma \ref{lem_Huang}. Given an instance \[ J=\{H_k(j_{1,1}x_{i_{1,1}},\ldots,j_{1,K}x_{i_{1,K}}), \ldots,H_k(j_{m,1}x_{i_{m,1}},\ldots,j_{m,K}x_{i_{m,K}})\} \] to $\mathrm{CSP}(H_k)$, we will produce a new instance $J'$ by changing the sign of the variables according to $y^k$ in every other constraint. Namely, \begin{eqnarray*} J'&=&\{H_k(j_{1,1}x_{i_{1,1}},\ldots,j_{1,K}x_{i_{1,K}}), H_k(y^k_1j_{2,1}x_{i_{2,1}},\ldots,y^k_Kj_{2,K}x_{i_{2,K}}), \ldots \\ &&\ldots, H_k(j_{m-1,1}x_{i_{m-1,1}},\ldots,j_{m-1,K}x_{i_{m-1,K}}), H_k(y^k_1j_{m,1}x_{i_{m,1}},\ldots,y^k_Kj_{m,K}x_{i_{m,K}})\}~. \end{eqnarray*} Note that if $J$ is random then so is $J'$. Also, if $J$ is satisfiable with a satisfying assignment $u$, then, by lemma \ref{lem_Huang}, $u$ satisfies in $J'$ exactly the constraints with odd indices. Next, we will produce a sample $S\in\left(\{\pm 1\}^{2Kn}\times \{0,1\}\right)^{m}$ from $J'$ as follows. We will index the coordinates of vectors in $\{\pm 1\}^{2Kn}$ by $[K]\times \{\pm 1\}\times[n]$. We define a mapping $\Psi$ from the collection of $H_k$-constraints to $\{\pm 1\}^{2Kn}$ as follows -- for each constraint $C=H_k(j_1x_{i_1},\ldots,j_Kx_{i_K})$ we define $\Psi(C)\in \{\pm 1\}^{2Kn}$ by the formula \[ \left(\Psi(C)\right)_{l,b,i}=\begin{cases} -1 & (b,i)=(-j_l,i_l) \\ 1 & \textrm{otherwise} \end{cases} \] Finally, if $J'=\{C'_1,\ldots,C'_m\}$, we will produce the sample \[ S=\left\{(\Psi(C'_1),1),(\Psi(C'_2),0),\ldots,(\Psi(C'_{m-1}),1),(\Psi(C'_m),0)\right\} ~. \] The theorem follows from the following claim: \begin{claim} \ \begin{enumerate} \item If $J$ is a random instance then $S$ is $\left(\frac{9}{100}m,\frac{1}{5}\right)$-scattered. \item If $J$ is a satisfiable instance then $S$ is realizable by a $\mathrm{DNF}$ formula with $\le 2^K$ clauses. \end{enumerate} \end{claim} Proposition \ref{prop:scat_crit} implies part 1. We proceed to part 2. Like every boolean function on $K$ variables, $H_k$ is expressible by a $\mathrm{DNF}$ expression of $2^K$ clauses, each of which contains all the variables. Suppose then that \[ H_k(x_1,\ldots,x_K)=\vee_{t=1}^{2^K}\wedge_{r=1}^Kb_{t,r}x_r~. \] Let $u\in \{\pm 1\}^n$ be an assignment to $J$. Consider the following $\mathrm{DNF}$ formula over $\{\pm 1\}^{2Kn}$ \[ \phi_{u}(x)=\vee_{t=1}^{2^K}\wedge_{r=1}^K\wedge_{i=1}^n x_{r,(u_i b_{t,r}),i}~, \] where, as mentioned before, we index coordinates of $x \in \{\pm 1\}^{2Kn}$ by triplets in $[K] \times \{\pm 1\} \times [n]$. We claim that for every $H_k$-constraint $C$, $\phi_u(\Psi(C))=C(u)$. This suffices, since if $u$ satisfies $J$ then $u$ satisfies exactly the constraints with odd indices in $J'$. Therefore, by the definition of $S$ and the fact that $\forall C, \phi_u(\Psi(C))=C(u)$, $\phi_{u}$ realizes $S$. Indeed, let $C(x)=H_k(j_1x_{i_1},\ldots,j_Kx_{i_K})$ be a $H_k$-constraint. We have \begin{align*} \phi_u(\Psi(C))=1 ~~&\iff \exists t\in[2^K]\,\forall r\in[K], i\in [n],\; (\Psi(C))_{r,(u_ib_{t,r}),i}=1 \\ &\iff \exists t\in[2^K]\,\forall r\in[K], i\in [n]\; (u_ib_{t,r},i)\ne(-j_r,i_r) \\ &\iff \exists t\in[2^K]\,\forall r\in[K],\; u_{i_r}b_{t,r}\ne -j_r \\ &\iff \exists t\in[2^K]\,\forall r\in[K],\; b_{t,r}= j_ru_{i_r} \\ &\iff C(u)=H_k(j_1u_{i_1},\ldots,j_Ku_{i_K})=1 ~. \end{align*} \hfill $\Box$ By a simple scaling argument we can prove corollary \ref{cor:dnf_small}. {\par\noindent {\bf Proof}\space\space} (of corollary \ref{cor:dnf_small}) By theorem \ref{thm:dnf_few_clauses}, it is $\mathrm{SRCSP}$-hard to learn $\mathrm{DNF}^{n}$. Since $\mathrm{DNF}^{n}\subset \mathrm{DNF}_{n^2}$, we conclude that it is $\mathrm{SRCSP}$-hard to learn $\mathrm{DNF}_{n^2}$. To establish the corollary, we note that $\mathrm{DNF}_{n^2}$ can be efficiently realized by $\mathrm{DNF}_{n^{\epsilon}}$ using the mapping $f:\{\pm 1\}^{n}\to \{\pm 1\}^{n^{\frac{2}{\epsilon}}}$ that pads the original $n$ coordinates with $n^{\frac{2}{\epsilon}}-n$ ones. \hfill $\Box$ \subsection{Agnostically learning halfspaces} {\par\noindent {\bf Proof}\space\space} (of theorem \ref{thm:halfspaces}) Let ${\cal H}$ be the hypothesis class of halfspaces over $\{-1,1,0\}^n$, induced by $\pm 1$ vectors. We will show that agnostically learning ${\cal H}$ is $\mathrm{SRCSP}$-hard. While we defined the class of $\mathrm{HALFSPACES}$ over instances in $\{\pm 1\}^n$, proving the hardness of learning ${\cal H}$ (which is defined over $\{-1,1,0\}^n$) suffices for our needs, since ${\cal H}$ can be efficiently realized by $\mathrm{HALFSPACES}$ as follows: Define $\psi:\{-1,1,0\}\to \{\pm 1\}^2$ by \[ \psi(\alpha)= \begin{cases} (-1,-1) & \alpha = -1 \\ (1,1) & \alpha = 1 \\ (-1,1) & \alpha = 0 \end{cases}~. \] Now define $\Psi:\{-1,1,0\}^n\to \{\pm 1\}^{2n}$ by \[ \Psi(x)=(\psi(x_1),\ldots,\psi(x_n))~. \] Also define $\Phi:\{\pm 1\}^n\to \mathbb \{\pm 1\}^{2n}$ by \[ \Phi(w)=(w_1,w_1,w_2,w_2,\ldots,w_n,w_n)~. \] It is not hard to see that for every $w\in \{\pm 1\}^n$ and every $x\in \{-1,1,0\}^n$, $h_w(x)=h_{\Phi(w)}(\Psi(x))$. Therefore, ${\cal H}$ is efficiently realized by $\mathrm{HALFSPACES}$. We will use assumption \ref{hyp:not_sat} with respect to the majority predicate $\maj_K:\{\pm 1\}^K\to \{0,1\}$. Recall that $\maj(x)=1$ if and only if $\sum_{i=1}^Kx_i>0$. The following claim analyses its relevant properties. \begin{claim}\label{claim:maj_pred} For every odd $K$, \begin{itemize} \item $\Uval(\maj_K) =1-\frac{1}{K+1}$. \item $\mathrm{VAR}_0(\maj_K)=\frac{K+1}{2}$. \end{itemize} \end{claim} {\par\noindent {\bf Proof}\space\space} It is clear that $\maj_K$ has $\frac{K+1}{2}$ $0$-variability. We show next that $\Uval(\maj_K)=1-\frac{1}{K+1}$. Suppose that $K=2t+1$. Consider the distribution ${\cal D}$ on $\{\pm 1\}^K$ defined as follows. With probability $\frac{1}{2t+2}$ choose the all zero vector, and with probability $\frac{2t+1}{2t+2}$ choose a vector uniformly at random among all vectors with $t+1$ ones. It is clear that $\E_{x\sim {\cal D}}[\maj_K(x)]=1-\frac{1}{2t+2}$. We claim that ${\cal D}$ is pairwise uniform, therefore, $\Uval(\maj_K)\ge 1-\frac{1}{2t+2}$. Indeed for every distinct $i,j\in [K]$, \begin{align*} &\Pr_{x\sim {\cal D}}\left((x_i,x_j)=(0,1)\right)=\Pr_{x\sim {\cal D}}\left((x_i,x_j)=(1,0)\right)=\frac{2t+1}{2t+2}\cdot \frac{t+1}{2t+1}\cdot \frac{t}{2t}=\frac{1}{4} ~, \\ &\Pr_{x\sim {\cal D}}\left((x_i,x_j)=(1,1)\right)=\frac{2t+1}{2t+2}\cdot \frac{t+1}{2t+1}\cdot \frac{t}{2t}=\frac{1}{4}~, \end{align*} and $\Pr_{x\sim {\cal D}}\left((x_i,x_j)=(0,0)\right)=\frac{1}{4}$. Next, we show that $\Uval(\maj_K)\le 1-\frac{1}{2t+t}$. Let ${\cal D}$ be a pairwise uniform distribution on $\{\pm 1\}^K$. We have $\E_{x\sim {\cal D}}\left[\sum_{i=1}^K\frac{x_i+1}{2}\right]=\frac{K}{2}$ therefore, by Markov's inequality, \[ \E_{x\sim {\cal D}}\left[\maj_K(x)\right]=\Pr_{x\sim {\cal D}}\left(\maj_K(x)=1\right)=\Pr_{x\sim {\cal D}}\left(\sum_{i=1}^K\frac{x_i+1}{2}\ge t+1\right)\le \frac{2t+1}{2(t+1)}~. \] Since this is true for every pairwise uniform distribution, $\Uval(\maj_K)\le \frac{2t+1}{2(t+1)}=1-\frac{1}{K+1}$. \hfill $\Box$ Fix $\alpha\ge 1$. We will use theorem \ref{thm:basic_agnostic} to show that there is no efficient algorithm that approximately agnostically learns ${\cal H}$ with approximation ratio of $\alpha$, unless the $\mathrm{SRCSP}$ assumption is false. Let $c>1$ and denote $\beta=\frac{1}{10\alpha}$. It suffices to show that there is a polynomial ensamble ${\cal D}=\{{\cal D}_n^{m(n)}\}_{n=1}^\infty$ that is $(\Omega(n^c),\alpha\beta+\frac{1}{n})$-scattered and it is $\mathrm{SRCSP}$-hard to distinguish between a ${\cal D}$-random sample and an $\beta$-almost realizable sample. By assumption \ref{hyp:not_sat} and claim \ref{claim:maj_pred}, for large enough odd $K$, it is $\mathrm{SRCSP}$-hard to distinguish between a random instances of $\mathrm{CSP}(\maj_K)$ with $m(n)= n^c$ constraints and instances with value $\ge 1-\beta$. Consider the following ensamble ${\cal D}=\{{\cal D}_n^{2m(n)}\}_{n=1}^\infty$: pick $m=m(n)$ independent uniform vectors $x_1,\ldots,x_m\in\{x\in\{-1,1,0\}^n\mid |\{i\mid x_i\ne 0\}|=K|\}$. Then, consider the sample $S=\{(x_1,1),(-x_1,0),\ldots,(x_m,1),(-x_m,0)\}$. The theorem follows from the following claim: \begin{claim}~ \begin{itemize} \item ${\cal D}$ is $(\Omega(n^c),\alpha\beta+\frac{1}{n})$-scattered. \item It is $\mathrm{SRCSP}$-hard to distinguish between a ${\cal D}$-random sample and an $\beta$-almost realizable sample. \end{itemize} \end{claim} {\par\noindent {\bf Proof}\space\space} The first part follows from proposition \ref{prop:scat_crit}. Next, we show that it is $\mathrm{SRCSP}$-hard to distinguish between a ${\cal D}$-random sample and $\beta$-almost realizable sample. We will reduce from the problem of distinguishing between a random instance with $m(n)$ constraints and an instance with value $\ge 1-\beta$. Given an instance $J$ with $m(n)$ constraints, we will produce a sample $S$ of $2m(n)$ examples by transforming each constraint into two examples as follows: for the constraint $C(x)=\maj(j_1x_{i_1},\ldots,{j_K}x_{i_K})$ we denote by $u(C)\in \{x\in\{-1,1,0\}^n\mid |\{i\mid x_i\ne 0\}|=K|\}$ the vector whose $i_{l}$ coordinate is $j_l$. We will produce the examples $(u(C),1)$ and $(-u(C),0)$. It is not hard to see that if $J$ is random then $S\sim {\cal D}^{2m(n)}_n$. If the value of $J$ is $\ge 1-\beta$, indicated by an assignment $w\in \{\pm 1\}^n$, it is not hard to see that $h_w\in{\cal H}$ $\epsilon$-almost realizes the sample $S$. This concludes the proof of the claim. \hfill $\Box$ Combining all the above we conclude the proof of theorem \ref{thm:halfspaces}. \hfill $\Box$ \begin{remark}\label{rem:halfsapces} What is the real approximation ratio of agnostically learning halfspaces in $n$ dimension? Taking a close look at the above proof, we see that in some sense, by the $\mathrm{SRCSP}$ assumption with $\maj_{K}$, it is hard to agnostically learn halfspaces with approximation ratio of $\Omega\left(K\right)$. If we let $K$ grow with $n$ (this is {\em not} allowed by the $\mathrm{SRCSP}$-hypothesis), say $K=\frac{1}{100}n$, we can hypothesize that it is hard to agnostically learn halfspaces with approximation ratio of about $n$. The approximation ratio of the best known algorithms is somewhat better, namely, $\frac{n}{\log(n)}$. But this is not very far from our guess. Therefore, one might hypothesize that the best possible approximation ratio is, say, of the form $\frac{n}{\poly\left(\log(n)\right)}$. Given a rigorous treatment to the above intuition is left as an open question. \end{remark} \begin{remark}\label{rem:large_margin} The problem of {\em learning large margin halfsapces} is an important variant of the problem of learning halfspaces. Here, we assume that the instance space is the unit ball in $\mathbb R^d$. For $1>\gamma>0$, the $\gamma$-margin error of a hyperplane $h$ is the probability of an example to fall on the wrong side of $h$ or at a distance $\le\gamma$ from it. The $\gamma$-margin error of the best $h$ (with respect to a distribution ${\cal D}$) is denoted $\Err_\gamma(\mathcal{D})$. An $\alpha(\gamma)$-approximation algorithm receives $\gamma,\epsilon$ as input and outputs a classifier with error rate $\le \alpha(\gamma)\Err_\gamma(\mathcal{D}) + \epsilon$. Such an algorithm is efficient if it uses $\poly(\frac{1}{\gamma},\frac{1}{\epsilon})$ samples and runs in time polynomial in the sample size. For a detailed definition, the reader is referred to \cite{danielyLinSha13The}. It is not hard to see that the proof of theorem \ref{thm:halfspaces} shows that it is hard to approximately learn large margin halfspaces with any constant approximation ratio. Taking considerations as in remark \ref{rem:halfsapces}, one might hypothesize that the correct approximation ratio for this problem is about $\frac{1}{\gamma}$. As in the case of learning halfspaces, best known algorithms \cite{LongSe11, BirnbaumSh12} do just a bit better, namely, they have an approximation ratio of $\frac{1}{\gamma\sqrt{\log(1/\gamma)}}$. Therefore, one might hypothesize that the best possible approximation ratio is $\frac{1}{\gamma\poly\left(\log(1/\gamma)\right)}$. We note that a recent result \cite{danielyLinSha13The} shows that this is the best possible approximation ratio, if we restrict ourselves to a large class of learning algorithms (that includes SVM with a kernel, regression, Fourier transform and more). \end{remark} \subsection{Learning automata} For a function $q:\mathbb N\to\mathbb N$, let $\mathrm{AUTO}_{q(n)}$ be the class of functions $h:\{\pm 1\}^n\to \{0,1\}$ that can be realized by a finite automaton with $q(n)$ states. \begin{theorem}\label{thm:auto} For every $\epsilon>0$, it is $\mathrm{SRCSP}$-hard to learn $\mathrm{AUTO}_{n^\epsilon}$. \end{theorem} \begin{note} The theorem remains true (with the same proof), even if we restrict to acyclic automata. \end{note} {\par\noindent {\bf Proof}\space\space} By a simple scaling argument, as in the proof of corollary \ref{cor:dnf_small}, it is enough to show that it is $\mathrm{SRCSP}$-hard to learn $\mathrm{AUTO}_{n^2+1}$. By theorem \ref{thm:dnf_few_clauses}, it is $\mathrm{SRCSP}$-hard to learn $\mathrm{DNF}^{\log_2(n)}$. To establish the theorem, we will show that if a function $h:\{\pm 1\}^n\to\{0,1\}$ can be realized by a $\mathrm{DNF}$ formula with $\log_2(n)$ clauses, then it can be realized by an automaton with $n^2+1$ states. For simplicity, assume that $n$ is a power of $2$. Given a $\mathrm{DNF}$ formula $R$ with $k:=\log_2(n)$ clauses, we will construct an acyclic automaton as follows. For each variable we will have $n$ states (corresponding to subsets of $[k]$). In addition, we will have a start state. From the start state, the automaton will jump to the state $(1,A)$, where $A$ is the set of the indices of all the clauses in $R$ that are not violated by the value of $x_1$. After reading $x_2$ the automaton will jump to the state $(2,A)$, where $A$ is the set of the indices of all the clauses in $R$ that are not violated by the values of $x_1$ and $x_2$. In this manner, after reading $x_1,\ldots, x_n$ the automaton will be at the state $(n,A)$, where $A$ is the set of the indices of all the clauses in $R$ that are satisfied by $x_1,\ldots,x_n$. The automaton accepts if and only if $A\ne \emptyset$. Clearly, this automaton calculates the same function as $R$. \hfill $\Box$ \subsection{Toward intersection of $4$ halfspaces}\label{sec:4half} For $1\le l\le k$ Consider the predicate $T_{k,l}:\{\pm 1\}^k\to \{0,1\}$ such that $T_{k,l}(x)=1$ if and only if $x$ has at least $l$ ones. For example, $T_{k,1}$ is the $\mathrm{SAT}$ predicate, $T_{k,\lfloor \frac{k}{2}\rfloor+1}$ is the $\mathrm{MAJ}$ predicate and $T_{k,k}$ is the $\mathrm{AND}$ predicate. Define $P_{k}:\left(\{\pm 1\}^{k}\right)^{8}\to\{0,1\}$ by \begin{eqnarray*} P_{K}(x^1,\ldots,x^8)&=&\left(\wedge_{j=1}^4 T_{k,\lceil \frac{k}{2}\rceil-1}(x^j)\right) \wedge \neg \left(\wedge_{j=5}^8 T_{k,\lceil \frac{k}{2}\rceil-1}(x^j)\right)~. \end{eqnarray*} \begin{proposition} There is $k_0$ such that for every odd $k\ge k_0$ we have \begin{enumerate} \item Assuming the unique games conjecture, $P_k$ is heredity approximation resistant. \item For some constant $1>\alpha>0$, it is $\mathbf{NP}$-hard to distinguish between satisfiable instances to $\mathrm{CSP}(P_k)$ and instances with value $\le \alpha$. \end{enumerate} \end{proposition} {\par\noindent {\bf Proof}\space\space} We start with part 1. By \citep{austrin2009approximation}, it suffices to show that there is a pairwise uniform distribution that is supported in $P_k^{-1}(1)$. Denote $Q(x^1,\ldots x^4)=\wedge_{j=1}^4 T_{k,\lceil \frac{k}{2}\rceil-1}(x^j)$ and $R(x^1,\ldots x^4)=\neg\left(\wedge_{j=1}^4 T_{k,\lceil \frac{k}{2}\rceil-1}(x^j)\right)$. Note that if ${\cal D}_Q$ is a pairwise uniform distribution that is supported in $Q^{-1}(1)$ and ${\cal D}_R$ is a pairwise uniform distribution that is supported in $R^{-1}(1)$, then ${\cal D}_{Q}\times {\cal D}_R$ is a pairwise uniform distribution that is supported in $P_k^{-1}(1)$. Therefore, it suffices to show that such ${\cal D}_Q$ and ${\cal D}_R$ exist. We first construct ${\cal D}_{Q}$. Let ${\cal D}_k$ be the following distribution over $\{\pm 1\}^k$ -- with probability $\frac{1}{k+1}$ choose the all-one vector and with probability $\frac{k}{k+1}$, choose at random a vector with $\lceil\frac{k}{2}\rceil-1$ ones (uniformly among all such vectors). By the argument of claim \ref{claim:maj_pred}, ${\cal D}_k$ is pairwise uniform. Clearly, the distribution ${\cal D}_Q={\cal D}_k\times{\cal D}_k\times{\cal D}_k\times{\cal D}_k$ over $\left(\{\pm 1\}^k\right)^4$ is a pairwise uniform distribution that is supported in $Q^{-1}(1)$. Next, we construct ${\cal D}_{R}$. Let $k_0$ be large enough so that for every $k\ge k_0$, the probability that a random vector from $\{\pm 1\}^k$ will have more than $\lceil \frac{k}{2}\rceil$ minus-ones is $\ge \frac{3}{8}$ (it is easy to see that this probability approaches $\frac{1}{2}$ as $k$ approaches $\infty$. Therefore, such $k_0$ exists). Now, let $Z\in \{0,1\}^4$ be a random variable that satisfies: \begin{itemize} \item $Z_1,\ldots,Z_4$ are pairwise independent. \item For every $1\le i\le 4$, $\Pr(Z_i=1)=\frac{3}{8}$. \item $\Pr(Z=(0,0,0,0))=0$. \end{itemize} In a moment, we will show that a random variable with the above properties exists. Now, let $B\subset \{\pm 1\}^k$ be a set with $|B|\ge \frac{3}{8}\cdot 2^k$ such that every vector in $B$ has more than $\lceil\frac{k}{2}\rceil$ minus-ones. Consider the distribution ${\cal D}_R$ of the random variable $(X^1,\ldots,X^4)\in \left(\{\pm 1\}^k\right)^4$ sampled as follows. We first sample $Z$, then, for $1\le i\le 4$, if $Z_i=1$, we choose $X^i$ to be a random vector $B$ and otherwise, we choose $X^i$ to be a random vector $B^c$. We note that since $Z_1,\ldots,Z_4$ are pairwise independent, $X^1,\ldots, X^4$ are pairwise independent as well. Also, the distribution of $X^i,\;1=1,\ldots,4$ is uniform. Therefore, ${\cal D}_R$ is pairwise uniform. Also, since $\Pr(Z=(0,0,0,0))=0$, with probability $1$, at least one of the $X^i$'s will have more than $\lceil\frac{k}{2}\rceil$ minus-ones. Therefore, ${\cal D}_R$ is supported in $R^{-1}(1)$. It is left to show that there exists a random variable $Z\in \{0,1\}^4$ as specified above. Let $Z$ be the random variable defined as follows: \begin{itemize} \item With probability $\frac{140}{192}$ $Z$ is a uniform vector with a single positive coordinate. \item With probability $\frac{30}{192}$ $Z$ is a uniform vector with $2$ positive coordinates. \item With probability $\frac{22}{192}$ $Z$ is a uniform vector with $4$ positive coordinates. \end{itemize} Clearly, $\Pr(Z=(0,0,0,0))=0$. Also, for every distinct $1\le i,j\le 4$ we have \[ \Pr(Z_i=1)=\frac{140}{192}\cdot\frac{1}{4}+\frac{30}{192}\cdot\frac{1}{2}+ \frac{22}{192}=\frac{3}{8} \] and \[ \Pr(Z_i=1,Z_j=1)=\frac{30}{192}\cdot\frac{1}{6}+ \frac{22}{192}=\left(\frac{3}{8}\right)^2~. \] Therefore, the other two specifications of $Z$ hold as well. We proceed to part 2. The reduction is quite simple and we only sketch it. By adding dummy variables, it is enough to prove that it is $\mathrm{NP}$-hard to distinguish between satisfiable instances of $\mathrm{CSP}(T_{k,\lceil\frac{k}{2}\rceil-1})$ and instances with value $\le \alpha$ for some constant $0<\alpha<1$. We will show somewhat stronger property, namely, that if $1\le l\le k-2$, then for some $0<\alpha<1$, it is $\mathrm{NP}$-hard to distinguish between satisfiable instances of $\mathrm{CSP}(T_{k,l})$ and instances with value $\le \alpha$. We will reduce from the problem of distinguishing between satisfiable instances to $3$-$\mathrm{SAT}$ and instances with value $\le \frac{8}{9}$. This problem is $\mathrm{NP}$-hard \citep{haastad2001some}. Given an instance $J$ to $3$-$\mathrm{SAT}$, we will produce an instance $R(J)$ to $\mathrm{CSP}(T_{k,l})$ as follows. Its variables would be the variables of $J$ together with some new variables. For every constraint $C(x)=j_1x_{i_1}\vee j_2x_{i_2}\vee j_3x_{i_3}$ in $J$, we will add $k+l-1$ new variables $x^C_1,\ldots,x^C_k$ and $y^C_4,\ldots,x^C_{l+2}$. These new variables will be used only in the new clauses corresponding to $C$. We will introduce the following constraints: we add the constraint $T_{k,l}(j_1x_{i_1},j_2x_{i_2},j_3x_{i_3},y^C_4,\ldots,y^C_{l+2},-x^C_{l+3},\ldots,-x^C_k)$. Also, for every $(j_1,\ldots,j_k)\in \{\pm 1\}^k$ with at most $(k-l)$ minus-ones we will add the constraint $T_{k,l}(j_1x^C_1,\ldots,j_kx^C_k)$. If $J$ is satisfiable, then $R(J)$ is satisfiable as well: simply set all new variables to $1$. On the other hand, if $\mathrm{VAL} (J)\le\frac{8}{9}$, then it is not hard to see that for every assignment to $R(J)$'s variables, for at least $\frac{1}{9}$ of $J$'s clauses, at least one of the new clauses corresponding to it will be unsatisfied. Since we introduce $\le 2^k$ constraints in $R(J)$ for each constraint in $J$, we conclude that $\mathrm{VAL}(R(J))\le 1-2^{-k}\frac{1}{9}$. Therefore, the theorem holds with $\alpha=1-2^{-k}\frac{1}{9}$. \hfill $\Box$ \begin{conjecture}\label{conj:inter_of_two} $P_k$ is heredity approximation resistant on satisfiable instances. \end{conjecture} \begin{theorem}\label{thm:inter_of_two} Assuming conjecture \ref{conj:inter_of_two}, it is $\mathrm{SRCSP}$-hard to learn $\mathrm{INTER}_{4}$ \end{theorem} {\par\noindent {\bf Proof}\space\space} (sketch) The proof goes along the same lines of the proof of theorem \ref{thm:halfspaces}. We will prove $\mathrm{SRCSP}$-hardness for learning intersections of two halfspaces over $\{-1,1,0\}^n$, induced by $\pm 1$ vectors. As in the proof of theorem \ref{thm:halfspaces}, $\mathrm{SRCSP}$-hardness of learning intersections of two halfspaces over the boolean cube follows from this. Fix $d>0$. It is not hard to check that $\mathrm{VAR}_0(P_k)\ge \lceil\frac{k}{2}\rceil-2$. Therefore, by conjecture \ref{conj:inter_of_two} and assumption \ref{hyp:sat}, for large enough odd $k$, it is $\mathrm{SRCSP}$-hard to distinguish between a random instance to $\mathrm{CSP}(P_k)$ with $n^d$ constraints and a satisfiable instance. We will reduce from this problem to the problem of distinguishing between a realizable sample and a random sample that is $(\Omega(m^d),\frac{1}{5})$-scattered. Since $d$ is arbitrary, the theorem follows. Given an instance $J$, we produce two examples for each constraint: for the constraint \begin{eqnarray*} C(x) &=&\left(\wedge_{q=1}^4 T_{k,\lceil \frac{k}{2}\rceil-1}(j_{q,1}x_{i_{q,1}},\ldots,j_{q,k}x_{i_{q,k}})\right) \\ && \;\;\;\;\;\;\;\;\;\wedge \neg \left(\wedge_{q=5}^8 T_{k,\lceil \frac{k}{2}\rceil-1}(j_{q,1}x_{i_{q,1}},\ldots,j_{q,k}x_{i_{q,k}})\right) \end{eqnarray*} we will produce two examples in $\{-1,1,0\}^{4n}\times \{0,1\}$, each of which has exactly $4k$ non zero coordinates. The first is a positively labelled example whose instance is the vector with the value $j_{q,l},\;1\le q\le 4,1\le l\le k$ in the $n(q-1)+i_{q,l}$ coordinate. the second is a negatively labelled example whose instance is the vector with the value $j_{q,l},\;5\le q\le 8,1\le l\le k$ in the $n(q-5)+i_{q,l}$ coordinate. It is not hard to see that if $J$ is satisfiable then the produced sample is realizable by intersection of four halfspaces: if $u\in \{\pm 1\}^n$ is a satisfying assignment then the sample is realized by the intersection of the $4$ halfspaces $\sum_{i=1}^{n} u_ix_{n(q-1)+i}\ge -1,\;\;q=1,2,3,4$. On the other hand, by proposition \ref{prop:scat_crit}, if $J$ is random instance with $n^d$ constraints, then the resulting ensamble is $(\Omega(n^d),\frac{1}{5})$ scattered. \hfill $\Box$ \subsection{Agnostically learning parity} For convenience, in this section the domain of hypotheses will be $\{0,1\}^n$ and the domain of predicates will be $\{0,1\}^K$ (instead of $\{\pm 1\}^n$ and $\{\pm 1\}^K$). For every $S\subset [n]$ define $\chi_S:\{0, 1\}^n\to \{\pm 1\}$ by $\chi_S(x)=\oplus_{i\in S}x_i$. Let $\mathrm{PARITY}$ be the hypothesis class consisting of all functions $\chi_S,\;S\subset [n]$. \begin{theorem}\label{thm:parity} For every constant $\alpha\ge 1$, it is $\mathrm{SRCSP}$-hard to approximately agnostically learn $\mathrm{PARITY}$ with an approximation ratio of $\alpha$. \end{theorem} {\par\noindent {\bf Proof}\space\space} Let $P_K:\{0, 1\}^K\to \{0,1\}$ be the parity predicate. That is, $P_K(x)=\oplus_{i=1}^Kx_i$. We first show that for $K\ge 3$, $\Uval(P_K)=1$. Indeed, a pairwise uniform distribution which is supported in $P_K^{-1}(1)$ is the following -- choose $(x_1,\ldots,x_{K-1})$ uniformly at random and then choose $x_{K}$ so that $P_K(x)=1$. Second, it is clear that $\mathrm{VAR}_0(P_K)=K$. Therefore, by assumption \ref{hyp:not_sat}, for every $\beta > 0$ and every $d$, for sufficiently large $K$, it is $\mathrm{SRCSP}$-hard to distinguish between instances to $\mathrm{CSP}(P_K)$ with value $\ge 1-\beta$ and random instances with $m^d$ constraints. Note that with the convention that the domain of $P_K$ is $\{0,1\}^K$, the constraints of instances to $\mathrm{CSP}(P_K)$ are of the form $C(x)=x_{i_1}\oplus\ldots \oplus x_{i_K}$ or $C(x)=x_{i_1}\oplus\ldots \oplus x_{i_K}\oplus 1$. We will reduce from the aforementioned problem to the problem of distinguishing between $\beta$-almost realizable sample and ${\cal D}$-random sample for a distribution ${\cal D}$ which is $\left(\Omega\left(n^d\right),\frac{1}{4}\right)$-scattered. Since both $\beta$ and $d$ are arbitrary, the theorem follows from theorem \ref{thm:basic_agnostic}. Given an instance $J$ to $\mathrm{CSP}(P_K)$, for each constraint $C(x)=x_{i_1}\oplus\ldots \oplus x_{i_K}\oplus b$ we will generate an example $(u_C,y_C)$ where $u_C$ is the vector with ones precisely in the coordinates $i_1,\ldots,i_K$ and $y_C=b$. It is not hard to verify that if $J$ is a random instance with $m^d$ constraints then the generated sample is $\left(\Omega\left(n^d\right),\frac{1}{4}\right)$-scattered. On the other hand, assume that the assignment $\psi\in \{0, 1\}^n$ satisfies $1-\beta$ fraction of the constraints. Consider the hypothesis $\chi_S$ where $S=\{i\mid x_i=1\}$. We have $\chi_S(x_C)=\oplus_{i\in S}(u_C)_i=\oplus_{q=1}^K\psi_{i_q}$. Therefore, $\psi$ satisfies $C$ if and only if $\chi_S$ is correct on $(u_C,y_C)$. Since $\psi$ satisfies $1-\beta$ fraction of the constraints, the generated sample is $\beta$-almost realizable. \hfill $\Box$ \section{Resolution lower bounds}\label{sec:res} In this section we prove theorem \ref{thm:res_main}. Let $P:\{0,1\}^K\to\{0,1\}$ be some predicate. Let $\tau=\{T_1,\ldots, T_r\}$ be a resolution refutation for a $\mathrm{CSP}(P)$ instance $J$. A basic parameter associated with $\tau$ is the {\em width}. The \emph{width} of a clause is the number of literals it contains, and the {\em width} of $\tau$ is $\mathrm{width}(\tau):=\max_{1\le i\le r}\mathrm{width}(T_i)$. We also define the width of an unsatisfiable instance $J$ to $\mathrm{CSP}(P)$ as the minimal width of a resolution refutation of $J$. Ben-Sasson and Wigderson \cite{BenWi99} have shown that if an instance to $\mathrm{CSP}(P)$ has a short resolution refutation, then it necessarily has a narrow resolution refutation. Namely, \begin{theorem}[\cite{BenWi99}]\label{thm:ben_sasson_wig} Let $J$ be an unsatisfiable instance to $\mathrm{CSP}(P)$. The length of every resolution refutation for $J$ is at least $2^{\Omega\left(\frac{\mathrm{width}^2(J)}{n}\right)}$. \end{theorem} Theorem \ref{thm:res_main} now follows from theorem \ref{thm:ben_sasson_wig} and the following two lemmas. \begin{lemma}\label{lem:width_criterion} Let $J$ be an unsatisfiable instance to $\mathrm{CSP}(P)$. Assume that for every subset $I$ of $l$ constraints from $J$, most of the constraints in $I$ have $\ge K-\mathrm{VAR}_0(P)-1$ variables that do not appear in any other constraint in $I$. Then $\mathrm{width}(J)\ge \frac{l}{6}$. \end{lemma} {\par\noindent {\bf Proof}\space\space} Let $\tau=\{T_1,\ldots,T_r\}$ be a resolution refutation to $J$. Define $\mu(T_i)$ as the minimal number $\mu$ such that $T_i$ is implied by $\mu$ constraints in $J$. \begin{claim}~ \begin{enumerate} \item $\mu(\emptyset)> l$. \item If $T_i$ is implied by $T_{i_1},T_{i_2},\;i_1,i_2<i$ then $\mu(T_i)\le \mu(T_{i_1})+\mu(T_{i_2})$. \end{enumerate} \end{claim} {\par\noindent {\bf Proof}\space\space} The second property clearly holds. To prove the first property, suppose toward a contradiction that $\mu(\emptyset)\le l$. It follows that there are $t\le l$ constraints $I\subset J$ that implies the empty clause, i.e., it is impossible to simultaneously satisfy all the constraints in $I$. By the assumption of the lemma, it is possible to choose an ordering $I=\{C_1,\ldots,C_t\}$ such that for every $1\le i\le t$, $C_i$ contains at least $K-\mathrm{VAR}_0(P)-1$ variables that do not appear in $C_1,\ldots,C_{i-1}$. Indeed, let us simply take $C_t$ to be a clause that contains at least $K-\mathrm{VAR}_0(P)-1$ variables that do not appear in the clauses in $I\setminus\{C_t\}$. Then, choose $C_{t-1}$ in the same way from $I\setminus\{C_t\}$ and so on. Now, let $\psi\in\{\pm 1\}^n$ be an arbitrary assignment that satisfies $C_1$. By the definition of $0$-variability, it is possible to change the values of the variables appearing in $C_2$ but not in $C_1$ to satisfy also $C_2$. We can continue doing so till we reach an assignment that satisfies $C_1,\ldots, C_t$ simultaneously. This leads to the desired contradiction. \hfill $\Box$ By the claim, and the fact that $\mu(C)=1$ for every clause that is implied by one of the constraints of $J$, we conclude that there is some $T_i$ with $\frac{l}{3}\le \mu=\mu(T_j)\le \frac{2l}{3}$. It follows that there are $\mu$ constraints $C_1,\ldots, C_\mu$ in $J$ that imply $T_j$, but no strict subset of these clauses implies $T_j$. For simplicity, assume that these constraints are ordered such that for every $1\le i\le \frac{\mu}{2}$, $C_i$ contains at least $K-\mathrm{VAR}_0(P)-1$ variables that do not appear in the rest of these constraints. The proof of the lemma is established by the following claim \begin{claim} For every $1\le i\le \frac{\mu}{2}$, $T_j$ contains a variable appearing only is $C_i$. \end{claim} {\par\noindent {\bf Proof}\space\space} Assume toward a contradiction that the claim does not hold for some $1\le i\le \frac{\mu}{2}$. Since no strict subset of $C_1,\ldots,C_\mu$ imply $T_j$, there is an assignment $\psi\in\{\pm 1\}^n$ such that for every $i'\ne i$, $C_{i'}(\psi)=1$ but $T_j(\psi)=0$. Since $C_1,\ldots,C_\mu$ imply $T_j$, we must have $C_i(\psi)=0$. Now, by the definition of $0$-variability, we can modify the values of the $K-\mathrm{VAR}_0(P)-1$ variables that appear only in $C_i$ to have a new assignment $\psi'\in\{\pm 1\}^n$ with $C_i(\psi')=1$. Since $T_j$ and the rest of the constraints do not contain these variables, we conclude that still for every $i'\ne i$, $C_{i'}(\psi)=1$ and $T_j(\psi)=0$. This contradicts the fact that $C_1,\ldots,C_\mu$ imply $T_j$. \hfill $\Box$ \hfill $\Box$ The next lemma shows that the condition in lemma \ref{lem:width_criterion} holds w.h.p. for a suitable random instance. For the sake of readability, it is formulated in terms of sets instead of constraints. \begin{lemma} Fix integers $k>r>d$ such that $r>\max\{17d,544\}$. Suppose that $A_1,\ldots, A_{n^d}\in \binom{[n]}{k}$ are chosen uniformly at random. Then, with probability $1-o_n(1)$, for every $I\subset [n^d]$ with $|I|\le n^{\frac{3}{4}}$ for most $i\in I$ we have $|A_i\setminus \cup_{j\in I\setminus \{i\}}A_j|\ge k-r$. \end{lemma} {\par\noindent {\bf Proof}\space\space} Fix a set $I$ with $2\le t\le n^{\frac{3}{4}}$ elements. Order the sets in $I$ arbitrarily and also order the elements in each set arbitrarily. Let $X_1,\ldots,X_{kt}$ be the following random variables: $X_1$ is the first element in the first set of $I$, $X_2$ is the second element in the first set of $I$ and so on till the $k$'th element of the last set of $I$. Denote by $R_i\;\;1\le i\le kt$ the indicator random variable of the event that $X_i=X_j$ for some $j<i$. We claim that if $\sum R_i< \frac{tr}{4}$, the conclusion of the lemma holds for $I$. Indeed, let $J_1\subset I$ be the set of indices with $R_i=1$, $J_2\subset I$ be the set of indices $i$ with $R_i=0$ but $X_i=X_j$ for some $j>i$ and $J=J_1\cup J_2$. If the conclusion of the lemma does not hold for $I$, then $|J|\ge \frac{tr}{2}$. If in addition $|J_1|=\sum R_i< \frac{tr}{4}$ we must have $|J_2|>\frac{tr}{4}>|J_1|$. For every $i\in J_2$, let $f(i)$ be the minimal index $j>i$ such that $X_i=X_j$. We note that $f(i)\in J_1$, therefore $f$ is a mapping from $J_2$ to $J_1$. Since $|J_2|>|J_1|$, $f(i_1)=f(i_2)$ for some $i_1<i_2$ in $J_2$. Therefore, $X_{i_1}=X_{f(i_1)}=X_{i_2}$ and hence, $R_{i_2}=1$ contradicting the assumption that $i_2\in J_2$. Note that the probability that $R_i=1$ is at most $\frac{tk}{n}$. This estimate holds also given the values of $R_1,\ldots, R_{i-1}$. It follows that the probability that $R_i=1$ for every $i\in A$ for a particular $A\subset I$ with $|A|=\lceil\frac{rt}{4}\rceil$ is at most $\left(\frac{tk}{n}\right)^{\frac{rt}{4}}$. Therefore, for some constants $C',C>0$ (that depend only on $d$ and $k$), the probability that $J$ fails to satisfy the conclusion of the lemma is bounded by \begin{eqnarray*} \Pr\left(\sum R_i\ge \frac{tr}{4}\right) &\le & \binom{tk}{\lceil\frac{tr}{4}\rceil}\left(\frac{tk}{n}\right)^{\frac{tr}{4}} \\ &\le & 2^{C\cdot t}\left(\frac{tk}{n}\right)^{\frac{tr}{4}} \\ &\le & 2^{C'\cdot t}\left(\frac{t}{n}\right)^{\frac{tr}{4}} \end{eqnarray*} The second inequality follows from Stirling's approximation. Summing over all collections $I$ of size $t$ we conclude that for some $C''>0$, the probability that the conclusion of the lemma does not hold for some collection of size $t$ is at most \[ \binom{n^d}{t}2^{C'\cdot t}\left(\frac{t}{n}\right)^{\frac{tr}{4}} \le n^{dt-\frac{1}{16}tr}\cdot2^{C'\cdot t} \le n^{-\frac{1}{272}tr}\cdot2^{C'\cdot t} \le n^{-2t}\cdot2^{C'\cdot t}\le C''\frac{1}{n} \] Summing over all $2\le t\le n^{\frac{3}{4}}$, we conclude that the probability that the conclusion of the lemma does not hold is at most $C'' n^{-\frac{1}{4}}=o_n(1)$. \hfill $\Box$ \section{On basing the $\mathrm{SRCSP}$ assumption on $\mathbf{NP}$-Hardness}\label{sec:base_on_np} Fix a predicate $P:\{\pm 1\}^{K}\to \{0,1\}$ and let $1\ge \alpha >\Lval(P)$. Let $L\subset \{0,1\}^*$ be some language. We say that $L$ can be efficiently reduced to the problem of distinguishing between random instances to $\mathrm{CSP}(P)$ with $Cn$ constraints and instances with value $\ge \alpha$, if there is an efficient probabilistic Turing machine that given $x\in \{0,1\}^n$, acts as follows: for some function $f:\mathbb N\to\mathbb N$, \begin{itemize} \item If $x\in L$ then $M(x)$ is an instance to $\mathrm{CSP}(P)$ with $f(n)$ variables, $C\cdot f(n)$ constraints and value $\ge \alpha$. \item If $x\notin L$ then $M(x)$ is a random instance to $\mathrm{CSP}(P)$ with $f(n)$ variables and $C\cdot f(n)$ constraints. \end{itemize} \begin{theorem}\label{thm:base_on_np} For every sufficiently large constant $C>0$, the following holds. Assume that the language $L\subset \{0,1\}^*$ can be efficiently reduced to the problem of distinguishing between random instances to $\mathrm{CSP}(P)$ with $m(n)\ge Cn$ constraints and instances with value $\ge \alpha$. Then, $L$ has a statistical zero knowledge proof. \end{theorem} \begin{corollary}\label{cor:base_on_np} For every sufficiently large constant $C>0$, the following holds. Assume that there is a reduction from an either an $\mathbf{NP}$-hard or $\mathbf{CoNP}$-hard problem to the problem of distinguishing between random instances to $\mathrm{CSP}(P)$ with $m(n)\ge Cn$ constraints and instances with value $\ge \alpha$. Then, the polynomial hierarchy collapses. \end{corollary} {\par\noindent {\bf Proof}\space\space} (of corollary \ref{cor:base_on_np}) Under the conditions of the corollary, by theorem \ref{thm:base_on_np}, we have $\mathbf{NP}\subset \mathbf{SZKP}$ or $\mathbf{CoNP}\subset \mathbf{SZKP}$. Since $\mathbf{SZKP}$ is closed under taking complement \citep{okamoto1996relationships}, in both cases, $\mathbf{NP}\subset \mathbf{SZKP}$. Since $\mathbf{SZKP}\subset \mathbf{CoAM}$ \citep{aiello1991statistical}, we conclude that $\mathbf{NP}\subset \mathbf{CoAM}$, which collapses the polynomial hierarchy \citep{bogdanov2006worst}. \hfill $\Box$ {\par\noindent {\bf Proof}\space\space} (of theorem \ref{thm:base_on_np}) Let $C>0$ be a constant large enough so that, with probability $\ge \frac{1}{2}$, a random instance to $\mathrm{CSP}(P)$ with $Cn$ constraints will have value $\le \alpha$. Consider the following problem. The input is a circuit $\Psi:\{0,1\}^n\to \{0,1\}^m$ and a number $t$. The instance is a YES instance if the entropy\footnote{We consider the standard Shannon's entropy with bits units.} of $\Psi$, when it acts on a uniform input sampled from $\{0,1\}^n$, is $\le t-1$. The instance is a NO instance if this entropy is $\ge t$. By \citep{goldreich1999comparing} this problem is in $\mathbf{SZKP}$. To establish the proof, we will show that $L$ can be reduced to this problem. Assume that there is a reduction from the language $L$ to the problem of distinguishing between random instances to $\mathrm{CSP}(P)$ with $m(n)\ge Cn$ constraints and instances with value $\ge \alpha$. Let $M$ and $f$ be a Turing machine and a function that indicate that. By a standard argument, it follows that there is an efficient deterministic Turing machine $M'$ that given $x\in \{0,1\}^n$ produces a circuit $\Psi$ whose input is $\{0,1\}^{g(n)}$ for some polynomially growing function and whose output is an instance to $\mathrm{CSP}(P)$, such that, for a uniformly randomly chosen input $z\in \{0,1\}^{g(n)}$, \begin{itemize} \item If $x\in L$ then $\Psi(z)$ is a (possibly random) satisfiable instance to $\mathrm{CSP}(P)$ with $f(n)$ variables and $m(f(n))$ constraints. \item If $x\notin L$ then $\Psi(z)$ is a random instance to $\mathrm{CSP}(P)$ with $f(n)$ variables and $m(f(n))$ constraints. \end{itemize} Since the number of instances to $\mathrm{CSP}(P)$ with $m(f(n))$ constraints is $\left(\binom{f(n)}{K}2^K\right)^{m(f(n))}$, in the second case, the entropy of $\Psi$ is $q(n):=m(f(n))\log_2\left(\binom{f(n)}{K}2^K\right)$. On the other hand, in the first case, the entropy is at most the entropy of a random instance to $\mathrm{CSP}(P)$ with $m(f(n))$ constraints and value $\ge \alpha$. By the choice of $C$, the number of such instances is at most half of the total number of instances with $m(f(n))$ constraints. Therefore, the entropy of $\Psi$ is at most $m(f(n))\log_2\left(\binom{f(n)}{K}2^K\right)-1=q(n)-1$. Hence, using $M'$, we can reduce $L$ to the problem mentioned in the beginning of the proof. \hfill $\Box$ \paragraph{Acknowledgements:} Amit Daniely is a recipient of the Google Europe Fellowship in Learning Theory, and this research is supported in part by this Google Fellowship. Nati Linial is supported by grants from ISF, BSF and I-Core. Shai Shalev-Shwartz is supported by the Israeli Science Foundation grant number 590-10. We thank Sangxia Huang for his kind help and for valuable discussions about his paper \citep{huang2013approximation}. We thank Guy Kindler for valuable discussions.
\section{Introduction\label{Introduction}} It is generally assumed that the atmospheres of gas giant planets are analogous to the atmospheres of old brown dwarfs because they are (1) approximately the same composition, (2) approximately the same radius ($\sim$1 R$_{jup}$ supported by electron degeneracy pressure) and (3) slowly cooling due to a lack of internal fusion. However, early studies of the handful of directly imaged exoplanets suggest that planets may be cloudier and more turbulent than their older and more massive analogs \citep{2008Sci...322.1348M,2010ApJ...723..850B,2011ApJ...729..128C,2011ApJ...733...65B,2011ApJ...732..107S,2011ApJ...735L..39B,2013Sci...339.1398K}. A particularly large discrepancy exists in the L-band (3-4$\micron$), where atmospheric models that have been successfully used to fit brown dwarfs underpredict the [3.3$\micron$] fluxes of the HR 8799 planets by 1-2 magnitudes \citep{2012ApJ...753...14S}. Much of what is currently known about the physical properties of exoplanets comes from near-infrared (1-2.5$\micron$) photometry and spectroscopy \citep{2004A&A...425L..29C,2007ApJ...657.1064M,2008Sci...322.1348M,2010Natur.468.1080M,2011ApJ...729..128C,2012ApJ...753...14S,2010A&A...517A..76P,2010ApJ...723..850B,2011ApJ...733...65B,2013ApJ...768...24O,2013Sci...339.1398K,2011A&A...528L..15B,2013A&A...555A.107B,2013arXiv1306.0610C,2013ApJ...774...11K,2013arXiv1310.4183J}, due in part to the difficulty of working at longer wavelengths from the ground. However, self-luminous exoplanets emit the majority of their photons in the mid-infrared $>$3$\micron$, which besides making the region interesting in its own right, implies that less contrast is necessary to detect them against the bright glare of their host stars \citep[see Figure \ref{mid-IR examples} and][]{2006ApJ...653.1486H}. While the first generation of directly-imaged exoplanets were relatively warm and massive \citep[2M1207 b, HR 8799 bcde and $\beta$ Pic b are all $>$800 K and $>$5 M$_{jup}$;][]{2011ApJ...732..107S,2008Sci...322.1348M,2010Natur.468.1080M,2011A&A...528L..15B}, systems like these are known to be rare, while low-mass, close-in planets are ubiquitous \citep{2013ApJ...773..179W,2013arXiv1306.1233N,2013arXiv1309.1462B,2011ApJ...736...19B,2013ApJS..204...24B}. Low-mass planets, core-accretion planets, and elderly planets will all be cooler than the early exoplanet discoveries \citep{2008ApJ...683.1104F}. And ``cool'' planets emit an even larger fraction of their light in the mid-infrared (see Figure \ref{mid-IR examples}). To prepare for imaging ``cool'' planets, and to understand the ``warm'' planets that have already been found, it is critical that we expand our knowledge of exoplanet SEDs into the mid-infrared. In this paper, we present deformable secondary AO\footnote{These adaptive optics systems employ the minimum number of warm optics, minimizing the thermal infrared background from the telescope \citep{2000PASP..112..264L}.} imaging of the HR 8799 system, using LBTI/LMIRCam, and the 2M1207 system, using MagAO/Clio2. HR 8799 is a 20-160 Myr A5V star \citep{1969AJ.....74..375C,2006ApJ...644..525M,2008Sci...322.1348M,2010ApJ...716..417H,2011ApJ...732...61Z} with four directly imaged planets \citep{2008Sci...322.1348M,2010Natur.468.1080M}, whose masses and separations are difficult to explain with standard planet formation models \citep{2009ApJ...707...79D,2010ApJ...710.1375K}. 2M1207 is a 5-13 Myr M8 brown dwarf \citep{2002ApJ...575..484G} with a planetary-mass companion \citep{2004A&A...425L..29C} that is also difficult to explain with standard planet-formation models \citep{2005A&A...438L..25C}. Ignoring system architectures, HR 8799 bcde and 2M1207 b are unambiguously low-gravity, planetary-mass objects, whose atmospheres offer the first opportunities to characterize directly-imaged planets. And since all five planets have luminosities consistent with L$\rightarrow$T transition brown dwarfs, their atmospheres are ideal laboratories for studying how cloud properties are affected by a low-gravity, planetary environment. For HR 8799 we build upon previous [3.3$\micron$] imaging \citep{2012ApJ...753...14S} by using 6 narrow-band filters in the 3-4$\micron$ window to probe the spectral shape of the 3.3$\micron$ methane fundamental absorption feature. For 2M1207 b, we present photometry in the broader [3.3$\micron$] filter to determine if the object has the same extreme 3.3-3.8$\micron$ colors first seen in the HR 8799 planets. We present our observations, reductions, and photometry in Section 2, a comparison with field brown dwarfs in Section 3, SED modeling in Section 4, and our conclusions in Section 5. We also include filter curves and their tabulated properties in Appendix A. \section{Observations and Reductions\label{Observations and Reductions}} \subsection{HR 8799\label{HR 8799}} We observed the HR 8799 planetary system in six 5\% bandwidth filters from 3.0-3.8$\micron$ (see filter properties in Appendix A) on UT Nov. 2, 2012 using LBTI \citep{2012SPIE.8445E..0UH} and its 1-5$\micron$ camera LMIRCam \citep{2010SPIE.7735E.118S,2012SPIE.8446E..4FL}. The LMIRCam detector is a Hawaii-2RG, 5$\micron$-doped HgCdTe device, with 32 readouts. Currently we use 16 readouts, each of which reads 1024 rows by 64 columns, giving us a usable region of 1024x1024 (11x11''). Images from both LBT telescopes were corrected by the LBT's two deformable secondary adaptive optics systems\footnote{The LBTI wavefront sensors are functionally equivalent to the First Light Adaptive Optics Sytem \citep[FLAO][]{2010SPIE.7736E...7E}} and incoherently overlapped in LBTI's beam-combiner. The resulting images at the focal plane of LMIRCam achieved the diffraction-limited performance of a single LBT 8.4 meter aperture and the collecting area of the full 2$\times$8.4 meter LBT. Ground-based high-contrast observations are usually limited by instrumental quasi-static speckles, which can be removed by allowing the astronomical field to rotate with parallactic angle, while keeping the instrument rotation fixed \citep[][angular differential imaging]{2006ApJ...641..556M}. In this approach, sky-rotation and clock-time are the observational requirements, rather than integration time. For our observations of HR 8799, we switched filters every $\sim$60 seconds, rotating through the set and nodding every $\sim$10 minutes. By doing this, we were able to achieve adequate sky-rotation and clock-time in multiple filters simultaneously. We acquired seven minutes of data in each filter (42 minutes total) over a period of 2 hours (with a 1 hour gap due to a telescope malfunction), during which time the parallactic angle changed by 70 degrees. There were occasional scattered clouds during the night, but the adaptive optics wavefront sensor counts and thermal sky-background stayed consistent throughout the observations. The LBT's differential image motion monitor (DIMM) measured a natural seeing of 0.8" and the nearby Submillimeter Telescope Tau-meter measured a precipitable water vapor of $\sim$5-6mm. Our images were taken in correlated double sampling mode (reset-read-integrate-read) so that the first read (0.029 seconds) could be used as an unsaturated image of the star, while the second read (12.1, 12.1, 6.1, 4.0, 6.1, 3.0 seconds for the six filters respectively) saturated the star and filled the wells into the photon-noise regime away from the star. \subsubsection{Detector Non-Linearity\label{Non-Linearity}} Throughout Fall 2012, LMIRCam suffered from detector non-linearity, caused by an incorrectly set bias voltage, which has since been re-tuned. We constructed fluence-to-count calibration curves (linearity curves) by taking sky flats of varying integration times. Typical detector non-linearity is characterized by decreasing gain with well filling. LMIRCam's non-linearity had an S-shape where gain increased with well-filling before entering a linear regime and then turning over as the array reached saturation. Over the course of the semester, the detector linearity did change slightly, but from night-to-night, it was quite consistent. For the data reduction in this paper, we combined linearity from consecutive nights (UT Nov. 2, 2012 and UT Nov. 3 2012) to improve sampling, allowing an overall change in sky flux as a free parameter. The pixel-to-pixel linearity curves were mostly consistent (to within $\sim$5\%) after subtracting biases and correcting for flat-fielding effects. Pixels with linearity curves that were inconsistent with the rest of the array were classified as bad pixels (10\% of the array but mostly concentrated in a region that we avoided during our observations). Using the good pixels, we constructed a median linearity curve and scaled it, for each pixel, based on a flat-field. We used the set of curves to pre-process both correlated double sampling reads of every frame, individually. After reducing all of our data (see Section \ref{LBTI Reductions and Photometry}), the difference in the inferred planet photometry with and without the linearity correction is $\sim$40\%. To assess the quality of this correction, we took images of HR 8799 with various exposure times, ranging from our minimum integration time to the point where the stellar core began to saturate. After linearizing these data, the measured stellar flux remained constant to within 3\%, which validates our linearity correction. \subsubsection{Reductions and Photometry\label{LBTI Reductions and Photometry}} Our science data is comprised of alternating short and long exposures (correlated double sampling) in a series of six filters. For the short-integration images, we subtracted similar integration dark frames, taken at the end of the night, to remove pixel-to-pixel biases (which persist for long time-scales). We then removed the more rapidly changing residual channel biases (which are constant over each readout) by subtracting the median of the 64x8 pixel overscan region from each pixel in the corresponding 64x1016 light-sensitive region. We then repeated the bias subtraction steps on the long exposure images. Both sets of images were linearized and flattened. We then nod-subtracted the images, and subtracted the sigma-clipped median of each column to remove any residual electronic effects. The images were binned, 2x2, to suppress bad pixels. Known bad pixels (from flat-field maps) were masked in the process and 2x2 blocks of bad pixels were replaced with the average of surrounding pixels. The binned pixels have a plate scale of 0.0214''/pixel, which still over-samples the LBT's single-aperture diffraction-limited PSF (0.07'' FWHM) at our shortest wavelength (3$\micron$). The short and long integration images were registered independently by cross-correlation in Fourier space. The unsaturated (star) images were median combined with $3\sigma$ clipping. The saturated images were processed for angular differential imaging (ADI). For each image, the stellar centroid was determined from the Airy rings, and confirmed to be accurate by reducing the first half and second half of our data independently and verifying that the planets did not move. We then subtracted the stellar profile by fitting an azimuthal profile at every radius bin with the first four low-order modes of a fast-Fourier transform. Next, we subtracted the median of the stack of images with adequate sky rotation (a parallactic angle that has changed $>\lambda/D)$ at a given radius). Finally, we rotated and combined the images with a $3\sigma$ clipped median. The final reduced images are shown in Figure \ref{HR8799_image}. Planets c and d are easily detected in all six images. Planet b is detectable in four of the six images, given its known position, although its S/N is insufficient for detailed photometric modeling. The noise in the vicinity of planet e is too high to detect the planet in individual narrow-band images (it is detected when the images are coadded). It is possible planet e could be detected in individual narrow-band images using more sophisticated high-contrast algorithms, such as LOCI or PCA. However, the detections would not produce photometry that is adequate for this paper's analysis. We calibrated the photometry of planets c and d by adding artificial negative planets to the positions of the planets before the ADI step \citep[as described in ][]{2012ApJ...753...14S}. This is necessary to calibrate out the self-subtraction that is intrinsic to high-contrast imaging pipelines, such as ADI. Our absolute photometry, assuming a stellar apparent magnitude of 5.22 \citep[all bands;][]{2010ApJ...716..417H} and a distance modulus of 2.98$\pm$0.06\citep{2007A&A...474..653V}, is presented in Table \ref{LBT/Magellan}. Errors were determined empirically from the areas surrounding the planets by inserting artificial negative planets with a range of brightnesses at a precision of 0.05 mags. The values obtained are consistent with the broad-band measurements of \citet{2008Sci...322.1348M} and \citet{2012ApJ...753...14S} (see Tables \ref{LBT/Magellan} and \ref{lit phot}). \subsection{2M1207\label{2M1207}} We observed 2MASS J12073346-3932539 (2M1207) in a 12\% bandwidth 3.3$\micron$ (see filter information in Appendix A) with the Magellan adaptive optics system \citep[MagAO;][]{2012SPIE.8447E..0XC} and its 14x28'' FOV near-infrared imager, Clio2 \citep{2006SPIE.6269E..27S,2004SPIE.5492.1561F} on UT April 14, 2013. Conditions were photometric, with 1.0'' seeing (from DIMM measurements) and 4.3 mm precipitable water vapor (from the nearby and similar altitude La Silla weather station). Previous AO imaging of 2M1207 \citep{2004A&A...425L..29C,2005A&A...438L..25C,2007ApJ...657.1064M} has used VLT/NACO's infrared wavefront sensor \citep{2003SPIE.4839..140R}, which takes advantage of 2M1207 A's very red (V-K=7) color. The Magellan AO system has a visible light pyramid wavefront sensor \citep{2011SPIE.8149E...1E}, which is able to lock on fainter targets than most typical Shack-Hartmann systems; however, it cannot lock on 2M1207 itself (R$\sim$19.3). As a result, we locked on a nearby guide-star, 2MASS J12073100-3932281, which is 38.4'' offaxis and has an R magnitude of $\sim$14.2. The resulting off-axis correction was sufficient to resolve the $\sim$0.8'' binary, 2M1207 A-b. Clio2 was mounted on the Nasmyth port of Magellan-II (Clay), with the rotator turned on to keep 2M1207 from moving with respect to the off-axis guide star. We obtained 529 [3.3$\micron$] images of 2M1207 with 1.5 second detector integration times and 4 coadds, totaling 53 minutes of integration time over a period of 2 hours. The inefficiencies were caused by concurrent engineering work, since this was the first time the system had been used to track an off-axis target through transit at such a high elevation. Clio2 was configured in its coarse plate-scale mode of 27 mas/pixel. We nodded back-and-forth by 3'' in an ABBA pattern to subtract the background, and dithered around the chip to aid the removal of bad pixels. \subsubsection{Reductions and Photometry\label{MagAO Reductions and Photometry}\label{magaoreductions}} We processed the data by flat-fielding, interpolating over bad pixels, nod-subtracting the frames, and registering the images. We frame-selected the best 80\% of the images based on cross-correlation to remove frames with poor image quality. We then averaged the frames together, which produced a higher S/N detection of 2M1207 b than a median combination. The final image is shown in Figure \ref{2M1207_image}. 2M1207 A and b are clearly resolved. To cleanly separate 2M1207 A and b for photometry, we subtracted a 5-pixel gaussian smoothed image of the binary from itself (unsharp masking). This adequately removed 2M1207 A's halo from 2M1207 b's position leaving `b' on a flat background. We used the IDL Astronomy User's Library's \textit{DAOPHOT-Type} procedures\footnote{http://idlastro.gsfc.nasa.gov/} to perform PSF-fitting photometry, using 2M1207 A as a PSF for 2M1207 b (both unsharp masked). The resulting PSF-subtracted image had a small negative residual ($\sim$0.2 mags), which we calibrated, along with our statistical uncertainty, using aperture photometry. For the aperture photometry, we used a 1.5$\lambda$/D aperture on-source (which was background-noise limited after unsharp masking) and around the array to determine statistical errors. We find that 2M1207 b is 3.86$\pm$0.10 mags fainter than 2M1207 A in the [3.3$\micron$] filter. Our hybrid PSF-fitting/aperture photometry approach was meant to ensure that an elongated PSF (due to anisoplanicity) did not bias our aperture photometry. However, we note that subtracting the image from itself after a 180 degree rotation (which removes 2M1207 A but not b), and then doing aperture photometry produces a result that is consistent at 1$\sigma$. Because Clio2's 3.3$\micron$ filter is very sensitive to variable telluric water vapor, we did our absolute calibration using data from the space-based Wide-field Infrared Survey Explorer \citep[WISE;][]{2010AJ....140.1868W}. In the WISE1 filter (3.4$\micron$), 2M1207 is 11.556$\pm$0.023 mags. Subtracting the contribution of 2M1207 b ($\Delta$mag$\sim$3.86 at 3.3$\micron$, 2M1207 A is 11.59$\pm$0.02 mags in the WISE1 filter. To convert the WISE1 photometry to our [3.3$\micron$] bandpass, we calculated the [3.3$\micron$]-WISE1 colors\footnote{We used the 3.3$\micron$ filter curve from the manufacturer (JDSU; see Appendix A), the WISE1 filter curve from http://www.astro.ucla.edu/\textasciitilde wright/WISE/passbands.html, a model Alpha Lyr spectrum from \citet{1992AJ....104.1650C}, and a telluric transmission function (1.0 airmasses, 4.3 mm precipitable water vapor at Cerro Pachon) from http://www.gemini.edu/sciops/telescopes-and-sites/observing-condition-constraints/ir-transmission-spectra.} of a range of DUSTY model atmospheres \citep{2001ApJ...556..357A} that have been used to fit 2M1207 A \citep{2007ApJ...657.1064M,2012A&A...540A..85P}. From 2400 K to 3100 K and log(g) from 4.0 to 5.5, [3.3$\micron$]-WISE1=0.01 with a range of $<$0.01. Therefore, 2M1207 A is 11.60$\pm$0.03 and 2M1207 b is 15.46$\pm$0.10 in the [3.3$\micron$] filter. Using a weighted average of parallax measurements \citep[52.8pc$\pm$1.0pc; ]{2007ApJ...669L..41B,2007ApJ...669L..45G,2008AA...477L...1D}, we find that 2M1207 b has an absolute [3.3$\micron$] magnitude of 11.85$\pm$0.14. A summary of our photometry is presented in Table \ref{LBT/Magellan}. \section{Mid-Infrared Colors of Exoplanets\label{mircmd}} While there are still relatively few examples of directly imaged exoplanets, hundreds of brown dwarfs have been observed throughout the optical and infrared, and their properties are relatively well characterized compared to exoplanets. Because brown dwarfs and gas-giants share similar compositions, temperatures, and radii, the large sample of brown dwarfs is a baseline with which to compare directly imaged exoplanets. In Figure \ref{cmd} we plot color-magnitude diagrams of brown dwarfs and directly-imaged exoplanets to illustrate similarities and differences in their appearances. The brown dwarf data is assembled from \citet{2012ApJS..201...19D}, who include 2MASS photometry \citep{2006AJ....131.1163S}, WISE photometry \citep{2010AJ....140.1868W} and, when available, L' photometry \citep{1998ApJ...509..836L,2001ApJ...548..908L,2002ApJ...564..452L,2004AJ....127.3516G,2007ApJ...655.1079L,2010MNRAS.408L..56L} for a large sample of M, L and T-type brown dwarfs with parallax measurements. We use \citet{2012ApJS..201...19D}'s set of ``normal'' brown dwarfs (excluding low-gravity, moving group members, and low-metallicity), removing a small fraction that have large photometric errors ($>$0.1 in color). For the center and right color-magnitude plots ($\rm M_{L'}$ vs. [3.3$\micron$]-L' and $\rm M_{M}$ vs. L'-M) we use WISE band 1 and 2 measurements for the brown dwarf sequence and color correct\footnote{The color correction comes from fitting the complete set of BT-SETTL models \citep{2012RSPTA.370.2765A}, which cover a wide range of temperatures, and hence fully sample 3.3-WISE1 and M-WISE2 slopes. We find that $[3.3\micron]-L'=-0.001+0.885\times(WISE1-L')+0.462567\times(WISE1-L')^2$ over a range of [0.0,2.0] in WISE1-L' (with corresponding polynomial errors of 0.003, 0.016 and 0.009), and $L-M'=-0.020+1.177\times(L'-WISE2)-0.002\times(L'-WISE2)^2$ over a range of [-1.0,1.5] in L'-WISE2 (with corresponding polynomial errors of (0.002, 0.008, and 0.007).} them to [3.3$\micron$] and M respectively so that they can be compared to ground-based measurements of exoplanets. For the exoplanets, we use photometry of the four HR 8799 planets \citep{2008Sci...322.1348M,2010Natur.468.1080M,2011ApJ...739L..41G,2012ApJ...753...14S}, 2M1207 b \citep[][this work]{2004A&A...425L..29C}, $\beta$ Pic b \citep{2011A&A...528L..15B,2013A&A...555A.107B}, and GJ 504 b \citep{2013ApJ...774...11K}. The exoplanets show a distinct separation from the field brown dwarfs in the left and center color-magnitude diagrams ($\rm M_{H}$ vs. H-Ks and $\rm M_{L'}$ vs. [3.3$\micron$]-L'). Brown dwarfs cool through the M dwarf and L dwarf sequence, becoming fainter and redder in the $\rm M_{H}$ vs. H-Ks color-magnitude diagram (left panel of Figure \ref{cmd}). As the brown dwarfs cool below an absolute H magnitude of $\sim$14, the sequence moves sharply to bluer colors where the brown dwarfs have a spectral type of `T'. The L$\rightarrow$T transition is thought to reflect the change from cloudy to cloud-free atmospheres \citep[e.g.][]{2006ApJ...640.1063B,2008ApJ...689.1327S,2010ApJ...723L.117M}, as the clouds dissipate (rain) and/or sink below the photosphere \citep{2001ApJ...556..872A,2002ApJ...571L.151B,2003ApJ...585L.151T,2004AJ....127.3553K}. The planets in this range of absolute H magnitude appear to be an extension of the L dwarf sequence, implying that planets maintain cloudy photospheres at cooler effective temperatures than field brown dwarfs. The distinction is thought to be gravity dependent\footnote{When comparing planets and brown dwarfs of the same effective temperature, the planets are younger and lower mass than the brown dwarfs.} \citep{2006ApJ...640.1063B,2008ApJ...689.1327S,2006ApJ...651.1166M,2010ApJ...723..850B,2013arXiv1310.0457L}. In low-gravity objects, hydrostatic equilibrium implies that a given species' condensation temperature (where cloud decks form) will occur at a lower pressure (higher in the atmosphere) than in an equivalent temperature high-gravity object \citep{2012ApJ...754..135M}. Thus the cloudy to cloud-free transition for low-gravity exoplanets should occur at lower effective temperatures and fainter H-magnitudes than for more massive brown dwarfs. For the exoplanets plotted in Figure 4a, $\beta$ Pic b \citep[$\sim$1700 K;][]{2013A&A...555A.107B}, HR 8799 cde \citep[$\sim$1100 K;][]{2011ApJ...729..128C,2012ApJ...753...14S}, 2M1207 b \citep[$\sim$1000 K;][]{2011ApJ...732..107S,2011ApJ...735L..39B} and HR 8799 b \citep[$\sim$900 K;][]{2011ApJ...729..128C} are all consistent with cloudy photospheres. The recently discovered GJ 504 b \citep[$\sim$510;][]{2013ApJ...774...11K} is approximately the same H-Ks as the other planets while being much fainter overall. However its J-H colors \citep[see Figure 11 of][]{2013ApJ...774...11K} are consistent with cloud-free, late T-type brown dwarfs, and narrow-band photometry shows methane absorption \citep{2013arXiv1310.4183J}. At around the same temperature range where field brown dwarfs are transitioning from cloudy to cloud free, carbon is converted from CO to CH$_{4}$, producing strong methane absorption bands at 1.7 $\micron$ and 3.3$\micron$ \citep{1997ApJ...491..856B,2002ApJ...564..466G}. For temperatures cooler than $\sim$1200K, virtually all of the carbon monoxide is removed if the reaction is allowed to go to chemical equilibrium \citep{2002Icar..155..393L,2011ApJ...735L..39B}. However, CO absorption is observed in field T dwarfs \citep{1997ApJ...489L..87N,2009ApJ...695..844G}, implying that the reaction does not reach equilibrium faster than CO from hot brown dwarf interiors is mixed into their cooler photospheres \citep{2000ApJ...541..374S}. While nearly all field brown dwarfs cooler than $\sim$1200K show CH$_{4}$ absorption in the near-infrared, exoplanets with similar effective temperatures do not \citep[HR 8799 bcd and 2M1207 b;][]{2007ApJ...657.1064M,2008Sci...322.1348M,2010ApJ...723..850B,2011ApJ...733...65B,2013Sci...339.1398K,2013ApJ...768...24O}. In the mid-infrared, the CH$_{4}$ fundamental absorption band is at 3.3$\micron$ and a strong CO band is at 4.7$\micron$, while intermediate wavelengths ($\sim$4$\micron$) have relatively weak molecular opacity \citep{1997ApJ...491..856B}. The center panel of Figure \ref{cmd} ($\rm M_{L'}$ vs. [3.3$\micron$]-L'), which is sensitive to CH$_{4}$ absorption, shows the field brown dwarfs becoming redder at later spectral types, due to increased methane absorption. The planets (HR 8799 bcde and 2M1207 b) are systematically bluer than the field brown dwarfs, which could be caused by a lack of methane, which absorbs at 3.3$\micron$ and/or clouds, which smooth out spectral features (see Section \ref{SED modeling} for detailed modeling). The right panel of Figure \ref{cmd} ($\rm M_{M}$ vs. L'-M), which is sensitive to CO absorption, also shows the brown dwarf sequence becoming redder at later spectral types, although in this case, the effect is caused by decreasing CO absorption. The planets (HR 8799 bcd) are roughly consistent with the brown dwarf sequence (within their large error bars), indicating that they may have a similar non-equilibrium excess of CO. \section{Spectral Energy Distributions and Modeling\label{SED modeling}} The color-magnitude diagrams described in Section \ref{mircmd} suggest that exoplanets maintain cloudy photospheres at lower temperatures than field brown dwarfs, and that non-equilibrium CO$\leftrightarrow$CH$_{4}$ chemistry is at least partially responsible for their 3-5$\micron$ SEDs. While there has been relative success fitting the near-infrared (1-2.5$\micron$) SEDs of these objects with cloudy and/or non-equilibrium chemistry models \citep{2008Sci...322.1348M,2011ApJ...729..128C,2011ApJ...737...34M,2011ApJ...733...65B,2011ApJ...732..107S,2011ApJ...735L..39B,2013Sci...339.1398K}, there has been more difficulty when the mid-infrared (3-5$\micron$) SEDs are considered in parallel \citep{2010ApJ...716..417H,2012ApJ...753...14S,2012ApJ...754..135M,2013arXiv1307.1404L}. Here we attempt to explain the broad-wavelength SEDs of planets, incorporating our new photometry of HR 8799 c, HR 8799 d, and 2M1207 b. Sections 4.1-4.3 review models from previous work, in the context of our new data. Section 4.4-4.5 present new models. Section 4.6 summarizes our aggregate findings. The models broadly fall under three families, which, for simplicity, we denote by their group leaders: \textit{Burrows}, \textit{Barman}, and \textit{Marley}. A listing of all of the models and their basic properties is given in Table \ref{model listing}. The photometry plotted in figures is listed in Tables \ref{LBT/Magellan} (new data) and \ref{lit phot} (literature data). \subsection{Previous Work: Cloudy/Thick-Cloudy Atmospheres\label{thick-cloudy subsec}} At the time of their initial discovery it was noticed that the HR 8799 planets had colors that looked like L dwarfs (cloudy with no methane absorption), while their luminosities were more consistent with field T dwarfs \citep{2008Sci...322.1348M}. A similar phenomenon had been seen in the planetary-mass object, 2M1207 b. But as a single object, it could be explained as a ``normal'' brown dwarf seen through an edge-on disk \citep{2007ApJ...657.1064M}. Since the HR 8799 planets are also redder than the field L dwarfs (see Figure \ref{cmd}), \citet{2011ApJ...729..128C} and \citet{2011ApJ...737...34M} used thick clouds to explain their broad SEDs, and \citet{2011ApJ...732..107S} used the same models to explain 2M1207 b. The adopted models from \citet{2011ApJ...737...34M} and \citet{2011ApJ...732..107S} are shown in Figure \ref{thick clouds fig}. For HR 8799 c and d, the models under-predict the [3.3$\micron$] flux by a factor of $\sim$4 \citep{2012ApJ...753...14S}. Our new narrowband photometry of HR 8799 c and d (shown in the right panel) are consistent with the broad-band photometry (in the left panel) and show a shallow slope with a slight break between L$_{NB3}$ and L$_{NB4}$ (which is consistent with methane's opacity profile; see Figure \ref{filter profiles}), rather than the deep methane absorption feature seen in the thick-cloud models. Our new photometry of 2M1207 b shows that it is also much brighter at [3.3$\micron$] than the thick-cloud models predict (by a factor of$\sim$2, but this is very dependent on the precise choice of cloud thickness), and it has a similar, if not more extreme 3.3$\micron$-3.8$\micron$ slope as the HR 8799 planets. \subsection{Previous Work: Cloudy Atmospheres with Non-Equilibrium CO$\leftrightarrow$CH$_{4}$ Chemistry} After obtaining a near-infrared spectrum of HR 8799 b, \citet{2011ApJ...733...65B} showed that in addition to clouds, non-equilibrium CO$\leftrightarrow$CH$_{4}$ chemistry, caused by vertical mixing, is required to explain the spectrum of HR 8799 b. Subsequently, \citet{2011ApJ...735L..39B} and \citet{2011ApJ...739L..41G} applied similar models to 2M1207 b, and HR 8799 c and d respectively. \citet{2012ApJ...753...14S} obtained [3.3$\micron$] photometry of all four HR 8799 planets, which were much brighter than the thick-cloudy/equilibrium chemistry atmosphere models (described in Section \ref{thick-cloudy subsec}), and somewhat brighter than the non-equilibrium models from \citet{2011ApJ...733...65B} and \citet{2011ApJ...739L..41G}. To investigate the effects of non-equilibrium chemistry, \citet{2012ApJ...753...14S} used the thick cloud atmospheres from \citet{2011ApJ...737...34M} and removed CH$_{4}$ until the models were consistent with the [3.3$\micron$] photometry; however, this made the models too bright compared to existing L' (3.8$\micron$) photometry. The non-equilibrium models from \citet{2012ApJ...753...14S} are shown in Figure \ref{non-eq thick clouds fig}. These models show a sharp flux increase at the edge of the methane absorption feature, similar to what is seen in the equilibrium chemistry models, but shifted in wavelength. This is inconsistent with the smooth slope seen in our narrowband photometry. While the non-equilbrium chemistry models predict a fairly sharp increase at the edge of the methane absorption feature for HR 8799, the same is not true for 2M1207 b. In Figure \ref{non-eq thick clouds fig}, we plot the \citet{2011ApJ...735L..39B} model of 2M1207 b along with our new [3.3$\micron$] photometry. The photometry is consistent with the model. 2M1207 b is redder than the HR 8799 planets in the near-infrared. This pushes the models towards thicker clouds, which wash out the methane absorption feature at 3.3$\micron$ in concert with non-equilibrium chemistry. For HR 8799, the near-infrared driven cloud properties are not quite thick enough to wash out the 3-4$\micron$ SED, even with non-equilibrium chemistry. \subsection{Previous Work: ``Patchy'' Linear Combinations of Cloudy Models} Based on the earlier observations that 2M1207 b and the HR 8799 planets have L dwarf SEDs at T dwarf luminosities, \citet{2012ApJ...753...14S} attempted to fit the HR 8799 data with linear combinations of patchy cloud models, where the SED would be dominated by an L dwarf model, while a cooler, cloudier model suppressed its luminosity without drastically changing its shape. The adopted models from \citet{2012ApJ...753...14S} are shown in Figure \ref{linear comb clouds fig}. The combined models show an almost linear slope from 3-4$\micron$. Our narrowband photometry has the same general shape as the models, but with, perhaps, a slight jump in flux between $L_{NB3}$ and $L_{NB4}$ (particularly for HR 8799 d), which would be better captured by the non-equilibrium chemistry models. Note that the narrowband 3-4$\micron$ photometry are systematically higher than the models (which were chosen to fit the broad-band photometry), but are consistent within error bars. \subsection{Patchy Cloud Models (Cloudy/Cloudier and Cloudy/Cloud-Free)} While the patchy cloud models from \citet{2012ApJ...753...14S} are encouraging, they are constructed by linearly combining a pair of one-dimensional atmospheric models which have different temperature profiles, including in the deep atmosphere where global circulation and convection would presumably homogenize their thermal profiles. A self-consistent two column model needs to account for the energy flux carried by both components and should employ the same thermal profile. Such a model is presented in \citet{2010ApJ...723L.117M} which allows for two atmosphere columns with arbitrary spatial coverage that combine to carry a specified thermal flux given by $\sigma T_{\rm eff}^4$. Using this formalism we construct illustrative models showing how heterogeneous clouds affect a planet's SED. Figure \ref{patchy clouds fig} shows three models demonstrating the effect of self-consistently calculated cloud patchiness\footnote{The models are generally described in \citet{2008ApJ...689.1327S} and \citet{2012ApJ...754..135M}, with cloud model utilizing parameter $f_{\rm sed}$, from \citet{2001ApJ...556..872A} and \citet{2010ApJ...723L.117M}}: very thick (opaque) clouds ($f_{\rm sed}=0.25$), very thick clouds with patches of thinner clouds (same cloud model as the very thick clouds but with 10\% opacity), and very thick clouds with clear patches. The models with no clear patches both wash out the near-infrared water absorption features \citep[the absorption is seen in the spectra of ][]{2011ApJ...733...65B,2013Sci...339.1398K} and suppress too much of the planet's near-infrared flux, pushing it to the mid-infrared. Small clear patches (here 10\% of the planet surface area) allow flux to escape from the holes and bring the near-infrared flux up to the measured values, but the resulting cooler atmosphere results in a large methane absorption feature at [3.3$\micron$], which is not seen in our mid-infrared photometry. \subsection{Patchy Cloud Models with Non-Equilibrium Chemistry\label{patchy non-eq subsec}} In Figure \ref{patchy clouds fig} the $\sim$3.3$\micron$ absorption feature arises from relatively cool methane found very high in the atmosphere, at pressures of 10 to 20 mbar and temperatures of 875 to 900 K. Because this is close to the $\rm CH_4 / CO$ equilibrium conditions and the deeper atmosphere is well within the region of CO stability, moderate vertical mixing can mix CO into the upper atmosphere and substantially lower the methane mixing ratio. Figure \ref{non-eq patchy clouds fig} compares the cloudy/cloud-free model spectrum from Figure \ref{patchy clouds fig} with a spectrum computed following the prescription of \citet{2006ApJ...647..552S} for the same temperature and cloud profile but with eddy diffusivity $K_{zz} = 10^6\,\rm cm^2\, s^{-1}$. With such mixing the methane number mixing ratio drops from $\sim 10^{-4}$ to $\sim 10^{-7}$ and the absorption feature is absent. Such levels of mixing are consistent with those observed in some brown dwarfs \citep{2009ApJ...702..154S} and are broadly consistent with the strong gravity wave flux observed in numerical models of convection under these conditions \citep{2010A&A...513A..19F}. The model spectra provide a qualitatively good fit to both HR 8799 c and d, although we have not attempted to fine-tune the models to fit the curvature seen in the narrowband filter SED, or constrain the exact model parameters. \subsection{Summary of Modeling\label{model summary}} For the HR 8799 planets, we find that a combination of patchy clouds and non-equilibrium CO$\leftrightarrow$CH$_{4}$ chemistry is necessary to fit the observed data. Our 3-4$\micron$ photometry, in particular, is not fit well by models that employ patchy clouds without non-equilibrium chemistry, or non-equilibrium chemistry without patchy clouds. For 2M1207 b, non-equilibrium chemistry models, which employ thick, but not patchy clouds, provide an adequate fit, and there is no reason to add extra free parameters. This result suggests that all three planets have vertical mixing/non-equilibrium chemistry in their atmospheres, and all three planets have thick clouds, but only HR 8799 c and d necessarily have holes in their clouds. It is possible that the HR 8799 planets are in the process of entering the L$\rightarrow$T transition, where cloud patchiness is observed as periodic variability in some field brown dwarfs \citep[e.g.,][]{2009ApJ...701.1534A,2012ApJ...750..105R}, while 2M1207 b, which is younger, has not yet evolved to that point. As is the case for the brown dwarfs, the patchy clouds should cause variability in the HR 8799 planets. Vertical mixing as evidenced by non-equilibrium chemistry appears to be persisting in the HR 8799 planets, even as the clouds are starting to dissipate. For young, low-mass objects, the disappearance of clouds and the appearance of methane absorption might happen on different timescales and for different reasons. Our work demonstrates that there are reasonable models that can explain the broad-wavelength appearances of the HR 8799 planets and 2M1207 b. A dedicated modeling effort is needed to constrain the ranges of important physical parameters. \section{Summary and Conclusions} While the first generation of directly imaged planets were primarily discovered in the near-infrared (1-2.5$\micron$), cooler planets emit the majority of their light at longer wavelengths, where it is therefore critical to study their general appearances. In particular, the 3-5$\micron$ region contains a significant amount of information about gaseous and condensed absorbers, while also being accessible to current ground-based telescopes. In this work, we have obtained photometry of HR 8799 c and d, and 2M1207 b in the 3-4$\micron$ window, which contains the methane fundamental absorption feature, and is a region where previous work has struggled to explain planet SEDs with atmospheric models. The directly-imaged exoplanets are separated from the field brown dwarf population in $\rm M_{H}$ vs. H-K and $\rm M_{L''}$ vs. [3.3$\micron$]-L' color-magnitude diagrams, suggesting that the exoplanets are characteristically different than similar temperature, but older and more massive field brown dwarfs. Based on new and existing photometry, we attempt to model the complete SEDs of HR 8799 c and d, and 2M1207 b. While models with thick clouds and vertical-mixing/non-equilibrium chemistry can explain the appearance of 2M1207 b, an additional free parameter for cloud patchiness is necessary to fit HR 8799 c and d. Cloud patchiness is observed in the atmosphere of Jupiter and may signify the onset of the L$\rightarrow$T transition, where clouds dissipate, and/or sink below the photosphere of a planet/brown dwarf. In this scenario, the HR 8799 planets might be beginning their L$\rightarrow$T transition, while 2M1207 b is still too young. Patchy clouds manifest as periodic variability in some brown dwarfs. Our modeling predicts that HR 8799 c and d should be variable, although the magnitude of this result depends on the scale of the clouds holes and the orientation of the planets with respect to our line of sight \citep{2013ApJ...762...47K}. Based on Figure 8, the variability should be evident at all wavelengths, with peak amplitudes in the J-band, and the short-wavelength side of the L-band. The fact that the HR 8799 planets and 2M1207 b have different colors than field brown dwarfs suggests that it will be difficult to categorize brown dwarfs and extrasolar planets with sparse collections of infrared photometry. Late-type brown dwarfs found by WISE, and ultra-cool exoplanets found by JWST imaging will both need to be studied in detail with multi-wavelength photometry and spectroscopy. Additionally, parallax measurements will be essential for breaking the degeneracy between nearby, young objects, and distant, older objects that might have similar colors. Given that the planets studied in this work all have very shallow 3.3$\micron$-3.8$\micron$ slopes, it is worth considering whether L-band exoplanet imaging surveys should use a broader bandpass than the typically used L' filter (3.4-4.1$\micron$). For HR 8799-like planets, a broader filter (3.0-4.1$\micron$) would improve sensitivity. However, such a filter might not be a good match for cooler planets, if methane absorption becomes prominent at lower temperatures. Currently, only a handful of exoplanets have been directly imaged. As their numbers grow, it will be possible to determine how youth, gravity, composition and formation history affect their appearances. And as the population of directly-imaged planets extends to cooler temperatures and lower masses, the mid-infrared will be a critical wavelength range for understanding their aggregate properties. \acknowledgements The authors thank Travis Barman for his insightful comments and for supplying his 2M1207 b model. We also thank the anonymous referee for his/her excellent suggestions. This work would not have been possible without the dedication of the LBTI staff, in particular Vidhya Vaitheeswaran, who programmed LBTI's rapid filter changing capabilities. AS was supported by the NASA Origins of Solar Systems Program, grant NNX13AJ17G. VB is supported by the NSF Graduate Research Fellowship Program (DGE-1143953). The Large Binocular Telescope Interferometer is funded by the National Aeronautics and Space Administration as part of its Exoplanet Exploration program. LMIRCam is funded by the National Science Foundation through grant NSF AST-0705296. \clearpage \begin{deluxetable}{lccccccccccc} \tablecolumns{8} \tabletypesize{\scriptsize} \tablecaption{LBTI Photometry of HR 8799 c and d and Clio Photometry of 2M1207 b} \tablewidth{0pt} \tablehead{ \colhead{} & \colhead{L$_{NB1}$} & \colhead{L$_{NB2}$} & \colhead{L$_{NB3}$} & \colhead{L$_{NB4}$} & \colhead{L$_{NB5}$} & \colhead{L$_{NB6}$} & \colhead{\[[3.3$\micron$]} \\ \colhead{} & \colhead{(3.04$\micron$)} & \colhead{(3.16$\micron$)} & \colhead{(3.31$\micron$)} & \colhead{(3.46$\micron$)} & \colhead{(3.59$\micron$)} & \colhead{(3.78$\micron$)} & \colhead{(3.31$\micron$)} } \startdata \sidehead{\textit{Contrast with Respect to Host Star}} HR 8799 c & 10.12$\pm$0.15 & 10.17$\pm$0.20 & 9.97$\pm$0.15 & 9.52$\pm$0.15 & 9.32$\pm$0.15 & 9.32$\pm$0.15 & \\ HR 8799 d & 10.12$\pm$0.20 & 10.02$\pm$0.25 & 10.12$\pm$0.20 & 9.52$\pm$0.15 & 9.27$\pm$0.15 & 9.27$\pm$0.15 & \\ 2M1207 b & & & & & & & 3.86$\pm$0.10 \\ \sidehead{\textit{Absolute Magnitudes}} HR 8799 c & 12.35$\pm$0.15 & 12.40$\pm$0.20 & 12.20$\pm$0.15 & 11.75$\pm$0.15 & 11.55$\pm$0.15 & 11.55$\pm$0.15 & \\ HR 8799 d & 12.35$\pm$0.20 & 12.25$\pm$0.25 & 12.35$\pm$0.20 & 11.75$\pm$0.15 & 11.50$\pm$0.15 & 11.50$\pm$0.15 & \\ 2M1207 b & & & & & & & 11.85$\pm$0.14 \\ \enddata \tablecomments{The listed absolute magnitudes do not include a distance modulus uncertainty of 0.06 mag \citep{2007A&A...474..653V}, which is correlated between all bands.} \label{LBT/Magellan} \end{deluxetable} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f1.ps} \vspace{0.1in} \caption{Characteristic examples of exoplanet-to-star contrasts (i.e. flux ratios) as a function of wavelength, showing (1) that gas-giant exoplanets can be detected with lower contrasts in the mid-infrared (3-5$\micron$) than in the near-infrared (1-2$\micron$), and (2) that this difference increases at lower temperatures. While the planets that have been directly imaged to date ($\beta$ Pic b and HR 8799 c, d and e on this plot) are relatively warm (1600 K and 1000 K, respectively), it is likely that the majority of self-luminous exoplanets are much cooler. Planets that formed by core-accretion \citep[approximated by the ``cold-start'' models;][]{2008ApJ...683.1104F} are never hotter than $\sim$700 K. Planets around average-aged stars (5 Gyr) are never hotter than $\sim$400 K, regardless of formation history. Jupiter, which may be a ubiquitous outcome of planet formation, is only $\sim$130 K. \citep[Models from ][]{2011ApJ...737...34M,2003ApJ...596..587B}. \label{mid-IR examples}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=0,width=\columnwidth]{f2.ps} \caption{LBTI/LMIRCam images of the HR 8799 planetary system in 6 narrowband filters. Each wavelength was observed for 7 minutes with the incoherently combined 2$\times$8.4 meter LBT aperture. HR 8799 c and d are visible in all 6 filters. HR 8799 b is visible in 4 of the 6 filters, although it is low S/N, given the short integration times. HR 8799 e is also detectable at low S/N in the longest wavelength filters, given its known position. The planet positions are circled in the lower-right panel. North is up and East is left. The images are each 3.5'' across. \label{HR8799_image}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=0,width=0.5\columnwidth]{f3.ps} \caption{MagAO/Clio2 images of the 2M1207 system. The binary is clearly resolved despite being 38'' off-axis from its faint (R$\sim$14.2) AO guide-star. The slight elongation of 2M1207 A is from the anisoplanatism of the off-axis adaptive optics reference star. 2M1207 b appears to be circular because of the image stretch, which just shows its diffraction-limited core (FWHM=0.1''). North is up and East is left. The field-of-view shown here is 2 by 2.5 arcseconds. \label{2M1207_image}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f4.ps} \caption{Color-magnitude diagrams showing the positions of directly imaged extrasolar planets with respect to field brown dwarfs. In the left-hand panel, the HR 8799 planets and 2M1207 b appear to be a continuation of the L dwarf sequence, implying that the planets have clouds at fainter absolute H-band magnitudes than the field brown dwarfs. The recently discovered GJ 504 b also appears below the L dwarf sequence, although other near-infrared colors place it with the field T dwarfs. In the center panel, the directly imaged planets are systematically bluer than the field brown dwarfs, suggesting that their 3.3$\micron$ methane absorption feature is somewhat diminished through some combination of clouds and non-equilibrium CO$\leftrightarrow$CH$_{4}$ chemistry. The right-hand panel shows the directly-imaged planets as relatively consistent with the field brown dwarfs, although with large error bars. For all three plots, error bars in the upper-right corner represent the maximum error bar for the brown dwarf sample. \label{cmd}} \end{center} \end{figure} \clearpage \begin{deluxetable}{lcccccccccccc} \tabletypesize{\scriptsize} \tablecaption{Literature Photometry (Absolute Magnitudes)} \tablewidth{0pt} \tablehead{ \colhead{Filter} & \colhead{$\lambda_{eff}$} & \colhead{HR 8799 c} & \colhead{HR 8799 d} & \colhead{2M1207 b} & \colhead{References} } \startdata F090M & 0.90$\micron$ & & & 18.83$\pm$0.25 & 1 \\ F110M & 1.10$\micron$ & & & 16.98$\pm$0.15 & 1 \\ J & 1.25$\micron$ & 14.65$\pm$0.17 & 15.26$\pm$0.43 & 16.37$\pm$0.20 & 2,2,3 \\ F145M & 1.45$\micron$ & & & 15.42$\pm$0.03 & 1 \\ F160W & 1.60$\micron$ & & & 14.63$\pm$0.02 & 1 \\ H & 1.63$\micron$ & 14.18$\pm$0.14 & 14.23$\pm$0.2 & 14.46$\pm$0.21 & 4,4,5 \\ Ks & 2.15$\micron$ & 13.13$\pm$0.08 & 13.11$\pm$0.12 & 13.30$\pm$0.11 & 2,2,5 \\ 3.3 & 3.3$\micron$ & 12.2$\pm$0.11 & 12.0$\pm$0.11 & & 4,4 \\ L' & 3.8$\micron$ & 11.74$\pm$0.09 & 11.56$\pm$0.16 & 11.65$\pm$0.14 & 2,2,5 \\ M & 4.7$\micron$ & 12.05$\pm$0.14 & 11.67$\pm$0.35 & & 6,6 \\ \enddata \tablerefs{ (1) \citet{2006ApJ...652..724S} (2) \citet{2008Sci...322.1348M} (3) \citet{2007ApJ...657.1064M} (4) \citet{2012ApJ...753...14S} (5) \citet{2004A&A...425L..29C} (6) \citet{2011ApJ...739L..41G} } \label{lit phot} \end{deluxetable} \clearpage \begin{deluxetable}{lcccccccccccc} \rotate \tabletypesize{\scriptsize} \tablecaption{Atmospheric Models Used} \tablewidth{0pt} \tablehead{ \colhead{Figure} & \colhead{Object} & \colhead{Model Family} & \colhead{T$_{eff}$} & \colhead{Clouds} & \colhead{Chemistry} & \colhead{Reference} } \startdata \ref{thick clouds fig} & HR 8799 c & Burrows & 1000 K & AE-type & Equilibrium & \citet{2011ApJ...737...34M} \\ \ref{thick clouds fig} & HR 8799 d & Burrows & 900 K & AE-type & Equilibrium & \citet{2011ApJ...737...34M} \\ \ref{thick clouds fig} & 2M1207 b & Burrows & 1000 K & A-type & Equilibrium & \citet{2011ApJ...732..107S} \\ \hline \ref{non-eq thick clouds fig} & HR 8799 c & Burrows & 1000 K & AE-type & Equilibrium & \citet{2011ApJ...737...34M} \\ \ref{non-eq thick clouds fig} & HR 8799 c & Burrows & 1000 K & AE-type & 0.1$\times$CH$_{4}$, 10$\times$CO & \citet{2012ApJ...753...14S} \\ \ref{non-eq thick clouds fig} & HR 8799 c & Burrows & 1000 K & AE-type & 0.01$\times$CH$_{4}$, 100$\times$CO & \citet{2012ApJ...753...14S} \\ \ref{non-eq thick clouds fig} & HR 8799 d & Burrows & 900 K & AE-type & Equilibrium & \citet{2011ApJ...737...34M} \\ \ref{non-eq thick clouds fig} & HR 8799 d & Burrows & 900 K & AE-type & 0.1$\times$CH$_{4}$, 10$\times$CO & \citet{2012ApJ...753...14S} \\ \ref{non-eq thick clouds fig} & HR 8799 d & Burrows & 900 K & AE-type & 0.01$\times$CH$_{4}$, 100$\times$CO & \citet{2012ApJ...753...14S} \\ \ref{non-eq thick clouds fig} & 2M1207 b & Burrows & 1000 K & A-type & Equilibrium & \citet{2011ApJ...732..107S} \\ \ref{non-eq thick clouds fig} & 2M1207 b & Burrows & 1000 K & A-type & Equilibrium & \citet{2011ApJ...732..107S} \\ \textbf{\ref{non-eq thick clouds fig}} & \textbf{2M1207 b} & \textbf{Barman} & \textbf{1000 K} & \textbf{best-fit thickness} & \textbf{Kzz}$\bm{=10^{8}}$ & \textbf{\citet{2011ApJ...735L..39B}} \\ \hline \ref{linear comb clouds fig} & HR 8799 c & Burrows & 700K/1400K & A-type/AE-type & Equilibrium & \citet{2012ApJ...753...14S} \\ \ref{linear comb clouds fig} & HR 8799 d & Burrows & 700K/1400K & A-type/AE-type & Equilibrium & \citet{2012ApJ...753...14S} \\ \hline \ref{patchy clouds fig} & HR 8799 c & Marley & 1100 K & $f_{\rm sed}$=0.25 & Equilibrium & this work \\ \ref{patchy clouds fig} & HR 8799 c & Marley & 1100 K & $f_{\rm sed}$=0.25 with 30\% thin clouds & Equilibrium & this work \\ \ref{patchy clouds fig} & HR 8799 c & Marley & 1100 K & $f_{\rm sed}$=0.25 with 10\% cloud free & Equilibrium & this work \\ \ref{patchy clouds fig} & HR 8799 d & Marley & 1100 K & $f_{\rm sed}$=0.25 & Equilibrium & this work \\ \ref{patchy clouds fig} & HR 8799 d & Marley & 1100 K & $f_{\rm sed}$=0.25 with 30\% thin clouds & Equilibrium & this work \\ \ref{patchy clouds fig} & HR 8799 d & Marley & 1100 K & $f_{\rm sed}$=0.25 with 10\% cloud free & Equilibrium & this work \\ \hline \ref{non-eq patchy clouds fig} & HR 8799 c & Marley & 1100 K & $f_{\rm sed}$=0.25 with 10\% cloud free & Equilibrium & this work \\ \textbf{\ref{non-eq patchy clouds fig}} & \textbf{HR 8799 c} & \textbf{Marley} & \textbf{1100 K} & \textbf{$\bm{f_{\rm sed}}$=0.25 with 10\% cloud free} & \textbf{Kzz}$\bm{=10^{6}}$ & \textbf{this work} \\ \ref{non-eq patchy clouds fig} & HR 8799 d & Marley & 1100 K & $f_{\rm sed}$=0.25 with 10\% cloud free & Equilibrium & this work \\ \textbf{\ref{non-eq patchy clouds fig}} & \textbf{HR 8799 d} & \textbf{Marley} & \textbf{1100 K} & \textbf{$\bm{f_{\rm sed}}$=0.25 with 10\% cloud free} & \textbf{Kzz}$\bm{=10^{6}}$ & \textbf{this work} \\ \enddata \tablecomments{The \textit{Burrows} models are generally described in \citet{1997ApJ...491..856B} and \citet{2006ApJ...640.1063B}, with cloud parameterizations from \citet{2011ApJ...737...34M}. The \textit{Barman} models are comprehensively described in \citet{2011ApJ...733...65B} and references therein. The \textit{Marley} models are generally described in \citet{2008ApJ...689.1327S} and \citet{2012ApJ...754..135M}, with cloud parameterizations from \citet{2001ApJ...556..872A} and \citet{2010ApJ...723L.117M}. Adopted models are highlighed in bold.} \label{model listing} \end{deluxetable} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f5.ps} \vspace{0.1in} \caption[]{Absolute photometry for HR 8799 cd and 2M1207 b compared to thick-cloud atmosphere models. The photometric data are shown as red error bars, the model spectra are shown as blue curves, and the model photometric predictions are shown as blue horizontal lines spanning the filter half-maximums. The thick-cloud models are adopted from \citet{2011ApJ...737...34M} for HR 8799 cd and \citet{2011ApJ...732..107S} for 2M1207 b, using the grid from \citet{2011ApJ...737...34M}. Two cloud parameterizations are used: A-type, which are the thickest clouds in \citet{2011ApJ...737...34M} and AE-type, which are slightly thinner.\\\\ For all three objects, the broad-band photometry (left-side of the figure) is generally well-fit by the models, except for the [3.3$\micron]$ filter. Our new narrow-band 3-4$\micron$ photometry (right-side of the figure) is consistent with broader-band photometry at similar wavelengths, although systematically brighter within errors. \label{thick clouds fig}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f6.ps} \vspace{0.1in} \caption[]{Same as Figure \ref{thick clouds fig} but with models that include non-equilibrium CO$\leftrightarrow$CH$_{4}$ chemistry or parameterized CO/CH$_{4}$ opacity. The \textit{Burrows} models are from \citet{2012ApJ...753...14S} and have similar clouds properties as Figure \ref{thick clouds fig} but with diminished CH$_{4}$ and enhanced CO opacity. The \textit{Barman} model, which is from \citet{2011ApJ...735L..39B} also employs thick clouds, and uses chemical reaction rates to self-consistently calculate CH$_{4}$ and CO mixing ratios assuming a vertical diffusion coefficient of $K_{zz}=10^{8}cm^{2}/s$.\\\\ The HR 8799 planets are not well-fit by the \textit{Burrows} models, which predict a sharp edge to the methane absorption feature, even with suppressed methane opacity. Our new narrow-band 3-4$\micron$ photometry suggest a smoother slope. Note that analogous models from \citet{2011ApJ...733...65B} and \citet{2011ApJ...739L..41G} predict qualitatively similar behavior (when considering models that have radii consistent with evolutionary tracks). However, the \textit{Barman} model fits 2M1207 b quite well, including our new [3.3$\micron$] photometry. Combined with \citet{2011ApJ...735L..39B}'s fit of the 2M1207 b near-infrared spectrum, it appears that thick clouds and non-equilibrium chemistry can explain all of the existing data for 2M1207 b. \label{non-eq thick clouds fig}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f7.ps} \vspace{-1.3in} \caption[]{Same as Figure \ref{thick clouds fig} but with models that linearly combine two different thick-cloud atmosphere models to represent non-isotropic emission through regions with different temperatures and different cloud properties. The models are from \citet{2012ApJ...753...14S} using the grid from \citet{2011ApJ...737...34M}.\\\\ The HR 8799 planets are reasonably well-fit by the data. However, the modeling approach of linearly combining two atmosphere models is not entirely self-consistent, so additional modeling (Section 4.4) is needed to study patchy cloud models. \label{linear comb clouds fig}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f8.ps} \vspace{-1.3in} \caption[]{Same as Figure \ref{thick clouds fig} but with self-consistent patchy clouds calculated using the formalism of \citet{2010ApJ...723L.117M}. The thick cloud models use $f_{\rm sed}$=0.25 \citep{2001ApJ...556..872A}, which are very thick clouds. The models with patches of thin clouds also use $f_{\rm sed}$=0.25, but with 30\% of the surface having 10\% of the normal cloud opacity. In the third set of models, 10\% of the surface has no cloud opacity.\\\\ The HR 8799 planets are not well-fit by any of the models. The thick cloud models and models with thin cloud patches both wash out the near-infrared SED where spectra of HR 8799 b, c, and 2M1207 b show strong water absorption bands \citep{2011ApJ...733...65B,2013Sci...339.1398K,2010A&A...517A..76P}. The models with cloud-free holes still have too much methane absorption from $\sim$3.1-3.6$\micron$. \label{patchy clouds fig}} \end{center} \end{figure} \clearpage \begin{figure} \begin{center} \includegraphics[angle=90,width=\columnwidth]{f9.ps} \vspace{-1.3in} \caption[]{Same as Figure \ref{thick clouds fig} but with self-consistent patchy clouds calculated using the formalism of \citet{2010ApJ...723L.117M} combined with non-equilibrium chemistry as described in \citet{2006ApJ...647..552S}. The non-equilibrium models use a diffusion coefficient of $K_{zz}=10^{6}cm^{2}/s$.\\\\ Both of the HR 8799 planets are reasonably well-fit by the patchy-cloud/non-equilibrium chemistry model, although the model clearly over-predicts the M-band (4.7$\micron$) flux of HR 8799 c. These models are a heuristic demonstration of the type of processes that are likely shaping the spectra of these directly imaged planets, although clearly more detailed model fits are required to find the best fitting parameters. \label{non-eq patchy clouds fig}} \end{center} \end{figure} \clearpage
\section{Introduction} The optical turbulence vertical profilers (OTVP) are widely used in ground-based astronomy to estimate the stratification of the optical turbulence ($\mbox{$C_N^2$}$ profiles) from the ground up to the top of the atmosphere ($\sim$ 20~km) or for sub-ranges of this interval \citep{roddier1981,rocca1974,vernin1983,avila1997,klueckers1998,fuensalida2004,egner2007,kornilov2003,wilson2002,azouit2005,avila2009,egner_masciadri2007,osborn2010,beckers2001,hickson2004,tokovinin2010}. Optical turbulence (OT) degrades the quality of images obtained with ground-based telescopes by introducing perturbations on the perfect wavefronts coming from the observed objects located outside the atmosphere at infinite distance. To correct the wavefront perturbations, recover the intrinsic angular resolution of telescopes and obtain images at diffraction limit, different typologies of adaptive optics (AO) techniques have been developed in the last decades \citep{beckers1993,davies2012}. However, all AO techniques depend on the OT status in the whole atmosphere. The knowledge of the spatio-temporal distribution of the OT in an astronomical site is therefore fundamental for many reasons: to optimize the AO techniques efficiency; to search for sites with favorable turbulence characteristics; to characterize the atmosphere of the existent observatories to prove the reliability of new AO techniques such as the wide-field-of-view AO \citep{neichel2009,vidal2010,basden2013}; to forecast the OT to schedule scientific programs to be carried out and instruments to be placed at the focus of telescopes. Different OTVPs have been developed in the last decades, each one characterized by some specific features. The Generalized SCIDAR (hereafter GS: \citet{fuchs1994,fuchs1998}) is a well known technique based on the autocorrelation of scintillation maps produced by binary stars on the pupil of a telescope. This technique permits to reconstruct the turbulence stratification from the ground up to 20-25~km with a variable vertical resolution that scales as \citep{vernin1983}: \begin{equation} \Delta H(h) = \frac{{0.78 \cdot \sqrt {\lambda | h - h_{gs}|} }} {\theta } \label{res} \end{equation} \noindent where $\lambda$ is the wavelength, $h$ is the height from the ground, $h$$_{gs}$ is the conjugated height under-ground (see Section \ref{gs}) and $\theta$ is the binary separation. Considering the typical values of the binaries separation and $h$$_{gs}$ used for the GS, we can say that the GS vertical resolution is typically of the order of 1~km on the whole atmosphere. The GS is based on a solid and simple optical principle so that several GSs have been built by different teams worldwide in the last two decades and used for site characterization on many among the major astronomical Observatories in the world \citep{avila1997,avila2004,fuensalida2004,klueckers1998,mckenna2003,garcia2006,egner2007,egner_masciadri2007,fuensalida2008,masciadri2010,garcia2011a,garcia2011b,avila2011,masciadri2012}. The GS is, at present, one of the most reliable OTVP to be used as a reference for testing new OTVPs on the whole 20~km. The main limitation is that it requires a telescope of at least 1~m (preferably 1.5~m) pupil size and the brightness of the binary stars has to be $\le$ 5-6 mag. This reduces the number of usable stars on sky. It is therefore not a good candidate to be used as an automatic monitor to be placed in astronomical observatories for systematic optical turbulence monitoring and/or for site searches but it is more suitable for dedicated experiments. The Multi Aperture Scintillation Monitor (MASS) entered in the astronomic context in more recent years \citep{kornilov2003}. It is based on a principle proposed by Ochs in 1976 \citep{ochs1976} that implies the observation of scintillation maps produced by single stars on the telescope's pupil. The turbulence profile is reconstructed thanks to the analysis of the spatial structure of the scintillation map through a set of different spatial filters and the use of a set of weighting functions. Kornilov introduced, as spatial filters, a set of four small concentric and circular apertures selected on the telescope's pupil \citep{kornilov2003}. The main attractive features of the MASS are: (a) a better sky coverage than the GS (single stars are much more numerous than binary stars with magnitude $\le$ 5-6~mag) and (b) a small pupil of the telescope (of the order of 8-15~cm) well suitable therefore to be employed as an automatic monitor. The main drawbacks are: (a) a lower vertical resolution that scales as $\Delta h$ $\sim$ 0.5$\cdot$h \citep{tokovinin2003a}. More precisely it provides OT profiles sampled on 6 layers centered at 0.5~km, 1~km, 2km, 4~km, 8~km and 16km a.g.l.; (b) the MASS is insensitive to the optical turbulence near the ground (below $\sim$ 0.5~km). It is therefore commonly considered as a free-atmosphere turbulence monitor. Turbulence ground layer contribution has been frequently estimated by subtracting the MASS estimate from the Differential Image Motion Monitor (DIMM)\footnote{The latter provides estimates of the integrated turbulence on the whole atmosphere (along the optical path).}; (c) because of the particular weighting functions (triangle shape) of the MASS, this method does not permit to identify precisely the height of the boundary layer and in general the height separating a layer from the contiguous one. Turbulence developed inside a vertical slab [h$_{1}$,h$_{2}$]~km is distributed indeed in two different contiguous weighting functions. Even if this is not particularly critical for the site selection it can create problems in some other contexts and applications such as the AO. MASS and DIMM have been widely used in the last decade to characterize the astroclimatic parameters of many astronomical sites in the world \citep{tokovinin2003b,lawrence2004,kornilov2010,lombardi2010,dali2010,thomas-osip2012}) but it has been developed by just one scientific group. An intense use of this instrument has been done also for new generation telescopes site searches such as those aiming at the identification of the Thirty Meters Telescope (TMT) \citep{schoeck2009} and the European Extremely Large Telescope (E-ELT) \citep{vazquez-ramio2012} sites. Curiously, little has been done so far in terms of analysis of the MASS reliability with other OTVP. The unique study has been carried out so far by Tokovinin \citep{tokovinin2005} in which MASS and GS measurements taken simultaneously at Mauna Kea have been statistically compared on a sample of four nights\footnote{Most of the analysis on reliability has been done so far with simulations and post verification of the exact reconstruction of the vertical stratification \citep{tokovinin2003a}. }. The conclusions of that paper state that a very good agreement (better than 20$\%$) of the integral of the turbulence in the free atmosphere as measured by the two instruments is observed. The authors report that a very satisfactory agreement is reached for layers at 8 and 16~km while the MASS systematically overestimates the OT in the layer at 0.5~km and underestimates the OT in layers at 2 and 4~km. These errors have been ascribed to the restoration technique. In other words, the total integral of the six layers is well reconstructed by the MASS but its distribution among the six predefined layers presents some systematic biases. Because measurements from a more recent and statistical rich site testing campaign done at Cerro Paranal (and called PAR2007) are now available \citep{dali2010}, we have performed in this paper a detailed cross-comparison of MASS and GS measurements with the aim to provide a quantitative estimation of the accuracy of the turbulence stratification and an as precise as possible analysis of the reliability of the two instruments. This observed OT stratification represents, indeed, for us a reference to calibrate a mesoscale atmospherical model to be used to predict the OT stratification \citep{masciadri2013a}. It is therefore for us extremely important to quantify the absolute values of these quantities. In Section \ref{obs} we present the PAR2007 site testing campaign instruments and measurements that we treated. In Section \ref{instr} we briefly remind the main physical principles which the two techniques of the GS and MASS are based on with some basic information on the DIMM too. In Section \ref{analy} we present the comparison of the different astroclimatic parameters (seeing in the free atmosphere, seeing in each individual layer, isoplanatic angle and wavefront coherence time) provided by the two instruments with the associated discussion. In Section \ref{disc} we comment our results with respect to the literature. In Section \ref{concl} we present the conclusions of our study. \section{Observations} \label{obs} Measurements we treated in this paper belong to the PAR2007 site testing campaign that took place at Cerro Paranal in November-December 2007 \citep{dali2010} and was promoted by ESO in the context of the E-ELT Design Study. A set of instruments, among those a MASS, a GS and a DIMM have been run during 20 nights in this couple of months. Simultaneous measurements of GS and DIMM on 20 nights and simultaneous measurements of GS and MASS on a sub-sample of 14 nights are available (Table \ref{tab:nights}). In 2 nights over 14 there are just a few simultaneous measurements. We considered therefore a sub-sample of 12 nights. Figure \ref{map} shows the position of the three instruments on the summit of Cerro Paranal. The GS is placed at the focus of an Auxiliary Telescope (AT). The GS and DIMM are located at 5 and 6 meters above the ground respectively. The distance between the two instruments is roughly 205~m over a basically flat surface. \begin{figure} \centering \includegraphics[width=5.5cm]{fig1_masciadri} \caption{Satellite image of the Paranal Observatory from Google Earth. Yellow arrows indicate the position of DIMM, MASS and the Generalized SCIDAR. The distance between the DIMM and the Cute-SCIDAR is 205~m. The distance between the MASS and the DIMM is 20~m. The telescope UT4 is located at mid-path between the DIMM and the Cute-SCIDAR. The Cute-SCIDAR is a Generalized SCIDAR. \label{map}} \end{figure} \begin{table} \caption{List of nights in which simultaneous DIMM and GS measurements are available in the PAR2007 site testing campaign (20 nights). The asterisk indicates the sub-sample of nights in which MASS and GS simultaneous measurements are available (14 nights). The double asterisk indicates the nights in which just a few (respectively 1 and 2) seeing simultaneous measurements are available. These two nights are not used in the statistic analysis.\label{tab:nights}} \begin{center} \begin{tabular}{c c c} \hline \multicolumn{3}{c}{Observation nights of PAR2007 Site Testing Campaign} \\ \multicolumn{3}{c}{with DIMM, GS and MASS} \\ \hline 2007/11/10 & 2007/11/21$^{*}$ & 2007/12/19$^{*}$ \\ 2007/11/11 & 2007/11/22$^{*}$ & 2007/12/20$^{*}$ \\ 2007/11/13 & 2007/11/24 & 2007/12/21$^{*}$\\ 2007/11/14 & 2007/12/15$^{**}$ & 2007/12/22$^{*}$\\ 2007/11/17 & 2007/12/16$^{**}$ & 2007/12/23$^{*}$ \\ 2007/11/18$^{*}$ & 2007/12/17$^{*}$ & 2007/12/24$^{*}$\\ 2007/11/20$^{*}$ & 2007/12/18$^{*}$ & \\ \hline \end{tabular} \end{center} \end{table} \section{Instruments} \label{instr} \subsection{Generalized SCIDAR: Cute-SCIDAR} \label{gs} The SCIDAR technique (Scintillation Detection and Ranging) has been originally proposed by \citet{rocca1974} and \citet{vernin1983} to measure the vertical optical turbulence distribution (the refractive index structure constant $C_{N}^{2}$ profiles) in the troposphere. The technique relies on the analysis of the scintillation maps generated by binary stars on the pupil plane of a telescope. The standard SCIDAR technique (called Classic Scidar) is insensitive to the turbulence near the ground. This fact represented in the past an important limitation for monitoring turbulence for astronomical applications because it is known that most of the turbulence develops in the low part of the atmosphere. To overcome this limitation \citet{fuchs1994} and \citet{fuchs1998} proposed a generalized version of the SCIDAR (called Generalized SCIDAR) in which the detector is virtually conjugated below the ground at a distance $h_{gs}$ permitting to extend the measurement range to the whole atmosphere (from the ground up to $\sim$ 20-25 km). The GS is based on the observation of binaries having an angular separation typically of $\sim$ 3-10 arcsec, magnitude $\le$ 5-6 mag and a $\Delta$$m$ $\sim$ 1 mag. When two plane wavefronts coming from a binary and propagating through the atmosphere meet a turbulent layer located at a height $h$ from the ground, they produce, on the detector plan, optically placed below the ground at a few kilometers ($h_{gs}$), two scintillation maps characterized by typical shadows appearing in couple at a distance $d$. Such a distance is geometrically related to the position of the turbulent layer as $d$ = $\theta$($h$+$h_{gs}$). The calculation of the auto-covariance (AC) of the scintillation map, normalized by the autocorrelation of the mean image, produces the so called 'triplet' i.e. three peaks. The central peak of the triplet is located at the centre of the AC frame; the lateral peaks are symmetrically located at a distance $d$ from the centre. The amplitude of the lateral peaks is proportional to the auto-covariance of the scintillation map therefore to the strength of the turbulent layer. In summary the amplitude of the later peaks of the triplet and their distance from the central peak give us the information on the strength and the height h of the turbulent layer respectively. In a multi-layers atmosphere, different turbulent layers located at different height $h_{i}$ produce several triplets all centred in the centre of the AC frame but with lateral peaks located at different distances $d_{i}$ from the centre of the AC frame. The $\mbox{$C_N^2$}$ profiles are obtained taking into account the information provided by every turbulent layer all together. From an analytical point of view, as it is indicated by \citet{avila1997}-Eq.1, to obtain the $\mbox{$C_N^2$}$ profiles one has first to calculate the difference of the perpendicular and parallel sections of the normalized auto-covariance (AC), and then inverting this equation (known as Fredholm equation). The GS used in the PAR2007 is called CUTE-Scidar III (see Fig.\ref{map}). It has been developed by the Instituto de Astrof\'isica de Canarias (IAC) team (V\'azquez-Rami\'o et al., 2008). CUTE-Scidar III instrument provides $\mbox{$C_N^2$}$ profiles in real time without dome and mirror seeing components and it has been used since 2002 to extensively monitor the optical turbulence at El Roque de Los Muchachos and Teide observatories in Canary Islands \citep{fuensalida2004,garcia2011a,garcia2011b}. The dome and mirror seeing is quantified using a technique based on evenness properties with the Fourier analysis \citep{fuensalida2008}. The GS principle, indeed, permits to measure $\mbox{$C_N^2$}$ profiles including the turbulence contribution coming from the dome. A dedicated procedure is necessary to disentangle that from the turbulence provided by the atmosphere. The temporal sampling of the raw measurements is $\sim$ 1 minute. $\mbox{$C_N^2$}$ profiles are sampled along the z-axis at 300~m but this spatial scale is totally arbitrary. The intrinsic $\mbox{$C_N^2$}$ profiles spatial resolution is given by Eq.\ref{res}. Measurements of the PAR2007 campaign done with the CUTE-Scidar III have been corrected \citep{masciadri2012} to eliminate the error introduced by the normalization of the autocorrelation of the scintillation maps by the autocorrelation of the mean pupil \citep{johnston2002,avila2009}. This error, if it is not corrected, induces an overestimation of the $\mbox{$C_N^2$}$(h) that depends on many parameters related to the optical set-up and the observed binaries. More precisely, it depends on the diameter of the pupil of the telescope $D$, the height h$_{gs}$ at which the detection plane is conjugated below the ground, the difference $\Delta$m between the stellar magnitudes of the binaries, the angular separation of the binary $\theta$ and the ratio $e$ between the central obscuration D$^{*}$ and the telescope pupil (e = D$^{*}$/D). \subsection{MASS} \label{mass} \begin{table} \caption{Configuration of the MASS weighting functions extremes: triangle's bases and peaks. Units in meters (m).\label{table1}} \begin{center} \begin{tabular}{c c c c c} \hline Layer & $h$ & $min_{i}$ & $max_{i}$\\ \hline 1 & 500 & 250 & 1000\\ 2 & 1000 & 500 & 2000\\ 3 & 2000 & 1000 & 4000\\ 4 & 4000 & 2000 & 8000\\ 5 & 8000 & 4000 & 16000\\ 6 & 16000 & 8000 & - \\ \hline \end{tabular} \end{center} \end{table} As anticipated in the Introduction, the MASS has been introduced in the astronomical context by Kornilov \citep{kornilov2003} with a slightly modified version of a technique proposed by \citet{ochs1976}. The MASS reconstructs the turbulence stratification in six layers distributed in the 20~km above the ground starting from the analysis of scintillation indices (four normal and six differential indices) of scintillation maps produced by single stars on a set of four small concentric and circular apertures selected on the telescope's pupil. Scintillation is measured by photo-multipliers. The vertical stratification is obtained by fitting a set of measured scintillation indices with a model having a small and fixed number of turbulent layers. The layers are located at 6 heights: 0.5, 1, 2, 4, 8 and 16~km above the ground. Their positions respect therefore the following law:\newline \begin{equation} h_{i}=2h_{i-1} \end{equation} where $i$ =1,..,6 and $h_{1}$ = 500~m. The scintillation indices depend on the integral of the $\mbox{$C_N^2$}$(h), the turbulent spectrum, the Fresnel diffraction term and the pupil filter. The product of the turbulent spectrum times the Fresnel diffraction term times the pupil filter constitutes the weighting function $W(h)$. The integral of the turbulence in each layer $J_{i}$ (with i=1,..,6 ) is the integral of the $\mbox{$C_N^2$}$ times the weighting function $W(h)$ (see Eq.1 in \citet{tokovinin2003b}):\newline \begin{equation} J_{i}=\int_{min_{i}}^{max_{i}}C_{N}^{2}(h)\cdot W(h)dh \end{equation} where $J$ is expressed in (m$^{1/3}$) and $min_{i}$ and $max_{i}$ represent the extremes of the layer $i$. The MASS weighting functions (WFs) are a sequence of triangles whose peaks are located at the layer heights (0.5, 1, 2, 4, 8 and 16~km) and the base of the triangles are reported in Table \ref{table1}. The sequence of the weighting functions in logarithmic scale along the whole atmosphere is shown in Fig.\ref{wf}. In this figure it is visible how each triangle is interlaced with the previous and successive one. For each height h in the [0.5, 16]~km range is valid the condition that the value of the weighting function $W(h)$ is always equal to 1. This condition guarantees the conservation of the turbulence energy in the vertical slab [0.5, 16]~km. The MASS device used during the PAR2007 site testing campaign was developed by the Kornilov's group at the Sternberg Astronomical Institute in Russia to be employed in the E-ELT site testing campaigns in the northern part of Chile and in Argentina. The temporal sampling of the raw measurements is of around 1 minute. The MASS position at the summit of Cerro Paranal is shown in Fig. \ref{map}. \begin{figure} \centering \includegraphics[width=7cm]{fig2_masciadri} \caption{MASS weighting functions as a function of the height from the ground and valid for the instrument 6-layers configuration. Y-axis is in logarithmic scale. \label{wf}} \end{figure} \subsection{DIMM} \label{dimm} DIMM measurements are routinely done at Cerro Paranal since 1988 in coincidence of the Very Large Telescope Site Testing campaign. The DIMM measures the integrated turbulence energy in the whole atmosphere (the whole optical path covered by the wavefront). The DIMM principle is based on the measurement of the variance of the fluctuations of the angle of arrival of the wavefront coming from single stars and passing through two small holes separated by a certain distance d within the pupil of a telescope \citep{sarazin1990}. The telescope pupil size is typically of 35-40~cm, the holes size $D$ = 6~cm and the distance $d$ = 14~cm. Values of $D$, $d$ and the pupil size of the small telescope can change depending on the prototype provided $d$ $\ge$ 2$\cdot$D. The seeing $\varepsilon$ is retrieved from the longitudinal $\sigma^{2}_{l}$ and transversal $\sigma^{2}_{t}$ variances of the fluctuations of the angles of arrival. The temporal sampling of measurements is of around 1 minute. In this paper DIMM measurements will be compared to the GS ones in order to have more solid arguments for our conclusions on GS versus MASS seeing comparison. We note that the methods used by the DIMM to calculate $\theta_{0}$ \citep{krause-polstorff1993} and $\tau_{0}$ \citep{sarazin2002} are approximated methods. We will not use therefore DIMM estimations as a reference to support the comparison between the GS and MASS estimation of these two parameters. It has been already put in evidence in the literature that important relative errors with respect to the GS taken as a reference can be obtained for $\tau_{0}$ using such methods \citep{masciadri2006,masciadri2013c}. In \citet{masciadri2013c} preliminary results on a comparison GS vs. DIMM for $\theta_{0}$ have been presented. The DIMM location at the summit of Cerro Paranal is shown in Fig.\ref{map}. \section{Analysis} \label{analy} \subsection{Seeing} \subsubsection{MASS versus GS} \label{mass_gs_see} The comparison MASS vs. GS has been done projecting the $\mbox{$C_N^2$}$ profiles of the GS on the weighting functions (Fig.\ref{wf}) that means re-binning the $\mbox{$C_N^2$}$ of the GS on the weighting functions of the MASS. This permits us to integrate the new GS $\mbox{$C_N^2$}$ profiles in each triangle and therefore to compare the relative MASS and GS seeing (or J) in each layer as well as in the sum of all the six layers. To perform the statistical analysis of the comparison we used the bias (Eq.\ref{eq:bias}), the root mean squared error (RMSE) (Eq.\ref{eq:rmse}) and the regression line passing by the origin\footnote{It has been proved in recent studies \citep{lascaux2013} that the correlation coefficient is not a suitable statistical estimator for this kind of analysis. }. The bias provides us information on systematic errors, the RMSE on the statistical errors plus the systematic errors obtained by the two instruments. The regression line tells us how close/far is the distribution of measurements from a $y$ = $x$ analytical linear relationship. \begin{equation} BIAS = \sum_{i=1}^{N}\frac{Y_i-X_i}{N} \label{eq:bias} \end{equation} \begin{equation} RMSE = \sqrt{\sum_{i=1}^{N}\frac{(Y_i-X_i)^2}{N}} \label{eq:rmse} \end{equation} where $X_{i}$ and $Y_{i}$ are the individual GS and the MASS seeing values respectively. $N$ is the total number of elements in the sample. From the bias and the RMSE it is possible to retrieve the bias-corrected RMSE:\newline \begin{equation}\begin{split} \sigma=\sqrt{\sum_{i=1}^{N}\frac{[(X_i-Y_i)-(\overline{X_i-Y_i})]^2}{N}}\\ =\sqrt{RMSE^2-BIAS^2} \label{eq:sigma} \end{split}\end{equation} $\sigma$ contains only the statistical errors and it will be discussed in some cases. \begin{figure*} \begin{center} \includegraphics[width=15cm]{fig3_masciadri} \end{center} \caption{Scattering plots of the MASS and GS measurements of turbulence (seeing values) related to the nights in which simultaneous measurements obtained by the two instruments are available: 12 nights (see Table \ref{tab:nights}). Measurements are sampled on a timescale of 1-minute. GS measurements are re-binned as if they were weighted by the triangle MASS weighting function in order to be compared to MASS measurements. Each figure of the panel reports the measurements distribution related to one of the six layers. The bias refers to ${\varepsilon}_{MASS} - {\varepsilon}_{SCIDAR}$. \label{fig1}} \end{figure*} \begin{figure*} \centering \includegraphics[width=6cm]{fig4a_masciadri} \includegraphics[width=6cm]{fig4b_masciadri} \caption{As Fig.\ref{fig1}. Scattering plot of the seeing as measured simultaneously by the MASS and the GS on the sub-sample of 12 nights (Table \ref{tab:nights}). Left: measurements coming from the first (layer 1) to the sixth (layer 6) layer are considered. Right: measurements coming from the second (layer 2) to the sixth (layer 6) layer are considered. \label{fig2}} \end{figure*} \begin{table} \caption{Number of measurements available during each night from the GS and the MASS to be used for the statistical analysis. \label{table_meas}} \begin{center} \begin{tabular}{c c c } \hline Night & N. of GS meas. & N. of MASS meas. \\ \hline 18/11/2007 & 391 & 64 \\ 20/11/2007 & 371 & 103 \\ 21/11/2007 & 147 & 57 \\ 22/11/2007 & 320 & 203 \\ 17/12/2007 & 345 & 33 \\ 18/12/2007 & 454 & 64\\ 19/12/2007 & 440 & 184\\ 20/12/2007 & 483 & 364 \\ 21/12/2007 & 492 & 97 \\ 22/12/2007 & 491 & 302 \\ 23/12/2007 & 402 & 270 \\ 24/12/2007 & 452 & 166 \\ \hline \end{tabular} \end{center} \end{table} \begin{table*} \centering \caption{Averages, standard deviations and relative errors of the seeing as measured simultaneously by the GS and the MASS on the sub-sample of 12 nights (Table \ref{tab:nights}). Measurements are resampled on a time scale of 1 minutes. Relative error is defined as (Eq.\ref{eq:rel}). Columns 8 and 9 report the relative errors expressed in terms of J (see Eq.\ref{j_epsi}) in this paper (col. 8) and in paper \citep{tokovinin2005} (col. 9).\label{tab:stat} } \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline \multicolumn{2}{|c|}{ } & \multicolumn{2}{|c|}{SCIDAR} & \multicolumn{2}{|c|}{MASS} & \multicolumn{3}{|c|}{ } \\ \hline Height & & Average & std. & Average & std. & Relative Error & Relative Error (J) & Relative Error (J) -T05 \\ (km) & & (arcsec) & (arcsec) & (arcsec) & (arcsec) & ($\%$) & ($\%$) & ($\%$) \\ \hline & 6 layers & 0.66 & 0.26 & 0.45 & 0.23 & -32 & -48 & -5 \\ & 5 layers & 0.50 & 0.21 & 0.40 & 0.19 & -20 & -30 & -\\ \hline 0.5 & Layer 1 & 0.34 & 0.18 & 0.12 & 0.17 & -65 & -82& +100 \\ 1& Layer 2 & 0.20 & 0.17 & 0.09 & 0.12 & -55 & -71& -41 \\ 2& Layer 3 & 0.15 & 0.12 & 0.17 & 0.13 & +13 & +25& -67\\ 4& Layer 4 & 0.16 & 0.10 & 0.10 & 0.09 & -37 &-60 & -57\\ 8& Layer 5 & 0.20 & 0.10 & 0.10 & 0.12 & -50 & -71& +14\\ 16 &Layer 6 & 0.17 & 0.08 & 0.19 & 0.07 & +18 & +20 & -25\\ \hline \end{tabular} \end{table*} Figure \ref{fig1} shows the scattering plots of MASS and GS measurements of turbulence (seeing values) in each of the six layers and related to the nights in which simultaneous measurements obtained by two instruments are available (12 nights - Table \ref{tab:nights}). Fig. \ref{fig2} shows the same scattering plot for the seeing as measured by the MASS and the GS integrated on all the six layers (from the layer located at 0.5~km up to the layer located at 16~km) and on the five layers (from layer located at 1~km up to the layer located at 16~km)\footnote{We preferred to treat in these figures the seeing and not the integrated turbulence J (see later Eq.\ref{j_epsi}) because J should be represented in log-scale and the distribution of the points as well as the regression line can provide distorted visual effects in log-scale.}. Table \ref{table_meas} reports the number of measurements from both GS and MASS used for this analysis. To perform the comparison on simultaneous measurements, the data-set has been resample on a time scale of 1 minute. We verified that results (Fig.\ref{fig1}, Fig.\ref{fig2} and Table \ref{tab:stat}) are substantially the same if we use a time scale of 10 minutes. Also we note that in Fig. \ref{fig1} there are frequent unrealistic zero values by the MASS. These features are produced by the restoration technique as said in Section 3 of \citet{tokovinin2005}. These values have not been considered in the statistical analysis. MASS measurements have been treated with the data-reduction software version Atmos 2.3 that means the same version used (for seeing and isoplanatic angle) in the analysis done in the context of the Thirty Meter Telescope (TMT) site testing study for the selection of the TMT site \citep{schoeck2009} and E-ELT site testing study for the selection of the E-ELT site \citep{vazquez-ramio2012}. Fig.\ref{fig1} tells us that, with exception of layer 3 and 6, the MASS presents in all the other layers, an underestimation of the seeing with respect to the GS (expressed by the bias). If we consider the total seeing on the whole six or five layers (Fig. \ref{fig2}) the underestimation of the MASS is still observed even if, in the five layers case, it is weaker. To better appreciate from a quantitative point of view these differences, Table \ref{tab:stat} reports averages, standard deviations and relative errors of the seeing as measured by the MASS and the GS in each layer and on the total six and five layers. The relative error is calculated as:\newline \begin{equation} r=\frac{\varepsilon _{MASS}-\varepsilon _{GS}}{\varepsilon _{GS}}\times100 \label{eq:rel} \end{equation} We can observe that, even if in absolute terms, all the differences are within 0.22 arcsec, the relative errors can be not necessarily negligible. Layers located at 0.5, 1 and 8~km (layers 1, 2 and 5) present the largest discrepancies between GS and MASS with relative errors respectively of -65, -55 and -50~$\%$. Layers located at 2 and 16~km (layers 3 and 6) show the smallest relative errors respectively of 13 and 18~$\%$. The relative error for the total integrated turbulence on the whole six layers is -32~$\%$. If we do not consider the layer at 0.5~km, i.e. the turbulence integrated in the five layers (from 2 to 6), the relative error remains within -20~$\%$. The same analysis could be done in terms of J i.e. the integral of the turbulence on the whole atmosphere that is related to the seeing $\varepsilon$ by a 5/3 power (see for example \citep{masciadri2010}):\newline \begin{equation} J = 9 \cdot 10^{-11} \cdot \lambda ^{1/3} \cdot \varepsilon ^{5/3} \label{j_epsi} \end{equation} where $\lambda$ is the wavelength and J unit is [m$^{1/3}$]. Due the power 5/3, the relative errors in J are larger as can be seen in column 8 of Table \ref{tab:stat}\footnote{The calculation of the relative error in J terms permits to discuss our results with respect to the \citet{tokovinin2005} paper.}. \begin{figure*} \begin{center} \includegraphics[width=6cm,angle=-90]{fig5a_masciadri} \includegraphics[width=6cm,angle=-90]{fig5b_masciadri} \end{center} \caption{Left: Distribution of the DIMM and GS measurements of total turbulence (seeing values) related to the nights in which simultaneous measurements obtained by the two instruments are available: 20 nights (see Table \ref{tab:nights}). Measurements are sampled on a timescale of 1 minute. The bias refers to ${\varepsilon}_{DIMM} - {\varepsilon}_{SCIDAR}$. Right: cumulative distribution of the total seeing related to the whole atmosphere obtained by the GS and the DIMM. The median values obtained for the two instruments are reported in the figure. \label{fig5}} \end{figure*} To complete the analysis, Fig.\ref{fig4} in Annex shows the temporal evolution of the J values as measured by the MASS and the GS on the total six layers during the whole night of observation for all the 12 nights (Table \ref{tab:nights}). The temporal evolution provides information on the trend during the night. The quantitative estimation is done in any case by Table \ref{tab:stat}. It has been decided to represent the integrated turbulence $J$ (and not the seeing) just because, from a visual point of view, the dynamic between the measurements of the two instruments is better suited for a cross-comparison and our interest is to put in evidence if there are systematic biases. Fig.\ref{layer1} to Fig.\ref{layer6} show the temporal evolution of MASS and GS integrated $J$ in each individual layer for the sub-sample of 12 nights (Table \ref{tab:nights}) in which measurements are available. All the figures have been resampled on a time scale of 30 minutes to make the figures more readable. From these figures it is visible that the underestimation that has been quantified in Table \ref{tab:stat} in some layers is not the result of sporadic events but it is typical of most of the nights. This tells us that, even if the two instruments did not always look at the same direction, the underestimation can hardly be attributed to a different line of sight. On the other side, looking at Fig.\ref{layer1} to Fig.\ref{layer6} it is possible to observe the substantial good agreement of turbulence estimation in layers located at 2 and 16~km (layers 3 and 6). At the same time it is possible to note that the trend of the turbulence observed by the two instruments is substantially well reconstructed, even in those cases in which an underestimation is observed. We finally note that (see Fig.\ref{fig1}) the $\sigma$ in the individual layers is within [0.06-0.15] arcsec range with the largest $\sigma$ in the layer closest to the ground and the smallest $\sigma$ in the highest layer as it is expected. This tells us that the turbulence increases its horizontal homogeneity (small $\sigma$) when we go at high heights. The $\sigma$ increases near the ground due to the orographic effects. To have some further elements that could confirm us that the problem in these estimates is represented by the MASS and not the GS we performed a comparison between GS and DIMM measurements that is discussed in the next Section. \subsubsection{DIMM versus GS} Fig. \ref{fig5}-left shows the scattering plot of the total turbulence as measured by the DIMM and GS on those nights of PAR2007 campaign in which simultaneous measurements are available (20 nights - see Table \ref{tab:nights}). Table \ref{table_meas_gsdimm} reports the number of measurements from both GS and DIMM used for this analysis. Measurements are resampled on a time scale of 1 minute\footnote{Also in this case we verified that results are basically the same if the resample is done on a time scale of 10 minutes (bias = 0.11~arcsec and RMSE = 0.33~arcsec) with respect to a bias of 0.11~arcsec and a RMSE of 0.36~arcsec for the time scale of 1 minute.}. This figure of merit tells us that the correlation between DIMM and GS is substantially good. Fig. \ref{fig5}-right shows the cumulative distribution of the same set of measurements in which it is visible that the two data-set have in most cases, the same statistical behavior. The bias of 0.11 arcsec tells us that the agreement is within 9$\%$ (the mean seeing for the GS and the DIMM are respectively 1.16 and 1.27~arcsec). Looking at Fig. \ref{fig5}-right we observe that for seeing $\le$ 1 arcsec the statistical behavior of the two data-set is almost perfect. A good agreement (within 10$\%$) is observed for the median value of the seeing. A relative error up to 20$\%$ is observed for seeing values in the [1-2]~arcsec range. The relative error is again within 10$\%$ for seeing larger than 2~arcsec. The excess of turbulence measured by the DIMM when the seeing is between 1 and 2 arcsec is due to the fact that the two instruments, even if they are located at the same height above the ground on a substantial flat surface, are separated by a distance of $\sim$ 205~m. The DIMM, in particular, is located at the extremity of the plateau obtained after the cut of the mountain summit performed to build the VLT and it points always at South i.e. towards the summit plateau. It has been noted on routinely observations (Lombardi private communication) that on the ridge of the plateau there is a tendency in creating a local excess of turbulence particularly when the seeing is strong. The local excess relates, of course, turbulence in the surface. Because of the fact that the DIMM points always at South, this excess is typically detected by the DIMM\footnote{It has been observed (Lombardi private communication) that if the DIMM points towards North this effect seems mitigated or even disappears. The GS has not a privileged direction for the lines of sight.}. This highly probably generates the slightly difference in seeing observed between the GS and the DIMM in the range [1-2] arcsec. We note however that this difference is much weaker in relative terms than that observed between the MASS and the GS. Even more important, the GS/DIMM difference can be justified by the local turbulence in the surface layer that can not be equally sensed by the two instruments but the difference between the MASS and the GS can not be explained by a different line of sight because first, the MASS vs. GS comparison treats just turbulence in the free atmosphere and therefore the argument of the local excess should be a no sense. Second, even if it is well possible that the MASS and the GS are looking at different line of sights in different instants during the night, this fact might eventually cause a temporary (i.e. locally in time) discrepancy but not a systematic bias between the two instruments at specific heights above the ground all along the time. Considering that the GS and the MASS have no privileged lines of sight, if one looks at the problem from a statistical point of view, the potential different line of sight between the two instruments during the night would produce at some instants an excess of turbulence from the GS and in some other instants an excess of turbulence from the MASS and not always an excess of turbulence from the GS. In other words, the line of sight direction is irrelevant to comment the bias observed between the MASS and the GS at some specific heights. With respect to our main analysis goal i.e. the comparison between the MASS and the GS, the cross-comparison between the DIMM and the GS supports the thesis that we are in front of an underestimation of the MASS and not of an overestimation of the GS. The second assumption should imply, indeed, that the GS seeing should be weaker than what is observed with a consequent lack of correlation with the DIMM. \begin{table} \caption{Number of measurements available during each night from the GS and the DIMM to be used for the statistical analysis. \label{table_meas_gsdimm}} \begin{center} \begin{tabular}{c c c } \hline Night & N. of GS meas. & N. of DIMM meas. \\ \hline 10/11/2007 & 322 & 490 \\ 11/11/2007 & 361 & 485 \\ 13/11/2007 & 290 & 473 \\ 14/11/2007 & 356 & 447 \\ 17/11/2007 & 303 & 459 \\ 18/11/2007 & 391 & 495 \\ 20/11/2007 & 371 & 490 \\ 21/11/2007 & 147 & 270 \\ 22/11/2007 & 320 & 427 \\ 24/11/2007 & 361 & 471 \\ 15/12/2007 & 299 & 518 \\ 16/12/2007 & 423 & 473 \\ 17/12/2007 & 345 & 471 \\ 18/12/2007 & 454 & 418 \\ 19/12/2007 & 440 & 513 \\ 20/12/2007 & 483 & 497 \\ 21/12/2007 & 492 & 469 \\ 22/12/2007 & 491 & 509 \\ 23/12/2007 & 402 & 482 \\ 24/12/2007 & 452 & 487 \\ \hline \end{tabular} \end{center} \end{table} \subsection{Isoplanatic angle: $\theta_{0}$} \label{mass_gs_iso} Fig. \ref{theta0} shows the isoplanatic angle ($\theta_{0}$) as observed by the MASS and the GS on those nights in which simultaneous observations are available. Measurements are resampled on a time scale of 1 minute and show a bias = -0.07~arcsec and a RMSE = 0.53~arcsec\footnote{We verified that if the resample is done on time scales of 10 and 30 minutes results are very similar for the bias (-0.09~arcsec) and slightly inferior for the RMSE (0.46~arcsec)} with a relative error of -3$\%$ (GS average: $\theta_{0}$ = 2.11~arcsec; MASS average: $\theta_{0}$ = 2.07~arcsec). The $\theta_{0}$ from the GS is obtained through the $\mbox{$C_N^2$}$:\\ \begin{equation} \theta _{0}=0.057\cdot \lambda ^{6/5}\cdot [\int_{0}^{\infty }h^{5/3}C_{N}^{2}(h)dh]^{-3/5} \end{equation} The $\theta_{0}$ from the MASS is retrieved from the calculation of the scintillation indices and it is an output of the software Atmos 2.3. This means that MASS and GS use different methods to calculate $\theta_{0}$. The sample of the simultaneous measurements between MASS and GS for $\theta_{0}$ is made of 12 nights, the same nights used for the statistical analysis of the seeing. We precise that, because of the method used by the MASS to provide measurements of the seeing and $\theta_{0}$, an observed seeing value is not automatically associated to the corresponding $\theta_{0}$. This is due to the fact that the $\theta_{0}$ is not obtained by the integral of the turbulence on the whole atmosphere but it is based on the scintillation indices. The simultaneous measurements of seeing and $\theta_{0}$ do not cover therefore necessarily the same temporal range. The value of the bias (Fig.\ref{theta0}) tells us that there is no a systematic error between the two instruments as confirmed also by the regression line. The good agreement observed for the $\theta_{0}$ is not in contradiction with the results obtained for the seeing (Section \ref{mass_gs_see}) that, on the contrary, put in evidence some discrepancies between the two instruments. It is indeed known that $\theta_{0}$ is mainly affected by the turbulence in the high part of the atmosphere because $\theta_{0}$ $\sim$ $h^{5/3}$$\cdot$$\mbox{$C_N^2$}$(h). Our analysis on the seeing told us that the MASS and GS correlation is substantially good in the layer 6 i.e. the highest layer (see Fig. \ref{fig1}, Table \ref{tab:stat} and Fig.\ref{layer1} to Fig.\ref{layer6}). In other words, the OT discrepancies observed in other parts of the atmosphere simply do not affect the $\theta_{0}$ in a significative way from a quantitative point of view. Fig.\ref{iso_te} in Annex shows the temporal evolution of the $\theta_{0}$ in all the 12 nights in which GS and MASS measurements are available. It is possible to observe how the temporal evolution of $\theta_{0}$ as measured by the two instruments is substantially well correlated in terms of measurements trend. We note that, in four of the twelve nights, the statistical sample of measurements is small. It would be, therefore, worth to analyze a richer statistical sample to confirm definitely these conclusions. As a final consideration we remind that, the fact that an agreement is observed from a statistical point of view, does not exclude that on individual nights, even for periods of a few hours, it is possible to have important discrepancies between the two instruments. For example, during the night 20/11 (from around 01:00 and 03:00 UT) and 24/12 (from around 04:00 and 06:00 UT) we observe a relative error of the order of 30-50 $\%$. This is due to the different turbulence estimation in the highest MASS layer (see Fig.\ref{layer6}). This kind of differences, in specific temporal intervals, might be attributed to different lines of sight of the two instruments. It is important to note, however, that, this is just a possible explanation. With that we mean that this is just an assumption and, in principle, we can not exclude that the difference is not due to a failure of an instrument. If we assume that the discrepancy is due to different lines of sight of the instruments this would imply the fact that in the high part of the atmosphere (layer 6), there are spatial inhomogeneities in the turbulence distribution on the horizontal plane enough large to produce differences in $\theta_{0}$ of the order of 30-50 $\%$. Such an horizontal inhomogeneity, also in the high part of the atmosphere, would confirm results obtained with a GS by \citet{masciadri2002}\footnote{This was the first evidence in the literature, at least at our knowledge, of the finite horizontal size of turbulence layers at different heights in the 20~km a.g.l. }. It is a fact that the two instruments were not looking at the same direction in that particular interval of time therefore the assumption might be reasonable. \begin{figure} \centering \includegraphics[width=6cm]{fig6_masciadri} \caption{Distribution of the MASS and GS measurements of the isoplanatic angle $\theta_0$ related to the nights in which simultaneous measurements obtained by the two instruments are available: 12 nights (see Table \ref{tab:nights}). Measurements are sampled on a timescale of 1 minute. The bias refers to $\theta_{0,MASS}$ - $\theta_{0,GS}$. \label{theta0}} \end{figure} \subsection{Wavefront coherence time: $\tau_{0}$} \label{mass_gs_tau0} The wavefront coherence time ($\tau_{0}$) calculated by the GS is obtained from:\newline \begin{equation} \tau_{0}=0.057\cdot \lambda^{6/5}\cdot \left [ \int_{0}^{\infty } \left | V(h) \right |^{5/3}\cdot C^{2}_{N}(h)\cdot dh\right ]^{-3/5} \end{equation} where the $\mbox{$C_N^2$}$ profiles are measured by the GS and the wind speed vertical profiles have been reconstructed by the atmospherical model Meso-Nh. In previous papers \citep{hagelin2010,masciadri2013a} it has been proved that such an atmospherical model provides very reliable wind speed vertical profiles. The most solid validation of the method has been carried out through a comparison of wind speed profiles provided by the model with radiosoundings on a sample of 53 launches distributed on 23 nights in summer and winter \citet{masciadri2013a}. The MASS wavefront coherence time on the whole atmosphere (hereafter total MASS $\tau_{0,tot}$) is defined as \citep{kornilov2011}: \begin{equation} \tau_{0,tot}=\left [ (\tau_{0,GL})^{-5/3}+(\tau_{0,MASS})^{-5/3} \right ]^{-3/5} \label{eq:tau_mass} \end{equation} where $\tau_{0,MASS}$ comes from the MASS covering the contribution of the free atmosphere and:\newline \begin{equation} \tau_{0,GL}=0.314\cdot \frac{r_{0,GL}}{V_{30m}} \end{equation} is the $\tau_{0}$ contribution associated to the boundary layer. $r_{0,GL}$ is the Fried parameter associated to the near-ground contribution of r$_{0}$ obtained subtracting the free atmosphere turbulent energy (MASS) from the total turbulent energy (DIMM). $V_{30m}$ is the wind speed measured at 30~m above the ground by the Automatic Weather Station (AWS) \citep{sandrock99}. $r_{0,GL}$ is obtained from:\newline \begin{equation} r_{0,GL}=\left [ (r_{0,DIMM})^{-5/3}-(r_{0,MASS})^{-5/3} \right ]^{-3/5} \end{equation} \noindent For the estimation of the MASS $\tau_{0,MASS}$ we used the Atmos 2.97.3 software version that implements a different method for the calculation of $\tau_{0}$ with respect to what was done in the Atmos 2.3 version \citep{kornilov2011}\footnote{The new $\tau_{0}$ is obtained indeed, taking into account a finite exposure time and a weighting function depending on the wind speed. }. It is known that the Atmos 2.3 version used for the seeing and $\theta_{0}$ does not provide reliable $\tau_{0}$ estimates \citep{kornilov2011}. For the $\tau_{0,tot}$ we have simultaneous measurements obtained with the MASS and the GS on a sample of 14 nights (all the 12 nights included in the sample used for the seeing and $\theta_{0}$ (Table \ref{tab:nights}) plus the nights of 15/12/2007 and 16/12/2007. Fig. \ref{tau0} shows the scattering plot of the $\tau_{0}$ as observed by the MASS and the GS on those nights in which simultaneous observations are available. Measurements are resampled on a time scale of 10 minutes\footnote{Due to the fact that the MASS $\tau_{0,tot}$ is obtained using measurements coming from the DIMM, the MASS and the AWS, we resampled the measurements on a time scale that is a common multiple of the sampling of the individual data-set.}. We can observe that the agreement between the two instruments is substantially good with a bias = 0.20~msec with a relative error of 1$\%$ (GS average: $\tau_{0}$ = 4.49~msec; MASS average: 4.54~msec). Fig.\ref{tau0_te} in Annex shows the temporal evolution of the $\tau_{0}$ in all the 14 nights in which GS and MASS measurements are available. It is possible to observe how the temporal evolution of $\tau_{0}$ as measured by the two instruments is substantially well correlated also in terms of measurements trend. If we look at the individual nights we note in a couple of nights (23/12 and 24/12) an important relative error (even $>$ 60 $\%$) in the interval of a few hours. \begin{figure} \centering \includegraphics[width=6cm]{fig7_masciadri} \caption{Distribution of the MASS and GS measurements of the wavefront coherence time $\tau_0$ related to the nights in which simultaneous measurements obtained by the two instruments are available: 14 nights. Measurements are sampled on a timescale of 10 minute. The bias refers to $\tau_{0,MASS}$ - $\tau_{0,GS}$. \label{tau0}} \end{figure} However, because of the fact that the $\tau_{0}$ comes from the product of the wind speed and the $\mbox{$C_N^2$}$ profiles and considering that we observed some discrepancies of the turbulence ($\mbox{$C_N^2$}$) detected on individual layers and on the free atmosphere by the MASS and the GS, one could wonder how an agreement of the MASS and GS $\tau_{0}$ is possible. One should expect an overestimation of the $\tau_{0}$ from the MASS because the MASS underestimate the seeing in the free atmosphere. Why is this not observed ? A probable explanation is the following. $\tau_{0,tot}$ is given by the addition of two contributions: the first one comes from the MASS ($\tau_{0,MASS}$); the second one comes substantially from the DIMM ($\tau_{0,GL}$) ($r_{0,MASS}$ is indeed much weaker than $r_{0,DIMM}$). Looking at Fig. \ref{tau0_te}, we note that, in the second part of the campaign (starting from 19/12), where the absolute value of $\tau_{0}$ is large, the MASS $\tau_{0}$ shows some tendency in giving a larger $\tau_{0}$ value with respect to the GS. We verified (Fig. \ref{tau0_gl_vs_fa}) that the nights characterized by a small $\tau_{0}$ ($\tau_{0}$ $<$ 5 msec) correspond to the cases in which the $\tau_{0,GL}$ contribution is dominant with respect to the $\tau_{0,MASS}$. The term 'dominant' means smaller because the $\tau_{0}$ respects Eq.\ref{eq:tau_mass}. In the nights characterized by a large $\tau_{0}$ ($\tau_{0}$ $>$ 5 msec) we can have both cases even if it is more frequent that the $\tau_{0,MASS}$ contribution is dominant. This indicates that the turbulence and/or the wind speed (or even both of them) in the boundary layer are so weak that the $\tau_{0}$ contribution from the free atmosphere becomes dominant. In those cases the overestimation of the MASS starts to be evident but on the whole sample of nights the DIMM contribution is dominant with respect to the MASS one. Considering that the typical $\tau_{0}$ median value of a good site is of the order of 4-5 msec, i.e. close to the threshold that we indicated, we think that it would be very useful, for precautionary principle, to do in the future the same comparison of MASS vs. GS $\tau_{0}$ on a richer statistical sample to confirm the conclusion of a good correlation and to confirm the fact that the overestimation tendency of the MASS remains statistically not so important. \section{Discussion} \label{disc} We observe that, even if our results as well as those obtained by \citep{tokovinin2005} (hereafter T05) show some discrepancies between the MASS and the GS measurements, the two studies reached in some cases similar results, in other cases different results. Table \ref{tab:stat} columns 8 and 9 report the relative errors calculated in this paper and in the T05 paper with respect to the integrated turbulence J in the free atmosphere and in the individual layers. As already said in Section \ref{mass_gs_see} the relative errors expressed in seeing (instead of J) are smaller (see col.7 of Table \ref{tab:stat}) but we calculated also the relative errors in J because T05 used J. A premise is necessary: due to the fact that the statistical samples of the two studies are relatively small and in any case not equal (12-14 nights for our study and 4 nights for the T05 study) we can not expect necessarily equivalence in the absolute value of the relative errors. What should be reasonable to expect is, however, at least a similar behavior for example, the evidence of an excess or a default or a matching of turbulence in specific regions of the atmosphere. For what concerns the vertical distribution of J in the individual layers the similarities between the two studies are: {\bf (a)} in the layer located at 16~km (layer 6) we both found a good agreement with a relative error within +20$\%$ (this paper) and within -25$\%$ (T05 paper) even if in opposite direction; {\bf (b)} in the layer located at 1~km (layer 2) we both found that MASS underestimates the seeing with respect to the GS with a relative error of -71$\%$ (this paper) and -41$\%$ (T05 paper); {\bf (c)} we both found a substantial underestimation of J by the MASS with respect to the GS in the layer located at 4~km (layer 4) with a relative error of -60$\%$ (this paper) and -57$\%$ (T05).\\ Differences between the two studies are: {\bf (a)} we found that the MASS underestimates J in the layer located at 0.5~km (layer 1) with a relative error of -82$\%$ while \citet{tokovinin2005} found that the MASS overestimates the GS with a relative error of +100$\%$; {\bf (b)} we found a substantial good agreement of J in layer located at 2~km (layer 3) with a relative error within +25$\%$ while T05 found that the MASS underestimates J with respect to the GS with a relative error of -67$\%$; {\bf (c)} we found that the MASS underestimates J in the layer located at 8~km (layer 5) with a relative error of -71$\%$ while T05 found a good agreement between the two instruments with a relative error within +14$\%$; {\bf (d)} we found a substantial underestimation by the MASS with respect to the GS of the J in the free atmosphere seeing (obtained as the contribution coming from all the six layers) with a relative error of -48$\%$ while T05 found a substantial good agreement between the two instruments with a relative error of -5$\%$. In synthesis our study put in evidence the fact that in almost all the layers (with exception of layers located at 2 and 16~km i.e. layers 3 and 6) there are important discrepancies between the turbulence observed by the GS and the MASS. Also we observe that the integration of the turbulence in the free atmosphere is underestimated by the MASS by a not negligible factor (-48$\%$ in J terms and -32$\%$ in seeing terms). T05 founded some not negligible discrepancies in layers located at 0.5 and 4~km but a substantial good agreement in estimating the free atmosphere by the two instruments (within 5$\%$). The most important difference between the results of the two studies is, therefore, the result on the free-atmosphere integration. The sample of 4 nights of T05 provides a statistical substantial good agreement while we obtain a statistical substantial underestimation of the MASS with respect to the GS on a sample of 12 nights. Our results are, in principle, more reliable from a statistical point of view (12 nights against 4 nights). We think, however, that it would be useful to repeat the calculation on a richer statistical sample to confirm these results. They indicate, indeed, biases on the MASS estimations when it is used as a monitor of the optical turbulence in the free-atmosphere. It is therefore important to fix this point to use correctly this instrument in this context. The other differences found in the two studies are more difficult to be justified. We think however that the most worried thing is the fact that visibly important discrepancies between the layers are detected in both studies. A richer statistical sample would certainly help in better understanding the status of art. We are promoting, in the context of an international working group \citep{masciadri2013b}, a site testing campaign aiming at establish absolute calibration of different vertical profilers among those, the MASS and the GS. It is evident that such not negligible discrepancies in individual layers can not be considered satisfactory for some applications such as the one in which we are interested on, for example the identification of a reference for a model validation. Discrepancies on individual layers are also very critical in application to the optimization of all kind of AO systems that are particularly sensitive to the position and strength of the individual turbulent layers. For the same reasons the MASS appears not very suitable for the calibration of the atmospherical models for the OT forecast. We report here few words on the consequences of our results on previous studies done in the context of the site search/selection. We note that, in spite of evident OT discrepancies on individual layers, $\theta_{0}$ appears well correlated between the MASS and the GS. There are therefore no major implications that might contradict results obtained previously. The free atmosphere seeing was probably underestimated (and the boundary layer overestimated because it is calculated by subtraction of the MASS contribution from the DIMM one) but we note that this happened for all the sites and, due to the fact that the site selection is an exercise done in relative terms, the consequences are therefore not so critical. For the $\tau_{0}$ we remind that, at the epoch of the TMT and E-ELT site selection the Atmos 2.97.3 software that we used was not available. The two teams used different approaches to calculate $\tau_{0}$. In the TMT case \citep{travouillon2009} the MASS software was corrected by a constant factor (1.73) found fitting measurements with wind speed radiosoundings and the order of magnitude of the final $\tau_{0}$ was included between 4.2 and 5.6 msec. In the E-ELT case \citep{vazquez-ramio2012} $\tau_{0}$ was calculated using the $\mbox{$C_N^2$}$ from MASS and DIMM plus the wind speed from the Global Data Assimilation System (GDAS) database, with a 1$^{\circ}$ horizontal resolution of the Air Resources Laboratory (ARL) of the National Oceanic and Atmospheric. It is known, however, that these data coming form the General Circulation Models have a too low resolution to well represent the wind speed in the low part of the atmosphere, particularly on mountain regions \citep{masciadri2013a}. Limiting our commentaries to the potential consequences of the results of this paper on previous $\tau_{0}$ estimations we can simply say that, apart the fact that in both cases they used the $\mbox{$C_N^2$}$ in the free atmosphere from the MASS that was, highly probably, underestimated, it would be a hazard to deduce any other consequences from our results. We limit ourselves to observe that the order of magnitude of the $\tau_{0}$ median values found in those studies fit with the expected order of magnitude. \section{Conclusions} \label{concl} In this study we present the results of a detailed comparison of the turbulence stratification in the free atmosphere as well as of the astroclimatic parameters (seeing, isoplanatic angle and wavefront coherence time) as measured by the MASS and the GS. The main results that we obtained tell us that:\newline \begin{itemize} \item the MASS underestimates the integrated turbulence (J or seeing) in the free atmosphere with respect to the GS with a relative error of -32$\%$ in seeing terms (-48$\%$ in J terms); a bias = -0.21~arcsec and a RMSE = 0.25~arcsec if we consider the 6 layers of the MASS. If we consider the highest 5 layers (from layer 2 to layer 6) the relative error decreases to -20$\%$ in seeing terms (-30$\%$ in J terms); the bias = -0.1 arcsec and the RMSE = 0.16~arcsec. We find important discrepancies between MASS and GS in all the individual layers (reaching relative errors as high as -65$\%$ in seeing terms and -82$\%$ in J terms) with exception of the layers located at 2 and 16~km (layers 3 and 6) in which the relative error remains limited to +18$\%$ in seeing terms (+20$\%$ in J terms); \item the unique study previously published on a similar topic (even if it was applied on a poorer statistical sample, \citet{tokovinin2005}) revealed relative errors on individual layers as large as those we observed in our study; \item the main difference between our results and those of \citet{tokovinin2005} is that they found a substantial good agreement of the free-atmosphere seeing as measured by the MASS and the GS on a statistical sample of 4 nights while we find that the MASS underestimates the free-atmosphere seeing with respect to the GS on a statistical sample of 12 nights. The relative error is -32$\%$ in seeing terms (-48$\%$ in J terms). Even if our estimations are statistically more reliable, it should be useful to confirm our results on a richer statistical sample. This result, if confirmed, should have important consequences on the main use of the MASS done so far i.e. as a monitor of the free-atmosphere seeing; \item the isoplanatic angle ($\theta_{0}$) appears substantially well correlated between the two instruments (bias = -0.07~arcsec; RMSE = 0.53~arcsec and relative error r = 3$\%$) because the most important contribution to $\theta_{0}$ comes from the OT in the highest layer (layer 6) that is well correlated between the two instruments; \item the wavefront coherence time ($\tau_{0}$) is well correlated between the two instruments (bias = 0.20~msec; RMSE = 1.52~msec and relative error r = 1$\%$). The underestimation of the MASS in quantifying the free atmosphere turbulence seems to have a minor effect on the $\tau_{0}$ because the boundary layer contribution from the DIMM ($\tau_{0,GL}$) has, in many cases, a dominant role with respect to the contribution from the MASS ($\tau_{0,MASS}$). It should be useful to repeat the same calculation on a richer statistical sample to confirm this thesis. We observed that, the cases in which biases induced by the overestimation of the free atmosphere turbulence from the MASS can be more evident, is when $\tau_{0}$ is large ($>$ 5 msec); \item we proved that, contrary to what is frequently believed, the boundary layer contribution of the $\tau_{0}$ has, most of the time, a more important impact from a quantitative point of view than the free atmosphere one.; \item even if the good correlation of some astroclimatic parameters has been observed in statistical terms ($\theta_{0}$ and $\tau_{0}$) this does not exclude that, on individual nights, on intervals of a few hours, important discrepancies are observed between the two instruments. We therefore warn in using appropriately the conclusions of this study depending on the contexts; \item we proved that conclusions reached in this study definitely would require a richer statistical sample of measurements to be confirmed. An international action is on-going \citep{masciadri2013b} to promote a site testing campaign including classical and new-generation vertical profilers aiming at quantifying the absolute values of the optical turbulence and the intrinsic uncertainty we have to consider for each astroclimatic parameter and at validating the new-generation instruments. The intention is to collect measurements on rich statistical samples (order of two weeks in summer and two weeks in winter). The MASS is supposed to be one of these instruments i.e. we will be able to verify these results on a richer statistical sample. Besides, such great relative errors on individual layers as well as on the free atmosphere suggests ancillary studies based on dedicated simulations to identify the origin of the MASS discrepancies; \item our results indicate that the MASS is not necessarily a suitable tool for some applications such as the calibration of atmospherical model for the OT forecasts and all applications in which it is crucial to know the location and the OT strength of individual layers such as the AO; \item concerning the impact that these results can have on site selection studies previously done (TMT and E-ELT) we showed that in some cases there were no consequence at all. In other cases we suspect some biases but, due to the fact that the site selection is a relative terms exercise and therefore the biases should be applied to all sites, we think that these results should have no major consequences on the most important goal of these activities i.e. the selection of the site for the respective telescopes. \end{itemize} \section*{Acknowledgments} This study is co-funded by the ESO contract: E-SOW-ESO-245-0933 (MOSE Project). We are very grateful to the ESO Board of MOSE (Marc Sarazin, Pierre-Yves Madec, Florian Kerber and Harald Kuntschner) for their constant support to this study. This study takes advantage of measurements performed during the PAR2007 site testing campaign promoted by ESO in the context of the E-ELT Design Study (FP6 Program). Different international teams from Laboratoire Lagrange - Universit\'e de Nice-Sophia Antipolis (France), IAC (Spain), ESO (Germany), Sternberg Astronomical Institute (Russia) participated and/or have been involved in the PAR2007 campaign.
\section{Introduction} The purpose of this article is to study the weak-type $(1,1)$ boundedness of the operator \[ \mathcal{A} f = \sum_{Q \in \mathcal{D}(I)} \alpha_Q \langle f \rangle_Q \mathbbm{1}_Q. \] Here $I$ denotes any finite interval in $\mathbb{R}$, $\langle f \rangle_I = \frac{1}{|I|}\int_I f$, $\mathcal{D}(I)$ denotes the dyadic grid consisting of dyadic subintervals of $I$ and $\{\alpha_J\}_{J \in \mathcal{D}(I)}$ is a Carleson sequence adapted to $I$, i.e.: $\alpha_J \geq 0$ for all $J \in \mathcal{D}(I)$ and \[ \sup_{J \in \mathcal{D}(I)} \frac{1}{|J|} \sum_{K \in \mathcal{D}(J)} \alpha_K |K| = C <\infty. \] These operators have recently appeared in the works of A. K. Lerner \cite{Lerner2010} and \cite{Lerner2012}, where $\alpha_K$ was a binary sequence, although the ideas go back to \cite{Garnett1982}. Hence, we will call them \textit{Lerner operators} in the sequel. Here we find the exact Bellman function describing the local boundedness of $\mathcal{A}$ from $L^1$ to $L^{1,\infty}$. It is easy to see that the operator $\mathcal{A}$ is bounded in $L^2$. This, together with a decomposition of Calder\'on-Zygmund type, can be used to prove an estimate of the form \[ \Bigl|\Bigl\{x\in I: \, \Bigl| \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x)\Bigr| > \lambda\Bigr\}\Bigr| \leq \frac{C}{\lambda} \int_{I} |f|. \] However, here we precisely describe how the best constant in the above inequality changes with respect to the parameters of the problem. The main result of the article is the following theorem: \begin{theorem}\label{MainTheorem} Let $A$, $\lambda$ and $t$ be positive numbers and $I$ an interval in $\mathbb{R}$, then \[ \sup \frac{1}{|I|}\Bigl|\Bigl\{x\in I: \, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda\Bigr\}\Bigr| = \begin{cases} \frac{2At}{A\lambda+t} & \text{if } 0 \leq t \leq A\lambda \leq \lambda, \\ \vspace{-10pt}\\ \sqrt{\frac{At}{\lambda}} &\text{if } 0 \leq A \leq \min\Bigl( \frac{t}{\lambda}, \frac{\lambda}{t} \Bigr), \\ 1 &\text{otherwise.} \end{cases} \] Where the supremum is taken over all nonnegative functions $f$ with $\langle f \rangle_I = t$ and all nonnegative sequences $\{\alpha_J\}_{J \in \mathcal{D}(I)}$ with Carleson constant at most $1$ which satisfy \[ \frac{1}{|I|} \sum_{J \in \mathcal{D}(I)} \alpha_J|J| = A. \] \end{theorem} We also provide a sequence of examples which, in the limit, attain the supremum of the previous result. See the last section for details on the structure of such examples. As an immediate corollary we have the following local weak-type (1,1) estimate: \begin{corollary} For any nonnegative $f \in L^1([0,1))$ and for any Carleson sequence $\{\alpha_J\}_{J \in \mathcal{D}([0,1))}$ with constant at most $1$ we have the sharp bound \[ \bigl| \bigl\{ x\in [0,1):\, \mathcal{A}f(x) > \lambda \bigr\} \bigr| \leq \begin{cases} \frac{2\|f\|_{L^1}}{\lambda + \|f\|_{L^1}} & \text{if } \|f\|_{L^1} \leq \lambda \\ 1 & \text{if } \|f\|_{L^1} \geq \lambda, \end{cases} \] which in particular implies that \[ \|\mathcal{A}f\|_{L^{1,\infty}([0,1))} \leq 2\|f\|_{L^1([0,1))}, \] and that the constant $2$ is sharp. \end{corollary} Operators similar to these were recently studied in \cite{Melas2005}, \cite{Melas2008}, \cite{Melas2009} and \cite{Melas2009a}, however their results are slightly different from ours. They consider the supremum taken over all functions $f$ satisfying \[ \int_I f = s \quad \text{and} \quad \int_I G(f) = t, \] where $G$ is a strictly convex function satisfying $G(x)/x \to \infty$ as $x \to \infty$. This does not include the question of boundedness from $L^1$ to $L^{1,\infty}$. Our method of proof is different than the one used in the articles cited above, where they use the deep combinatorial properties of these operators. See also the monograph \cite{OsekowskiBook} by A. Os{\setbox0=\hbox{e}{\ooalign{\hidewidth \lower1.5ex\hbox{`}\hidewidth\crcr\unhbox0}}}kowski for related results. We instead follow the ideas in \cite{Slavin2008} and \cite{Vasyunin2009} to solve the Bellman PDE and prove its sharpness. This problem is also closely related to studying Haar shifts, the main difference being that Haar shifts are not positive operators. It has been shown however, see \cite{Cruz-Uribe2012}, that Lerner-type operators can be used to bound Haar shifts. The reader can find results similar to ours in \cite{RVV-UnweightedMartingale}, \cite{NRVV-WeightedHaarShifts} and \cite{NV-Square}. The article is organized as follows. In Section 2 we explain how the Bellman function technique is used to compute the supremum in Theorem \ref{MainTheorem}. In Section 3 we give a supersolution to the Bellman variational problem which serves as an upper bound for the exact Bellman function. Finally, in Section 4 we show that the function we found in the previous section is the exact Bellman function, we also give a sequence of examples which, in the limit, extremize the inequality of Theorem \ref{MainTheorem}. \section*{Acknowledgements} The authors would like to thank Alexander Volberg for originally proposing the problem and for many valuable discussions. \section{The Bellman function technique} Consider the function defined in $\Omega = \{(t, A, \lambda):\, 0 \leq t, 0 \leq A \leq 1, \, \lambda \in \mathbb{R}\}$ \[ \mathbb{B}(t, A,\lambda) = \sup\Bigl\{ \frac{1}{|I|}\Bigl| \bigl\{x\in I:\, \sum_{J \in \mathcal{D}(I)}\alpha_J \langle f \rangle_J \mathbbm{1}_J > \lambda \bigr\} \Bigr|\Bigr\}, \] where the supremum is taken over all all nonnegative functions $f$ on $I$ with $\langle f \rangle_I = t$ and all Carleson sequences $\{\alpha_J\}_{j \in \mathcal{D}(I)}$ with constant at most $1$ and \[ A = \frac{1}{|I|}\sum_{J \subseteq I} \alpha_J |J|. \] Note that $I$ is \emph{not} a parameter in $\mathbb{B}$, this is because the supremum is invariant under dilations and translations in $I$, and hence independent of $I$. The Bellman function technique, which first appeared in the 1995 preprint version of \cite{Nazarov1999}, is based on showing that $\mathbb{B}$ solves a certain minimization problem. One first shows that $\mathbb{B}$ satisfies a kind of concavity property and explicitly computes $\mathbb{B}$ in a subdomain natural to the problem (this is usually easy). Then one shows that any continuous positive function satisfying these conditions majorizes $\mathbb{B}$, which reduces the problem to finding the smallest function which satisfies these properties. Finally one has to actually find such a function, this is usually the hardest part. The reader can find insightful introductions in \cite{Nazarov2001} and \cite{Nazarov1996}, see also \cite{Nazarov1999}, \cite{Slavin2008}, and \cite{Vasyunin2009} for more examples of this technique. Let us begin by describing more precisely the concavity property which $\mathbb{B}$ satisfies: \begin{lemma}[Main inequality]\label{Bellman.MainInequality} \begin{equation}\label{Bellman.MainInequality.eq} \mathbb{B}(t,A,\lambda) \geq \frac{1}{2}\Bigl( \mathbb{B}(t_1,A_1,\lambda') + \mathbb{B}(t_2,A_2,\lambda') \Bigr) \end{equation} whenever \[ t = \frac{t_1+t_2}{2}, \quad A = \frac{A_1+A_2}{2} + \alpha \quad \text{and} \quad \lambda = \lambda' + \alpha t \] and $\alpha \geq 0$. \end{lemma} \begin{proof} Consider any dyadic interval $I$, any function $f \geq 0$ satisfying \[ \langle f \rangle_{I_-} = t_1 \quad \text{and} \quad \langle f \rangle_{I_+} = t_2 \] and any Carleson sequence $\{\alpha_J\}_{J \in \mathcal{D}(I)}$ with constant at most $1$ on $I$ satisfying \[ \frac{1}{|I_-|} \sum_{J \in \mathcal{D}(I_-)} \alpha_J |J| = A_1, \quad \frac{1}{|I_-|} \sum_{J \in \mathcal{D}(I_+)} \alpha_J |J| = A_2 \quad \text{and} \quad \alpha_I = \alpha. \] Suppose also that $\lambda = \lambda' + \alpha t$. Since $\langle f \rangle_I = t$ then we must have \begin{align*} \mathbb{B}(t,A,\lambda) &\geq \frac{1}{|I|}\Bigl| \Bigl\{ x\in I: \, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda \Bigr\} \Bigr| \end{align*} since the supremum defining $\mathbb{B}$ is taken over a larger space. Observe now that \begin{multline*} \frac{1}{|I|}\Bigl| \Bigl\{ x\in I: \, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda \Bigr\} \Bigr| = \\ \frac{1}{2|I_-|}\Bigl| \Bigl\{ x\in I_- : \, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda \Bigr\} \Bigr| + \\ \frac{1}{2|I_+|}\Bigl| \Bigl\{ x\in I_+ : \, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda \Bigr\} \Bigr| \\ =\frac{1}{2|I_-|}\Bigl| \Bigl\{ x\in I_- : \, \sum_{J \in \mathcal{D}(I_-)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda - \alpha_I t \Bigr\} \Bigr| + \\ \frac{1}{2|I_+|}\Bigl| \Bigl\{ x\in I_+ : \, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda - \alpha_I t \Bigr\} \Bigr| \\ =\frac{1}{2|I_-|}\Bigl| \Bigl\{ x\in I_- : \, \sum_{J \in \mathcal{D}(I_-)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda' \Bigr\} \Bigr| +\\ \frac{1}{2|I_+|}\Bigl| \Bigl\{ x\in I_+ : \, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(x) > \lambda' \Bigr\} \Bigr| \end{multline*} and thus the claim follows. \end{proof} Also, we trivially see that $\mathbb{B}$ must satisfy the following ``obstacle'' condition: \begin{equation}\label{Bellman.Obstacle} \mathbb{B}(t,A,\lambda) = 1 \quad \text{whenever } \lambda < 0. \end{equation} As we described in the beginning of the section, the function $\mathbb{B}$ is a minimizer in the space of positive functions which satisfy these properties. The following proposition makes this precise: \begin{proposition} \label{Bellman.BellmanDoesIt} Suppose a continuous function $F$ satisfies inequality \eqref{Bellman.MainInequality.eq} together with the obstacle condition \eqref{Bellman.Obstacle}, then we must have \[ \mathbb{B}(t,A,\lambda) \leq F(t,A,\lambda). \] \end{proposition} \begin{proof} Let $f \geq 0$ be an integrable function on an interval $I$ and let $\{\alpha_J\}_{J \in \mathcal{D}(I)}$ be a Carleson sequence with constant at most $1$, then for all fixed $\lambda$ we have (by \eqref{Bellman.MainInequality.eq}) \begin{align*} F(\langle f \rangle_I, A, \lambda) &= F\biggl( \frac{\langle f \rangle_{I_-} + \langle f \rangle_{I_+}}{2}, \frac{A_-+A_+}{2} + \alpha_I, \lambda \biggr) \\ &\geq \frac{1}{2}\Bigl( F(\langle f \rangle_{I_-}, A_-, \lambda - \alpha_I \langle f \rangle_I) + F(\langle f \rangle_{I_+}, A_+, \lambda - \alpha_I \langle f \rangle_I) \Bigr), \end{align*} where $A = \frac{1}{|I|}\sum_{J \subseteq I} \alpha_J |J|$ and $A_{\pm}$ is defined analogously for $I_-$ and $I_+$. If we iterate this inequality we obtain \begin{align*} F(\langle f \rangle_I, A, \lambda) &\geq \frac{1}{2^N} \sum_{J \subset I,\, |J| = 2^{-N}|I|} F(\langle f \rangle_J, A_J, \lambda - \sum_{k=1}^N \alpha_{J^{(k)}} \langle f \rangle_{J^(k)} \mathbbm{1}_{J^{(k)}}(c_J) ), \end{align*} where $A_J = \frac{1}{|J|} \sum_{P \subseteq J} \alpha_P |P|$. If we assume a priori that the Carleson sequence $\alpha$ is finite then we can let $N \to \infty$ and obtain \begin{align*} F(\langle f \rangle_I, A, \lambda) &\geq \frac{1}{|I|} \int_I F(f(x), A(x), \lambda - \mathcal{A}f(x) ) \, dx \\ &\geq \frac{1}{|I|} \int_{\{x \in I : \lambda - \mathcal{A}f(x) < 0\}} 1 \, dx &&\text{by } \eqref{Bellman.Obstacle}\\ &= \frac{1}{|I|}|\{x\in I: \mathcal{A}f(x) > \lambda \}|. \end{align*} Here $A(x)$ is almost everywhere-defined as the limit of $A(J)$ as $J \to x$, this is easily seen to exist almost everywhere by the Lebesgue differentiation theorem. Letting the number of non-zero elements of $\{\alpha_J \}_{J \in \mathcal{D}(I)}$ tend to infinity and then taking the supremum in the definition of $\mathbb{B}$ we obtain \[ F(\langle f \rangle_I, A, \lambda) \geq \mathbb{B}(\langle f \rangle_I, A, \lambda). \] \end{proof} \begin{remark} Note that we don't know yet if the function $\mathbb{B}$ is continuous, thus finding a minimizer in the space of continuous functions might not give us the true Bellman function. It turns out, however, that assuming continuity (actually $C^1$ smoothness) we are able to find a positive function satisfying \eqref{Bellman.MainInequality.eq} and \eqref{Bellman.Obstacle} which moreover is best possible without the a priori assumption of smoothness. We show this in the last section. \end{remark} We have therefore seen that finding any positive continuous function $F$ satisfying \eqref{Bellman.MainInequality.eq} and \eqref{Bellman.Obstacle} will give us an upper bound for $\mathbb{B}$. In the next section we find such a function. \section{Finding the Bellman function candidate} Our goal now is to find the smallest continuous function $F$ satisfying \eqref{Bellman.MainInequality.eq} and \eqref{Bellman.Obstacle}. As we remarked after Proposition \ref{Bellman.BellmanDoesIt}, we will assume a priori that $F$ is $C^1$. Moreover, we will restrict the minimization space even more by requiring $F$ to have the same kind of homogeneity that the true $\mathbb{B}$ must have, i.e.: \[ \mathbb{B}(\eta t, A, \eta \lambda) = \mathbb{B}(t,A,\lambda) \quad \forall \eta >0, \lambda >0. \] This in principle might make our candidate for Bellman function larger than the one we could find without requiring such homogeneity. However, the optimal Bellman function satisfies this identity, so requiring $F$ to also satisfy it will not prevent us from finding it. Assuming smoothness we can write the Main Inequality \eqref{Bellman.MainInequality.eq} as a concavity condition, together with a monotonicity property along certain characteristics. More precisely, if $F$ is a smooth positive function, then \eqref{Bellman.MainInequality.eq} together with \eqref{Bellman.Obstacle} and the above homogeneity is equivalent to the following conditions: \begin{enumerate} \item $F$ is nonnegative, and concave in the first two variables. \item $F(t,A,\lambda)$ is increasing in the direction $(0,1,t)$. \item $F(st,A,s\lambda) = F(t,A,\lambda)$ for all $s > 0$. \item $F(t,A,\lambda) = 1$ whenever $\lambda <0$ \end{enumerate} Indeed, if we let $\alpha = 0$ in \eqref{Bellman.MainInequality.eq} we see that $\mathbb{B}$ is concave in the variables $(t,A)$. If we set $A_1 = A_2 = A$ and $t_1 = t_2 = t$ then we see, by varying $\alpha$, that $\mathbb{B}(t,A,\lambda)$ is increasing in the direction $(0,1,t)$. This shows that any smooth $F$ satisfying \eqref{Bellman.MainInequality.eq} and \eqref{Bellman.Obstacle}, and which is also homogeneous in the above sense, must also satisfy properties (1) through (4). Moreover, if $F$ is any smooth function satisfying properties (1) through (4), then it also must satisfy the main inequality \eqref{Bellman.MainInequality.eq} and the obstacle condition \eqref{Bellman.Obstacle}. To see this observe that using property (1) we obtain \eqref{Bellman.MainInequality.eq} but with $\alpha = 0$, now property (2) allows us to insert an $\alpha$ as in the hypotheses for the main inequality since it describes the path along which $F$ is increasing. The homogeneity and obstacle conditions are exactly (3) and (4) respectively, so this proves the equivalence. Using the homogeneity property, we can reduce to finding $M:(0,\infty) \times [0,1] \to [0,\infty)$ such that if \[ F(x,y,z) = \begin{cases} M(x/z,y) & \text{if } z > 0\\ 1 & \text{if } z \leq 0, \end{cases} \] then $F$ satisfies (1) through (4). These properties, when translated to the function $M$, become: \begin{enumerate} \item $M$ is concave. \item $M_y - x^2 M_x \geq 0$. \item $M(x,y) \to 1$ when $x \to \infty$. \end{enumerate} The second of these properties tells us that $M$ is increasing along the characteristics \[ \begin{cases} \dot{x}(t) &= -x^2 \\ \dot{y}(t) &= 1. \end{cases} \] Observe that these characteristics foliate $[0,\infty) \times[0,1]$. Also, if we move backwards in time along a characteristic which starts at $(x_0,1)$ with $x_0 \geq 1$, then this characteristic is above the curve $y = \frac{1}{x}$ and furthermore the characteristic tends to $(\infty, y_f)$ for some $0< y_f<1$. Using the fact that $M(x,y) \to 1$ as $x \to \infty$ and that we should decrease if we move backwards along these characteristics, we must have \[ M(x,y) \geq 1 \quad \text{whenever } y \geq \frac{1}{x}. \] However, we may assume (if our goal is to find the true Bellman function) that $M \leq 1$ since the true Bellman function $\mathbb{B}$ obviously cannot be larger than $1$, so we will actually impose \[ M(x,y) = 1 \quad \text{whenever } y \geq \frac{1}{x}. \] Observe that $\mathbb{B}(0,0,1)$ is $0$ and consider the straight line joining the point $(0,0)$ with $(x_1,y_1)$, where $x_1y_1 = 1$. Observe also that the pointwise minimum of any two positive continuous functions satisfying \eqref{Bellman.MainInequality.eq} and \eqref{Bellman.Obstacle} will give us a smaller function which also satisfies these properties. We know that the function $M$ should be $1$ at $(x_1,y_1)$ and that, along this line, $M$ should be concave. The smallest concave curve joining these two points is obviously a straight line, so if defining $M$ in this way produces a smooth concave function satisfying the monotonicity property (2) then the optimal $M$ should be such a function. Joining the point $(0,0)$ with the points $(x_1,y_1) \in [0,\infty) \times [0,1]$ satisfying $x_1y_1 = 1$ covers everything in the subdomain $0 \leq y \leq \min(x,x^{-1})$, so let us define $M$ here by \[ M(x,y) = \sqrt{xy}. \] This function is linear along straight lines joining $(0,0)$ with the boundary curve $xy = 1$ and is $1$ at this boundary. It is furthermore concave and satisfies the monotonicity property (2), so if we knew that $\mathbb{B}$ is continuous then $\mathbb{B}(x,y,1)$ must be defined as above in this subdomain. We are therefore left with defining $M$ in the upper triangle $\Omega_T = \{0 \leq x \leq y \leq 1\}$. Inspired by the linear behavior of $M$ in the first domain, we make the ansatz that $M$ is actually $1$-homogeneous in the whole domain. Let $f(x) = M(x,1)$ for $0 \leq x \leq 1$, then if $M$ is $1$-homogeneous we should have \[ M(x,y) = y f(x/y). \] If we want condition $(2)$ to hold then we should have \[ f(x/y)-(x/y)f'(x/y) - x^2f'(x/y) \geq 0. \] We expect this to be an equality on the boundary, which is when $y = 1$, so we will assume that \[ f(x) - f'(x)(x+x^2) = 0. \] This ordinary differential equation has the solutions \[ f(x) = C\frac{x}{1+x}, \] and we should furthermore have $f(1) = M(1,1) = 1$. So $ C= 2$ and therefore \[ f(x) = \frac{2x}{x+1} \implies M(x,y) = \frac{2xy}{x+y} \] whenever $1 \geq y \geq x \geq 0$. One easily verifies that $M$ satisfies all the requirements in this subdomain, so we just have to show that the whole function $M$ is concave, but this immediately follows from the fact that $M$ is concave in each subdomain and that $M$ is $C^1$ (as can be easily seen). This gives us that \[ M(x,y) = \begin{cases} \frac{2xy}{x+y} &\text{if } 0 \leq x \leq y \leq 1 \\ \sqrt{xy} & \text{if } 0 \leq y \leq \min(x,x^{-1}), \end{cases} \] which using the homogeneity gives us the full function of Theorem \ref{MainTheorem}. \section{Optimality} In this section we show that the function found in the previous section is actually the exact Bellman function. We first we need a simple technical lemma which will allow us to deduce that $\mathbb{B}(\cdot,\cdot,1)$ must be superlinear along lines joining $(0,0,1)$ to $(x,1,1)$. \begin{lemma}\label{DydadicConcavity.Lemma} Let $f:[0,1] \to [0,\infty)$ be a function which satisfies \begin{equation} \label{DyadicConcavity.eq} f\Bigl( \frac{x+y}{2} \Bigr) \geq \frac{1}{2}f(x) + \frac{1}{2}f(y) \end{equation} for all $0 \leq x \leq y \leq 1$. Then we must have \[ f(x) \geq f(1)x \] for all $x \in [0,1].$ \end{lemma} \begin{proof} We can assume without loss of generality that $x \in (0,1)$ and that $f(1) = 1$. Using \eqref{DyadicConcavity.eq} we have \begin{equation}\label{DyadicConcavity.eq2} f(x_0 + \lambda(1-x_0)) \geq \lambda \end{equation} for all dyadic rationals $\lambda \in [0,1]$, i.e.: numbers of the form $\lambda = k2^{-N}$ for $0 \leq k \leq 2^N$. For every $N \in \mathbb{N}$ let $k_N$ be the unique integer in $0 \leq k \leq 2^N x$ which satisfies \[ \Bigl| x - \frac{k}{2^N} \Bigr| < \frac{1}{2^N} \] (this exists because the sequence $k \mapsto k2^{-N}$ is an arithmetic sequence of step $2^{-N}$). Observe that then, if we define \[ x_N := \frac{2^Nx - k_N}{2^N - k_N} = \frac{x-\frac{k_N}{2^N}}{1-\frac{k_N}{2^N}}, \] we must have $0 \leq x_N \leq \frac{1}{2^N(1-x)}$, so in particular $x_N \to 0$ as $N \to \infty$. But then \[ \lambda := \frac{k_N}{2^N} = \frac{x-x_N}{1-x_N} \] is a dyadic rational and plugging it into \eqref{DyadicConcavity.eq2}, with $x_N$ playing the role of $x_0$, yields \[ f(x) \geq \frac{x-x_N}{1-x_N}, \] so letting $N \to \infty$ completes the proof. \end{proof} Using this lemma, together with the Main Inequality \eqref{Bellman.MainInequality.eq} we immediately have the following corollary: \begin{corollary} \label{EasyObstacle} We have the following identity: \[ \mathbb{B}(x,y,1) = M(x,y) \] for all $x,y$ in the subdomain $0 \leq y \leq \min(x,x^{-1})$. \end{corollary} \begin{proof} We showed in the previous section that $\mathbb{B}(x,y,1) \leq M(x,y)$ for all $(x,y) \in \Omega'$. To show the reverse inequality notice that the Main Inequality \eqref{Bellman.MainInequality.eq} together with Lemma \ref{DydadicConcavity.Lemma} imply \begin{equation}\label{EasyObstacle.eq} \mathbb{B}(x,y,1) \geq \lambda \mathbb{B}\Bigl(\frac{x}{\lambda},\frac{y}{\lambda},1 \Bigr). \end{equation} We would be done if we can show that $\mathbb{B}(x,y,1) = 1$ whenever $xy = 1$. Indeed, then we can just use equation \eqref{EasyObstacle.eq} with $\lambda = \sqrt{xy}$. Fix $(x,y) \in \Omega'$ with $xy = 1$ and consider the function \[ f_n = \frac{2^nx}{2^n -1}\mathbbm{1}_{[0,1-2^{-n})}. \] If $I$ is the interval $[0,1)$ then obviously $\langle f_n \rangle_I = x$. Consider also the Carleson sequence $\{\alpha_J\}_{J \in \mathcal{D}(I)}$ defined by \[ \alpha_J = \begin{cases} \frac{y}{1-2^{-n}} & \text{if } J = [2^{-n}(k-1),2^{-n}k) \text{ and } k \in \{1, \dots, 2^n-1\}\\ 0 & \text{otherwise.} \end{cases} \] Then we have \[ \frac{1}{|I|}\sum_{J \in \mathcal{D}(I)} \alpha_J |J| = y\sum_{k=1}^{2^n-1} \frac{2^{-n}}{1-2^{-n}} = y. \] Also, \[ \mathcal{A}f_n(t) = \begin{cases} \Bigl(\frac{2^n}{2^n-1}\Bigr)^2 & \text{if } 0 \leq t < 1- 2^{-n} \\ 0 & \text{otherwise,} \end{cases} \] hence \[ \mathbb{B}(x,y,1) \geq 1-2^{-n} \] for all $n \geq 1$. Letting $n \to \infty$ yields the claim. \end{proof} \begin{remark} Observe that using the constant function $f(t) = x\mathbbm{1}_I(t)$ and the one-term Carleson sequence which is $y$ on $I$ and $0$ everywhere else, one obtains that $\mathcal{A}f = xy\mathbbm{1}_I$, hence $\mathbb{B}(x,y,1) = 1$ for all $xy >1$. \end{remark} Using Lemma \ref{DydadicConcavity.Lemma} in the same way, we just have to show that $\mathbb{B}(x,1,1) = \frac{2x}{x+1}$ to prove that $\mathbb{B}(x,y,1) = M(x,y)$ in the rest of the domain, however this turns out to be harder. \begin{theorem}\label{MainTheorem2} Fix $x \in (0,1)$ and let $\epsilon > 0$. For any interval $I$ there exists a nonnegative function $f$ on $I$ with $\langle f \rangle_I = x$ and a Carleson sequence $\{\alpha_J\}_{J \in \mathcal{D}(I)}$ with Carleson constant at most one and verifying \[ \frac{1}{|I|}\sum_{J \in \mathcal{D}(I)} \alpha_J |J| = 1 \] such that \[ \frac{1}{|I|}\Bigl| I \cap \Bigl\{ \sum_{J \in \mathbb{D}(I)} \alpha_J \langle f\rangle_J \mathbbm{1}_J > 1 \Bigr\} \Bigr| = \frac{2x}{x+1} + O(\epsilon). \] \end{theorem} To prove this we will use the Main Inequality \eqref{Bellman.MainInequality.eq} iteratively to give a decomposition of $f$ consisting of constant functions on certain dyadic intervals, this also gives us the construction of the sequence $\{\alpha_J\}_{J \in \mathcal{D}(I)}$. The basic idea is to, starting with a point $(x,1)$ in $\Omega'$, use \eqref{Bellman.MainInequality.eq} to split this point into another point $(x_+,1)$ on the boundary and some point $(x_-,A_-)$. The point $(x_-,A_-)$ is then absorbed back into the initial point and we apply the same procedure to the point $(x_+,1)$ until we get to a point past the obstacle $xy \geq 1$ (where extremizers consist of constant functions together with one-term Carleson sequences as in the Remark after Corollary \ref{EasyObstacle}). In order to illustrate the idea we will first prove the lower bound for $\mathbb{B}$ without explicitly constructing the example. The way in which we prove the lower bound will make the construction more intuitive. \begin{theorem}\label{Optimality} The Bellman function $\mathbb{B}$ satisfies \[ \mathbb{B}(x,1,1) = \frac{2x}{x+1} \] for all $x \in [0,1]$. \end{theorem} \begin{proof} Let $E(x,y) = \mathbb{B}(x,y,1)$, then using the Main Inequality \eqref{Bellman.MainInequality.eq} we see that we have the following behavior: \[ E(t,1) \geq \frac{1}{2}E\Bigl( \frac{t_1}{1-\alpha t}, A_1 \Bigr) + \frac{1}{2} E\Bigl(\frac{t_2}{1-\alpha t}, A_2 \Bigr) \] whenever $t = \frac{t_1+t_2}{2}$ and $1 = \frac{A_1 + A_2}{2}+\alpha$. Letting $\epsilon > 0$, $x = t$ and $A_2 = 1$ we get \[ E(x,1) \geq \frac{1}{2}\Bigl( E\Bigl( x - 2\epsilon, 1 - \frac{2\epsilon}{x} \Bigr) + E\Bigl( x_+, 1 \Bigr)\Bigr), \] where \[ x_+ = x \frac{1+\epsilon}{1-\epsilon} + 2\epsilon. \] Since $\mathbb{B}$ is superlinear in the first two variables and $\mathbb{B}(0,0,1) = 0$, we must have \[ E\Bigl( x - 2\epsilon, 1 - \frac{2\epsilon}{x} \Bigr) \geq \Bigl( 1-\frac{2\epsilon}{x} \Bigr)E(x,1) \] so putting everything together we obtain \begin{equation}\label{Bellman.Optimality.It2} E(x,1) \geq \frac{x}{x+2\epsilon} E(x_+,1). \end{equation} If we define inductively $x_{n+1} = x_n \frac{1+\epsilon}{1-\epsilon} + 2\epsilon$ and $x_0 = x$, then we easily see that \[ x_n = \delta^n \Bigl( \frac{1}{1-\epsilon} + x \Bigr) - \frac{1}{1-\epsilon}, \] where $\delta = \frac{1+\epsilon}{1-\epsilon}$. We want to stop the iteration once $x_n \geq 1$, and this happens when \[ \delta^n \geq \frac{2-\epsilon}{1+x(1-\epsilon)}, \] let $N = N(\epsilon,x)$ be the smallest integer for which the above inequality does not hold. Then iterating \eqref{Bellman.Optimality.It2} $N$ times we get (since $E(1,1) = 1$) \[ E(x,1) \geq \prod_{j=0}^N \frac{x_j}{x_j+2\epsilon}, \] it just suffices to give a lower bound for the right hand side. To this end observe that \begin{align*} \prod_{j=0}^N \frac{x_j}{x_j+2\epsilon} &\geq \exp\Bigl(-\sum_{j=0}^N \log\Bigl(1 + \frac{2\epsilon}{x_j}\Bigr)\Bigr) \\ &\geq \exp \Bigl( - \sum_{j=0}^N \frac{2\epsilon}{x_j} \Bigr) \\ &= \exp\Bigl( -2\epsilon \sum_{j=0}^N \frac{1}{x_j} \Bigr). \end{align*} Let us estimate $-2\epsilon \sum_{j=0}^N \frac{1}{x_j}$. Using the explicit formula for $x_n$ we have \begin{align*} -2\epsilon \sum_{j=0}^N \frac{1}{x_j} &= -2\epsilon \sum_{j=0}^N \frac{1}{\delta^j \bigl( \frac{1}{1-\epsilon} + x \bigr)- \frac{1}{1-\epsilon}} \\ &= -2\epsilon\sum_{j=0}^N \biggl( \frac{1}{\delta^j \bigl( \frac{1}{1-\epsilon} + x \bigr)- \frac{1}{1-\epsilon}} - \frac{1}{\delta^j \bigl( 1 + x \bigr)- 1}\biggr) + \sum_{j=0}^N \frac{2\epsilon}{1-\delta^j(1+x)}. \end{align*} The first term tends to $0$ as $\epsilon \to 0$ and the second is a Riemann sum, indeed (recalling the definition of $N = N(x,\epsilon)$: \begin{align*} \sum_{j=0}^N \frac{2\epsilon}{1-\delta^j(1+x)} &= (1-\epsilon)\sum_{j=0}^N \frac{\delta^j\frac{2\epsilon}{1-\epsilon}}{\delta^j(1- \delta^j (1+x))} \\ &= (1-\epsilon) \sum_{j=0}^N f(\delta^j)(\delta^{j+1}-\delta^j) \\ &= \int_1^{\frac{2}{1+x}} f(y) \, dy + O(\epsilon), \end{align*} as $\epsilon \to 0$ and where \[ f(y) = \frac{1}{y(1-y(x+1))}. \] It is easy to see that \[ \int_1^{\frac{2}{1+x}} \frac{1}{y(1-y(x+1))} \, dy = \log\Bigl( \frac{2x}{x+1} \Bigr), \] which completes the proof of the lower bound. \end{proof} Let us now use these ideas to construct the example. There are two basic steps in the iteration: first we split the point $(x,1)$ into $(x_-,A_-)$ and $(x_+,1)$, then we absorb $(x_-,A_-)$ into $(x,1)$ and obtain a lower bound for $E(x,1)$ in terms of $E(x_+,1)$, we then iterate this until $x_+ >1$, where we stop because we know that $E(x_+,1)$ must be $1$ there. These two steps are imposing a certain self-similarity on $f$ and the Carleson sequence $\alpha$ in terms of $(f_+,\alpha_+)$. The following Lemma, which is based on the ideas from \cite{Vasyunin2009}, makes this precise. \begin{lemma} \label{TechnicalLemma} Fix an interval $I$ and let $g_+$ be a nonnegative function on $I_+$. Suppose also that $\alpha^+$ is a Carleson sequence adapted to $I_+$ with constant at most $1$ and such that \[ \frac{1}{|I_+|} \sum_{J \in \mathcal{D}(I_+)} \alpha^+_J |J| = 1. \] If $\langle g_+ \rangle_{I_+} = x \frac{1+\epsilon}{1-\epsilon} + 2\epsilon$ for some $x \in (0,1)$ and a sufficiently small $\epsilon > 0$, then we can construct a function $f$ on $I$ and a Carleson sequence $\alpha$ adapted to $I$ with constant at most $1$ such that $\langle f \rangle_I = x$, \begin{equation} \label{TechnicalLemma.eq.1} \frac{1}{|I|} \sum_{J \in \mathcal{D}(I)} \alpha_J |J| = 1 \end{equation} and \begin{equation} \label{TechnicalLemma.eq.2} \frac{1}{|I|} \Bigl| I \cap \Bigl\{ \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J > 1 \Bigr\} \Bigr| \geq \Bigl(\frac{x}{x+2\epsilon}\Bigr) \frac{1}{|I_+|} \Bigl| I_+ \cap \Bigl\{ \sum_{J \in \mathcal{D}(I_+)} \alpha_J^+ \langle g_+ \rangle_J \mathbbm{1}_J > 1 \Bigr\} \Bigr|. \end{equation} \end{lemma} \begin{proof} We will assume without loss of generality that $I = [0,1)$, also denote $\alpha = \frac{\epsilon}{x}$. Define $\alpha_J$ to be $\alpha$ if $J = I$ and $\alpha^+_J$ for $J \in \mathcal{D}(I_+)$. Define $f$ to be $(1-\epsilon)g_+$ on $I_+$ and denote $g_- = (1-\epsilon)^{-1}f \mathbbm{1}_{I_-}$, then \begin{multline*} \frac{1}{|I|}\Bigl| y \in I:\, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(y) > 1 \Bigr| = \\ \frac{1}{2|I_-|}\Bigl| y \in I_-:\, \sum_{J \in \mathcal{D}(I_-)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(y) > 1 - \epsilon \Bigr|\\ +\frac{1}{2|I_+|}\Bigl| y \in I_+:\, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle f \rangle_J \mathbbm{1}_J(y) > 1 - \epsilon \Bigr| \\ =\frac{1}{2|I_-|}\Bigl| y \in I_-:\, \sum_{J \in \mathcal{D}(I_-)} \alpha_J \langle g_- \rangle_J \mathbbm{1}_J(y) > 1 \Bigr| \\ +\frac{1}{2|I_+|}\Bigl| y \in I_+:\, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle g_+ \rangle_J \mathbbm{1}_J(y) > 1 \Bigr|. \end{multline*} Let $I_j = [e_{j}, e_{j+1})$, where $e_j = \frac{1}{2} - 2^{-j}$, and suppose that $\alpha_{\widehat{I_j}} = 0$ for $j \geq 1$ and $\alpha_{I_-} = 0$, then \begin{equation}\label{TechnicalLemma.Proof.eq.1} \frac{1}{2|I_-|}\Bigl| y \in I_-:\, \sum_{J \in \mathcal{D}(I_-)} \alpha_J \langle g_- \rangle_J \mathbbm{1}_J(y) > 1 \Bigr| = \frac{1}{2}\sum_{j= 1}^\infty 2^{-j}\frac{1}{|I_j|}\Bigl| y\in I_j:\, \mathcal{A}(g_-\mathbbm{1}_j)(y) > 1 \Bigr|. \end{equation} Let $\theta = 1-2\alpha$ and write \[ \theta = \sum_{j=1}^\infty 2^{-j}b_j \] for some binary sequence $\{b_j\}_{j \in \mathbb{N}}$ (i.e.: write $\theta$ in binary). For a given interval $J$ let $S_Jf$ be the scaled version of $f$ adapted to $J$, i.e.: if $J = [a,b)$ then \[ S_J f(x) = f\Bigl( \frac{x-a}{b-a} \Bigr). \] Abusing notation, let us also denote by $S_J \alpha$ the scaled version of the Carleson sequence $\alpha$ to the dyadic subinterval $J$ of $I$, then we have \begin{multline*} \frac{1}{|J|}\Bigl|\Bigl\{y \in J:\, \sum_{K \in \mathcal{D}(J)} (S_J \alpha)_K \langle S_J f \rangle_K \mathbbm{1}_K(y) > 1 \Bigr\}\Bigr|= \\ \frac{1}{|I|}\Bigl| \Bigl\{ y\in I:\, \sum_{K \in \mathcal{D}(I)} \alpha_K \langle f \rangle_K \mathbbm{1}_K(y) > 1 \Bigr\} \Bigr|. \end{multline*} Suppose that $(1-\epsilon)f$, when restricted to $I_j$, agrees with $S_{I_j}f$ for all $j \geq 1$ such that $b_j = 1$ and is $0$ otherwise. Suppose furthermore that the Carleson sequence $\alpha$ also satisfies the same similarity, i.e.: if we scale to $I$ the restriction of $\alpha$ to $I_j$ we obtain $\alpha$ again. If we denote by $\Xi$ the left-hand side in \eqref{TechnicalLemma.eq.2} then we could use \eqref{TechnicalLemma.Proof.eq.1} to obtain \[ \Xi = \frac{1}{2}\sum_{j=1}^\infty 2^{-j} b_j \Xi + \frac{1}{2|I_+|}\Bigl| y \in I_+:\, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle g_+ \rangle_J \mathbbm{1}_J(y) > 1 \Bigr|, \] hence \begin{align*} \Xi &= \Bigl(\frac{1}{1+2\alpha}\Bigr)\frac{1}{|I_+|}\Bigl| y \in I_+:\, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle g_+ \rangle_J \mathbbm{1}_J(y) > 1 \Bigr| \\ &= \Bigl(\frac{x}{x+2\epsilon}\Bigr)\frac{1}{|I_+|}\Bigl| y \in I_+:\, \sum_{J \in \mathcal{D}(I_+)} \alpha_J \langle g_+ \rangle_J \mathbbm{1}_J(y) > 1 \Bigr|, \end{align*} which is what we wanted. Note also that we could use the same method to compute the average of $f$ and it yields precisely the right amount: $x$. Therefore we just have to show that we can find a function $f$ and a Carleson sequence $\alpha$ satisfying these self-similarity conditions. Let us start with $f$: define the operator $T$ by \[ Tf = (1-\epsilon)\sum_{j=1}^\infty b_j \mathbbm{1}_{I_j} S_{I_j} f + (1-\epsilon)\mathbbm{1}_{I_+} g_+. \] We need to show that $T$ has a fixed point in $L^1(I)$; we will do this following the steps of the proof of the Banach fixed point theorem. Let $f_0 = (1-\epsilon)g_+ \mathbbm{1}_{I_+}$ and define inductively \[ f_{n+1} = Tf_n. \] We should show that $f_n$ is a Cauchy sequence in $L^1(I)$, but observe that \begin{align*} \|f_{n+1}-f_n\|_{L^{1}(I)} &= (1-\epsilon)\int_{I_-} \Bigl| \sum_{j=1}^\infty b_j \mathbbm{1}_{I_j} S_{I_j}(f_n) - \sum_{j=1}^\infty b_j \mathbbm{1}_{I_j} S_{I_j}(f_{n-1}) \Bigr| \\ &= (1-\epsilon)\sum_{j=1}^\infty b_j \int_{I_j} |S_{I_j}(f_n) - S_{I_j}(f_{n-1})| \\ &= (1-\epsilon)\sum_{j=1}^\infty b_j|I_j| \int_{I}|f_n-f_{n-1}| \\ &= (1-\epsilon)\int_I |f_n-f_{n-1}| \sum_{j=1}^\infty b_j 2^{-j-1} \\ &= \frac{(1-\epsilon)(1-2\alpha)}{2} \int_{I} |f_n-f_{n-1}|. \end{align*} The constant $\xi := \frac{(1-\epsilon)(1-2\alpha)}{2}$ is strictly less than $1$ and by induction we have \[ \|f_{n+1}-f_n\|_{L^1(I)} \lesssim \xi^{n}, \] hence the sequence is Cauchy. This finishes the proof of existence for $f$ since we can just define $f$ to be the limit in $L^1$ of the sequence $f_n$ defined above. To show the existence of the Carleson sequence we can follow the same steps as above, but now we don't have to deal with convergence issues. Indeed, start with a sequence as in the beginning of the proof and define inductively the $(n+1)$-th sequence $\alpha^{n+1}$ by inserting the entire dyadic tree of $\alpha^n$ at each $I_j$. At each step we are only changing the value of the sequence at deeper and deeper levels, so we can just define $\alpha_K$ as the the value of $\alpha_K^n$, where $n$ is the first integer at which the sequence $\alpha_K^n$ stabilizes. \end{proof} We are now ready to prove Theorem \ref{MainTheorem2}, we will use the same ideas and notation as in the proof of Theorem \ref{Optimality}. Given $\epsilon$, $I$ and $x \in (0,1)$ let $N$ be the smallest integer such that \[ \delta^n \geq \frac{2-\epsilon}{1+x(1-\epsilon)}. \] Let $f_1$ be the constant function $x$ on $I_+$ and let $\alpha^1$ be the one-term Carleson sequence which is $1$ at $I_+$. Now define the function $f_{n+1}$ and Carleson sequence $\alpha^{n+1}$ inductively by applying Lemma \ref{TechnicalLemma} to the function $g_+ := S_{I_+}(f_n)$ and the Carleson sequence $S_{I_+}(\alpha^n)$; let $f = f_N$ and $\alpha = \alpha^N$. Then, as in the proof of Theorem \ref{Optimality}, we have \[ \frac{1}{|I|}\Bigl| \Bigl\{ y\in I:\, \sum_{J \in \mathcal{D}(I)} \alpha_J \langle f \rangle_J \mathbbm{1}_J >1 \Bigr\} \Bigr| \geq \exp\Bigl( \sum_{j=0}^N \frac{2\epsilon}{1-\delta^j(1+x)} \Bigr), \] which we showed to be \[ \frac{2x}{x+1} + O(\epsilon), \] and this is what we wanted to prove.
\section{Introduction} The motivation for Predictable Feature Analysis (PFA) comes from typical reinforcement-learning settings, where an autonomous agent is placed in an environment and aims to achieve some goal. While many common scenarios are discrete (board-game-like) with rather few states, we consider more natural scenarios, where the input is a continuous signal over time and of high dimension like vision or some other sensory input.\footnote[1]{ Since it comes from a technical setup, the signal would still be discrete. However, we would not regard its discreteness as states but would conceptually treat it as a continuous signal.} PFA is intended as a tool to help the agent make sense of this vast amount of incoming data. Our approach is to look for information in the signal that helps to understand and manipulate the environment in the desired way. To achieve its goal, the agent must be able to plan its actions and thus needs to understand, how the environment behaves -- it needs a model that is capable of predicting the outcomes of possible actions. It has been frequently proposed that predictable features are crucial to obtain such a model, see \cite{BialekNemenmanEtAl-2001} for a review. In contrast to most common approaches from control theory, we attempt to perform the modeling without putting previously known, problem-specific information (usually a representation of the environment in form of differential equations and system theoretic setups) into the model, but look for a truly unsupervised, self-organized approach. Slow Feature Analysis (SFA) is an algorithm that has most characteristics we are looking for (and as such also served as the name-giving pattern for PFA). It is an algorithm that has proven valuable in several fields and problems concerning signal- and data analysis. The idea is that a drastic, yet reasonable dimensionality reduction can be obtained by focusing on slowly varying sub-signals, the so-called “slow features”. These are considered most relevant, because slowness usually indicates invariance and invariant problem representations are crucial for typical data-analysis and recognition tasks, such as regression and classification. Many of these tasks have proven to become much more feasible on the reduced signal after SFA has been applied. For instance, tasks like the self-organization of complex-cell receptive fields, the recognition of whole objects invariant to spatial transformations, the self-organization of place-cells, extraction of driving forces, or nonlinear blind source separation were successfully performed on the basis of SFA (see \cite{DahneWilbertEtAl-2010a, FranziusWilbertEtAl-2011, FranziusSprekelerEtAl-2007d, Wiskott-2003b, BlaschkeZitoEtAl-2007}). PFA extracts sub-signals from the input using the same methods like SFA does, but instead of the slowest features, it selects those that are best predictable by a certain prediction model. In section \ref{sec:criteria} we give the criteria a model must meet to be suitable for this purpose. While there are also model-independent notions of predictability like in the information bottleneck approach \cite{CreutzigGlobersonEtAl-2009, CreutzigSprekeler-2008}, focusing on concrete models has the advantage that if PFA finds predictable sub-signals, one is directly able to actually perform the prediction, since the appropriate model is given. The arising optimization problem turned out to be significantly harder than that of SFA, because it is a nested problem: The features extracted must be optimized for predictability, but judging their predictability is an optimization problem by itself. Optimizing these problems in turns usually converges to sub-optimal solutions that depend on the starting point. In this work we discuss the details of the PFA-problem and present a tractable algorithm, thus setting up the basis for future PFA-related work. There have been other approaches to use notions of predictability. For instance \cite{gepperth:simultaneous} considers scenarios involving embodied agents. They let two sensors predict each other in order to retrieve representation-invariant information. \cite{CreutzigSprekeler-2008} combines notions of predictability with SFA to better understand principles of sensory coding strategies. There also exists an ICA-based approach to predictability-driven dimensionality reduction, see \cite{Hyvarinen01}. ForeCA (Forecastable Component Analysis), an independently developed method, is based on the same paradigm as PFA, but proposes a model-independent approach \cite{goerg13}. A further difference is that PFA (optionally) searches for well predictable Systems, while ForeCA selects best predictable single components. In future work, we are going to compare PFA- and ForeCA-results to get a better understanding for strengths and weaknesses of the two approaches. Finally, there has also been a previous version of our PFA approach \cite{RichthoferWeghenkelWiskott-2012c} \section{Extracting predictable features} \label{sec:extraction} Given an input-signal $\mathbf{x}(t)$ with $n$ components, our goal is to extract a certain number ($r$) of well predictable output-components, referred to as “predictable features”. Since our approach is inspired by SFA, we start with a summary of that algorithm. \subsection{Recall SFA} \label{sec:sfa} In the SFA-setting, the optimized property is slow variation. Extraction is in principle performed by linear transformation and projection.\footnote[2]{In a strict sense, the transformation is affine because it clears the signal's mean. Additionally one can count the non-linear expansion as part of the extraction.} The parameters of these mappings are optimized over a finite training-phase $\trph$ consisting of equidistant time points. To make the method more powerful, a non-linear expansion $\mathbf{h}$ can be applied to the signal -- usually using monomials of low degree. In order to avoid the trivial constant solution, the output is constrained to have unit variance and zero mean. Additionally, the output-components must be pairwise uncorrelated. This way the repeated occurrence of the same component is avoided. Mean is defined using $\av{s(t)}_t~\coloneq~\frac{1}{\abs{\trph}} \sum_{t\in \trph} s(t)$ (average of a signal over the training phase). To fulfill the constraints, the expanded signal is sphered over the training-phase, i.e. its mean is shifted to zero and the covariance-matrix is normalized to the identity matrix: \begin{align} \label{sphering} \tilde{\mathbf{z}}(t) \quad &\coloneq \quad \mathbf{h}(\mathbf{x}(t)) - \av{\mathbf{h}(\mathbf{x}(t))}_t &&\text{(make mean-free)}\\ \mathbf{z}(t) \quad &\coloneq \quad \mathbf{S} \tilde{\mathbf{z}}(t) \qquad \text{with} \quad \mathbf{S} \coloneq \av{\tilde{\mathbf{z}} \tilde{\mathbf{z}}^T}^{-\frac{1}{2}} &&\text{(normalize covariance)} \end{align} Summing up all SFA-constraints, the following optimization problem is derived: \begin{align} \label{sfa} \text{For} \; i \in \{1, \ldots, r\} \notag \\ \begin{split} \optmin{\mathbf{a}_i \in \mathbb{R}^{n}} & \mathbf{a}_i^T \av{\dot{\mathbf{z}}\dot{\mathbf{z}}^T} \mathbf{a}_i\\ \subjectto & \mathbf{a}_i^T \av{\mathbf{z}} \hphantom{\mathbf{a}_i \mathbf{a}_j \mathbf{z}^T} \, = \quad 0 \quad \hphantom{\forall \; j < i} \quad \text{(zero mean)}\\ & \mathbf{a}_i^T \av{\mathbf{z} \mathbf{z}^T} \mathbf{a}_i \hphantom{\mathbf{a}_j} = \quad 1 \quad \hphantom{\forall \; j < i} \quad \text{(unit variance)}\\ & \mathbf{a}_i^T \av{\mathbf{z} \mathbf{z}^T} \mathbf{a}_j \hphantom{\mathbf{a}_i} = \quad 0 \quad \forall \; j < i \quad \text{(pairwise uncorrelated)} \end{split} \end{align} To describe the algorithm, we first define the extraction matrix $\mathbf{A}_r \coloneq \left(\mathbf{a}_1, \ldots, \mathbf{a}_r \right) \in \mathbb{R}^{n \times r}$ and the reduced identity $\mathbf{I}_r \in \mathbb{R}^{n \times r}$ consisting of the first $r$ euclidean unit vectors as columns. Now please note that because of the sphering, it holds that $\av{\mathbf{z}} = 0$ and $\av{\mathbf{z} \mathbf{z}^T} = \id$, thus having the constraints equal to \begin{equation} \exists \mathbf{A} \in \orth(n) \colon \qquad \mathbf{A}_r \quad = \quad \mathbf{A}\mathbf{I}_r \end{equation} where $\orth(n) \subset \mathbb{R}^{n \times n}$ denotes the space of orthogonal transformations, i.e. $\mathbf{A}\mathbf{A}^T = \mathbf{I}$. Choosing $\mathbf{a}_i$ as the eigenvectors of $\av{\dot{\mathbf{z}}\dot{\mathbf{z}}^T}$, corresponding to the eigenvalues in descending order, yields an $\mathbf{A}_r$ that solves \eqref{sfa} globally. We denote the extracted signal with $\mathbf{m} \coloneq \mathbf{A}_r^T \mathbf{z}$. \cite{WiskottBerkesEtAl-2011} describes this procedure in detail. \subsection{Modeling the PFA-problem} \label{sec:pfa} In order to measure predictability, we focus on a certain prediction-model. Because it is simple and very popular, we use \textbf{linear, auto-regressive prediction} as our default model -- it is successfully used in many fields for modeling time-related problems. A signal is regarded well predictable, if each value can be approximated by a linear combination of some ($p$) recent values. Expressing this formally, we face the problem of finding vectors $\mathbf{a}$ and $\mathbf{b}$ such that \begin{align} \mathbf{a}^T \mathbf{z}(t) \quad \appr^! \;& \quad b_{1} \mathbf{a}^T \mathbf{z}(t-1) + \ldots + b_{p} \mathbf{a}^T \mathbf{z}(t-p) \label{pfa-criterionARScalar}\\ = \;& \quad \mathbf{a}^T \hist_{\mathbf{z}, p}(t) \; \mathbf{b} \end{align} with $\hist$ defined as the signal's history over the recent $p$ time-steps: \begin{equation} \label{history} \hist_{\mathbf{z}, p, \Delta}(t) \quad \coloneq \quad \sum_{i=1}^p \quad \mathbf{z}(t-i\Delta) \mathbf{e}_i^T \quad \text{with} \quad \mathbf{e}_i \in \mathbb{R}^p, \quad \left(\mathbf{e}_1, \ldots, \mathbf{e}_p \right) = \mathbf{I}_{p, p}. \end{equation} Here $\mathbf{I}_{p, p}$ denotes the $p$-dimensional identity, thus $\mathbf{e}_i$ denotes the $i$-th $p$-dimensional Euclidean unit vector. $\Delta$ defaults to $1$: $\hist_{\mathbf{z}, p} \coloneq \hist_{\mathbf{z}, p, 1}$. \begin{figure}[ht] \centering \subfloat \psset{xunit=1.5,yunit=1.0} \begin{pspicture} (0.25, 1)(7, 3) \sffamily \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01, arrowsize=0.2, arrowlength=2, arrowinset=0]{->}(0.5, 2)(2, 2) \psframe(2, 1.5)(3.5, 2.5) \rput[c](2.75, 2){\large{PFA}} \rput[t](1.05, 1.8){$\mathbf{z}(t)$} \psline[doubleline=true, doublesep=0.02, doublecolor=orange, linewidth=0.01, arrowsize=0.1, arrowlength=2, arrowinset=0]{->}(3.5, 2)(6, 2) \rput[tl](3.7, 1.8){$\mathbf{a}^T \mathbf{z}(t)$} \pnode(4.5, 2){A} \pnode(5.25, 2){B} \pnode(6, 2){C} \cnode*(6.75, 2){0.06}{D} \psline[linewidth=0.01, linestyle=dashed](4.5, 1.85)(4.5, 2.15) \psline[linewidth=0.01, linestyle=dashed](5.25, 1.85)(5.25, 2.15) \psline[linewidth=0.01, linestyle=dashed](6, 1.85)(6, 2.15) \ncarc[linewidth=0.015, arcangle=70, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{C}{D} \ncarc[linewidth=0.015, arcangle=70, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{B}{D} \ncarc[linewidth=0.015, arcangle=70, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{A}{D} \end{pspicture} } \caption{Illustration of PFA} \label{fig:PFAIllustration} \end{figure} Like in SFA, we optimize the parameters over the training phase $\trph$ and also adopt the SFA constraints to avoid trivial or repeated solutions. The first steps of PFA are indeed equal to those in SFA, i.e. we also allow for a non-linear expansion and also start with a sphering-step. As far as possible, we use the notation that was introduced in \ref{sec:sfa}. The common way to extend \eqref{pfa-criterionARScalar} to multiple dimensions can be written as \begin{equation} \label{pfa-criterionARDiagMatrix} \mathbf{m}(t) \quad \appr^! \quad \mathbf{B}_{1} \mathbf{m}(t-1) + \ldots + \mathbf{B}_{p} \mathbf{m}(t-p) \quad \text{with}\quad \mathbf{B}_i \in \mathbb{R}^{n \times n} \text{, diagonal} \end{equation} In this form, it does not fulfill all criteria from section \ref{sec:criteria} (it is not orthogonal agnostic for $n > 1$). Nevertheless, we mention strategies to solve \eqref{pfa-criterionARDiagMatrix} in the appendix, section \ref{sec:extractIsolated}. Here we proceed by refining it to be suitable for PFA: \begin{equation} \label{pfa-criterionARGeneralMatrix} \mathbf{m}(t) \quad \appr^! \quad \mathbf{B}_{1} \mathbf{m}(t-1) + \ldots + \mathbf{B}_{p} \mathbf{m}(t-p) \quad \text{with}\quad \mathbf{B}_i \in \mathbb{R}^{n \times n} \hphantom{\text{, diagonal}}. \end{equation} The difference to the first formulation is that each extracted component's prediction may depend on all other extracted components. Note that \eqref{pfa-criterionARDiagMatrix} and \eqref{pfa-criterionARGeneralMatrix} are equal for $n = 1$. A massive advantage of model \eqref{pfa-criterionARGeneralMatrix} is that we can initially fit it to our data in full dimension and search for the best-fitted components afterwards. For \eqref{pfa-criterionARDiagMatrix}, this would not be possible, because the fitting-quality of each component is not invariant under the transformation used for extraction. We formalize the need for such a model-property in section \ref{sec:criteria}. To formalize fitting, we briefly introduce the notion of a general prediction model and of the fitting-error. We speak of a general prediction-model $\mathbf{g}$ as \begin{equation} \label{generalPredictionModel} \mathbf{g} \in \mathcal{G}: \qquad \mathbf{z}(t) \quad \appr^! \quad \mathbf{g}(\hist_{\mathbf{z}, p, \Delta}(t)) \end{equation} where $\mathcal{G}$ is the model-class, i.e. the set of possible realizations of $\mathbf{g}$. We measure the prediction error in an average least squares sense by \begin{align} &\err(\mathbf{g}, \mathbf{z}) \quad \coloneq \quad \pfaerrr{\mathbf{g}}{ \mathbf{z}}{ \hist_{\mathbf{z}, p, \Delta}} \label{totalPredictionError} \end{align} Now fitting a prediction-model on a given sample in a least squares sense can be expressed as \begin{equation} \label{fitting_g0} \optmin{\mathbf{g} \in \mathcal{G}} \pfaerr{\mathbf{g}}{\mathbf{z}} \end{equation} By $\mathbf{g}_{\mathbf{z}}^*$, we denote the global solution of \eqref{fitting_g0} and define the following shortcut-notation: \begin{align} &\err(\mathbf{z}) \quad \coloneq \quad \pfaerr{\mathbf{g}^*_\mathbf{z}}{ \mathbf{z}} \label{shortcutError} \end{align} To formalize \eqref{pfa-criterionARGeneralMatrix} as a prediction model in this notation, we combine the coefficient-matrices $\mathbf{B}_{i}$ to a single broad matrix and define $\mathbf{g}_{\mathbf{B}}$ and $\mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}}$: \begin{align} \mathbf{B} \hphantom{\mathbf{g}_{\mathbf{B}}(\hist_{\mathbf{z}, p}(t))\mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}}} &\coloneq \qquad \left( \mathbf{B}_{1}, \ldots, \mathbf{B}_{p} \right) \quad \in \; \mathbb{R}^{n \times np} \label{B} \\ \mathbf{g}_{\mathbf{B}}(\hist_{\mathbf{z}, p}(t)) \hphantom{\mathbf{B} \mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}}} &\coloneq \qquad \mathbf{B} \mvec(\hist_{\mathbf{z}, p}(t)) \label{gB} \\ \mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}} \hphantom{\mathbf{g}_{\mathbf{B}}(\hist_{\mathbf{z}, p}(t))\mathbf{B}} &\coloneq \qquad \lbrace \; \mathbf{g}_{\mathbf{B}} \colon \quad \mathbf{B} \in \; \mathbb{R}^{n \times np} \; \rbrace \end{align} By analytic optimization, we obtain the following regression formula to fit $\mathbf{g}_{\mathbf{B}} \in \mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}}$ to $\mathbf{m}~=~\mathbf{A}_r^T \mathbf{z}$: \begin{equation} \label{B-ReducedDimRankTransformSolution} \mathbf{B}_{\mathbf{z}}(\mathbf{A}_r) \quad = \quad \mathbf{A}^T_r \av{\mathbf{z}\mathbf{\zeta}^T}\underline{\mathbf{A}_r} \left(\underline{\mathbf{A}_r}^T\av{\mathbf{\zeta}\mathbf{\zeta}^T} \underline{\mathbf{A}_r} \right)^{-1} \end{equation} Here we used $\zeta(t) \coloneq \mvec(\hist_{\mathbf{z}, p}(t))$ and the following shortcut notation defined for any matrix $\mathbf{M}$: \begin{equation} \label{multiA} \underline{\mathbf{M}} \quad \coloneq \quad \mathbf{I}_{p, p} \otimes \mathbf{M} \quad = \qquad \underbrace{\!\!\!\!\!\!\left( \begin{smallmatrix} \mathbf{M}& & \mathbf{0} \\ & \ddots & \\ \mathbf{0} & & \mathbf{M} \end{smallmatrix} \right)\!\!\!\!\!\!}_{\text{$p$ times $\mathbf{M}$}} \end{equation} If $r=n$, $\mathbf{A}=\mathbf{I}$ and thus $\mathbf{A}_r=\mathbf{I}$, we write $\mathbf{W} \coloneq \mathbf{B}_{\mathbf{z}}(\mathbf{I}) = \av{\mathbf{z}\mathbf{\zeta}^T} \av{\mathbf{\zeta}\mathbf{\zeta}^T}^{-1}$. (See \ref{sec:notation} for an overview of all notation in this document.) It sometimes happens that $\av{\mathbf{\zeta}\mathbf{\zeta}^T}$ is not (cleanly) invertible due to some very small or even zero eigenvalues. We regard it best practice to project away the eigenspaces corresponding to eigenvalues below a critical threshold. The intuition behind this is that spaces corresponding to (almost-)zero eigenvalues indicate redundancies in the signal and should not be used for prediction anyway. To perform this, first compute an eigenvalue decomposition on $\av{\mathbf{\zeta}\mathbf{\zeta}^T}$. Replace eigenvalues below the threshold by $0$ and replace the other ones by their multiplicative inverse. After that undo the decomposition and use the resulting matrix as a proxy for $\av{\mathbf{\zeta}\mathbf{\zeta}^T}^{-1}$. For $r=n$ and $\mathbf{A} \in \orth(n)$, we have $\mathbf{B}_{\mathbf{z}}(\mathbf{A}) = \mathbf{A}^T\mathbf{W}\underline{\mathbf{A}}$. Since $\mathbf{z}$ is sphered, we can state the following compact notation of the PFA-problem: \begin{equation} \label{pfaAutoRegressiveWhitened} \optmin{\mathbf{A} \in \orth(n)} \; \err(\mathbf{A}_r^T \mathbf{z}) \end{equation} Inserting our default model, we have $\err(\mathbf{A}_r^T \mathbf{z}) = \pfaerrob{\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r) \underline{\mathbf{A}_r}^T \zeta}{ \mathbf{A}_r^T \mathbf{z}}$. However, because \eqref{B-ReducedDimRankTransformSolution} is an involved term, mainly due to the projection under an inversion symbol, \eqref{pfaAutoRegressiveWhitened} appears to be intractable by every method known to us\footnote[3]{Not counting evolutionary and other inherent local optimization approaches, since we aim for the global solution. Experiments showed us that locally optimal solutions are usually still of high error and of low relevance for the model.}. Instead of solving it directly, we propose the following tractable relaxation: \begin{equation} \label{pfaNonAutoRegressiveWhitened} \optmin{\mathbf{A} \in \orth(n)} \; \pfaerrob{\mathbf{I}_r^T\mathbf{B}_{\mathbf{z}}(\mathbf{A}) \underline{\mathbf{A}}^T \zeta}{ \mathbf{A}_r^T \mathbf{z}} \quad = \quad \bigav{\; \norm{\mathbf{A}_r^T(\mathbf{z}-\mathbf{W}\mathbf{\zeta})}^2 \; } \end{equation} Informally speaking, problem \eqref{pfaNonAutoRegressiveWhitened} asks for components that are optimally predictable, if the prediction may be based on the entire input signal, rather than just on the extracted components themselves. From now on we denote a global optimum of \eqref{pfaAutoRegressiveWhitened} with $\mathbf{A}_r^*$ and of \eqref{pfaNonAutoRegressiveWhitened} with $\mathbf{A}_r^{(0)}$. \begin{figure}[ht] \centering \subfloat \psset{xunit=1.5,yunit=1.0} \begin{pspicture} (0.25, 1)(7, 3 \sffamily \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01](0.5, 1.75)(2, 1.75) \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01, linestyle=dashed](2, 1.75)(3.5, 1.75) \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01, arrowsize=0.2, arrowlength=2, arrowinset=0]{->}(3.5, 1.75)(6, 1.75) \psframe(2, 1.25)(3.5, 3) \rput[c](2.75, 2.5){\large{PFA}} \rput[b](1.25, 2){$\mathbf{z}(t)$} \psline[doubleline=true, doublesep=0.02, doublecolor=orange, linewidth=0.01, arrowsize=0.1, arrowlength=2, arrowinset=0]{->}(3.5, 2.5)(6, 2.5) \rput[bl](3.7, 2.6){$\mathbf{m}(t) = \mathbf{A}_r^T \mathbf{z}(t)$} \pnode(4.5, 1.75){A} \pnode(5.25, 1.75){B} \pnode(6, 1.75){C} \cnode*(6.75, 2.5){0.06}{D} \psline[linewidth=0.01, linestyle=dashed](4.5, 1.5)(4.5, 2.6) \psline[linewidth=0.01, linestyle=dashed](5.25, 1.5)(5.25, 2.6) \psline[linewidth=0.01, linestyle=dashed](6, 1.5)(6, 2.6) \ncarc[linewidth=0.015, arcangle=10, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{C}{D} \ncarc[linewidth=0.015, arcangle=-7, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{B}{D} \ncarc[linewidth=0.015, arcangle=-10, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{A}{D} \end{pspicture} } \caption{Illustration of relaxation \eqref{pfaNonAutoRegressiveWhitened}} \end{figure} To solve \eqref{pfaNonAutoRegressiveWhitened} globally, we write it as \begin{equation} \label{pfaNonAutoRegressiveWhitenedRewrite} \optmin{\mathbf{A} \in \orth(n)} \; \tr \left( \mathbf{A}_r^T \bigav{\left(\mathbf{z} - \mathbf{W}\mathbf{\zeta}\right) \left(\mathbf{z} - \mathbf{W}\mathbf{\zeta}\right)^T} \mathbf{A}_r \right) \end{equation} and choose $\mathbf{A}$ such that it diagonalizes $\bigav{\left(\mathbf{z} - \mathbf{W}\mathbf{\zeta}\right) \left(\mathbf{z} - \mathbf{W}\mathbf{\zeta}\right)^T}$ and sorts the $r$ smallest eigenvalues to the upper left. This method can be described as performing PCA on the residuals of the least squares fit. By some calculus, one can show formal equivalence of this approach to the method proposed in \cite{boxTiao1977}. In section \ref{sec:proofOfRelaxationGap} we prove that if $\err((\mathbf{A}_r^*)^T \mathbf{z}) = 0$, then $\mathbf{A}_r^{(0)}$ is also a global solution of \eqref{pfaAutoRegressiveWhitened}.\footnote[4]{Note that if $\mathbf{A}_r^{(0)}$ is used as solution for \eqref{pfaAutoRegressiveWhitened}, the prediction model must be refitted to the reduced signal to get optimal prediction. For this, calculate $\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r^{(0)})$ as defined in \eqref{B-ReducedDimRankTransformSolution}.} More precisely speaking, the relaxation gap of \eqref{pfaNonAutoRegressiveWhitened} depends on $\err((\mathbf{A}_r^*)^T \mathbf{z})$ in a continuous manner and is zero, if that error is zero. If the optimal sub-signal has a significant prediction error, the solution obtained as $\mathbf{A}_r^{(0)}$ usually suffers from overfitting and is sub-optimal for \eqref{pfaAutoRegressiveWhitened}. In the following, we offer a heuristic method to overcome this overfitting. \subsection{Avoiding overfitting} \label{sec:overfitting} To reduce overfitting, we propose the heuristics that signals well predictable in terms of \eqref{pfaAutoRegressiveWhitened} yield a lower error-propagation to subsequent predictions than signals that are well predictable in terms of \eqref{pfaNonAutoRegressiveWhitened} but not in terms of \eqref{pfaAutoRegressiveWhitened}. We ground this on the intuition that the prediction of the latter ones is partly based on noisy data -- thus subsequent predictions inherit a higher error. To formalize this idea we define \begin{equation} \label{V} \mathbf{V} \quad \coloneq \quad \av{\mathbf{\zeta}(t+1)\mathbf{\zeta}^T(t)}_t \av{\mathbf{\zeta}\mathbf{\zeta}^T}^{-1} \end{equation} and can thus perform iterated prediction as follows: \begin{equation} \label{zVi} \mathbf{z}(t) \quad \approx \quad \mathbf{W}\mathbf{V}^i \mathbf{\zeta}(t-i) \end{equation} \begin{figure}[ht] \centering \captionsetup[subfloat]{labelformat=empty} \subfloat \psset{xunit=1.5, yunit=1.0} \begin{pspicture} (0.25, 1)(7.1, 3 \sffamily \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01](0.5, 1.75)(2, 1.75) \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01, linestyle=dashed](2, 1.75)(3.5, 1.75) \psline[doubleline=true, doublesep=0.19, doublecolor=yellow, linewidth=0.01, arrowsize=0.2, arrowlength=2, arrowinset=0]{->}(3.5, 1.75)(6, 1.75) \psframe(2, 1.25)(3.5, 3) \rput[c](2.75, 2.5){\large{PFA}} \rput[b](1.25, 2){$\mathbf{z}(t)$} \psline[doubleline=true, doublesep=0.02, doublecolor=orange, linewidth=0.01, arrowsize=0.1, arrowlength=2, arrowinset=0]{->}(3.5, 2.5)(6, 2.5) \rput[bl](3.7, 2.6){$\mathbf{m}(t) = \mathbf{A}^T \mathbf{z}(t)$} \pnode(4.5, 1.75){A} \pnode(5.25, 1.75){B} \pnode(6, 1.75){C} \cnode*(6.75, 1.75){0.06}{D} \cnode*(7.5, 1.75){0.06}{E} \cnode*(8.25, 1.75){0.06}{F} \cnode*(8.25, 2.5){0.06}{G} \psline[linewidth=0.01, linestyle=dashed](4.5, 1.5)(4.5, 2.6) \psline[linewidth=0.01, linestyle=dashed](5.25, 1.5)(5.25, 2.6) \psline[linewidth=0.01, linestyle=dashed](6, 1.5)(6, 2.6) \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{C}{D} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{B}{D} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{A}{D} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{D}{E} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{C}{E} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{B}{E} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{E}{F} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{D}{F} \ncarc[linewidth=0.015, arcangle=40, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{C}{F} \ncdiag[linewidth=0.015, angleA=90, angleB=-90, arrowsize=0.1, arrowlength=1, arrowinset=0.2]{->}{F}{G} \end{pspicture}} \caption{Illustration of iterated prediction} \end{figure} Based on this, we propose the following optimization problem: \begin{equation} \label{pfaNonAutoRegressiveIterated} \optmin{\mathbf{A} \in \orth(n)} \; \sum_{i=0}^k \quad \bigav{\; \norm{\mathbf{A}_r^T(\mathbf{z}-\mathbf{W}\mathbf{V}^i\mathbf{\zeta}(t-i)}^2 \; }_t \end{equation} We can solve it globally in a rather similar way like \eqref{pfaNonAutoRegressiveWhitened}. To do so, we write it as \begin{equation} \label{pfaNonAutoRegressiveIteratedRewrite} \optmin{\mathbf{A} \in \orth(n)} \; \tr \Big( \mathbf{A}_r^T \sum_{i=0}^k \bigav{\left(\mathbf{z} - \mathbf{W}\mathbf{V}^i\mathbf{\zeta}(t-i)\right) \left(\mathbf{z} - \mathbf{W}\mathbf{V}^i\mathbf{\zeta}(t-i)\right)^T}_t \mathbf{A}_r \Big) \end{equation} and then solve it by diagonalizing $\sum_{i=0}^k \bigav{\left(\mathbf{z} - \mathbf{W}\mathbf{V}^i\mathbf{\zeta}(t-i)\right) \left(\mathbf{z} - \mathbf{W}\mathbf{V}^i\mathbf{\zeta}(t-i)\right)^T}_t$ and sorting the lowest $r$ eigenvalues to the upper left. We denote the global solution of \eqref{pfaNonAutoRegressiveIterated} by $\mathbf{A}^{(k)}_r$. How to optimally choose $k$ for a certain problem is currently an open question, but we know from experiments that up to some value, increasing $k$ improves $\err((\mathbf{A}^{(k)}_r)^T \mathbf{z})$. Beyond that value, increasing $k$ lowers the quality again. Our intuition is that the critical value is related to the maximal time distance, over which the signal holds any auto-correlation -- investigating this formally will be subject of future work. To give some impression of the technique and as a proof of concept, we demonstrate it on a synthetic example. We know that basic trigonometric functions are losslessly predictable with our default model for $p=2$. Because of the theorem in section \ref{sec:proofOfRelaxationGap}, it would make no sense to work with a losslessly predictable signal. So we add some white noise $\eta(t)$ to it and define an example signal $\mathbf{x}(t) \coloneq (\sin(0.1 t)+0.7 \eta(t), \sin(0.2 t)+\eta(t), \sin(0.4 t)+5.3 \eta(t))^T$. We train the algorithm with $\Delta = 1$ once on $1000$ samples and once on $2000$ samples, always extracting two components, i.e. $r=2, p=2$. As a lower bound for any error not involving overfitting, we evaluate \eqref{pfaNonAutoRegressiveWhitened}. For our example we get a lower bound of $\approx 1.206$ for $1000$ samples and $\approx 1.22$ for $2000$ samples. These are plotted as red horizontal lines in the following. To measure the amount of overfitting, we add many dimensions of white random data to our signal and mix everything up by a random, orthogonal transformation. While $\mathbf{x}$ has always the same noise-seed, the added data is generated with different noise in every run. The following results are averaged over about 150 runs, and plot the prediction error against the noise-dimension:\footnote[5]{Note that the vertical axis ranges from $1.15$ onward to provide a better focus.} \begin{figure}[H] \centering \parbox{6cm}{ \includegraphics[width=7cm]{PFAIteration1-5Noise1000c.eps} \caption{$k=0,\ldots, 4$ from left to right$;$ $k=4$ dashed$;$ 1000 samples} \label{fig:k0to4at1000}} \quad \begin{minipage}{6cm} \includegraphics[width=7cm]{PFAIteration5-15Noise1000c.eps} \caption{$k=4,\ldots, 14$ from right to left$;$ $k=4$ dashed$;$ 1000 samples} \label{fig:k4to14at1000} \end{minipage} \begin{minipage}{6cm} \includegraphics[width=7cm]{PFAIteration1-5Noise2000c.eps} \caption{$k=0,\ldots, 4$ from left to right$;$ $k=4$ dashed$;$ 2000 samples} \label{fig:k0to4at2000} \end{minipage} \quad \begin{minipage}{6cm} \includegraphics[width=7cm]{PFAIteration5-15Noise2000c.eps} \caption{$k=4,\ldots, 14$ from right to left$;$ $k=4$ dashed$;$ 2000 samples} \label{fig:k4to14at2000} \end{minipage} \end{figure} Obviously $k=4$ yields the best results in this example. The result for $k=0$ with no additional noise can be considered to be the optimal solution. Solutions below the red line indicate overfitting conform with \eqref{pfaAutoRegressiveWhitened}. This kind of overfitting can only be reduced by using larger samples. That it decreases for higher noise-dimension indicates that the best solution is not found any more. So we conclude that in our example, the algorithm is robust for about $50$ dimensional noise if $1000$ samples are used and for about $100$-dimensional noise, if $2000$ samples are used. Future work will include more detailed research about these relationships. \pagebreak \section{Criteria for suitable prediction models} \label{sec:criteria} In this section we discuss what properties of a prediction model are crucial to make the procedure described in \ref{sec:pfa} feasible. \begin{definition}[Orthogonal agnosticity criterion] We say that a prediction-model $\mathcal{G}$ is \textbf{orthogonal-agnostic} on $\Omega_t$, if for every $\mathbf{A} \in \orth(n), \mathbf{g} \in \mathcal{G}$: \begin{equation} \label{orthogonalAgnosticity} \err(\mathbf{z}) \quad = \quad \err(\mathbf{A}^T \mathbf{z}) \end{equation} \end{definition} \eqref{orthogonalAgnosticity} means that the model fits equally well to any orthogonal transformation of the data. In section \ref{sec:proofOfRelaxationGap} we will need a more restrictive variant of this criterion that additionally considers projections of the data to subspaces: \begin{definition}[Projective orthogonal agnosticity criterion]\label{sec:projOrthAgn} We say that a prediction-model $\mathcal{G}$ is \textbf{projective orthogonal-agnostic} on $\Omega_t$, if for every $\mathbf{A} \in \orth(n)$, $r \leq n$ the following holds: \begin{equation} \label{projectiveOrthogonalAgnosticity} \pfaerrr{\mathbf{I}_r^T \mathbf{g}_{\mathbf{A}^T \mathbf{z}}^*}{\mathbf{A}_r^T \mathbf{z}}{ \mathbf{A}^T \hist_{\mathbf{z}, p}} \quad = \quad \pfaerrr{\mathbf{A}_r^T \mathbf{g}_{\mathbf{z}}^*}{\mathbf{A}_r^T \mathbf{z}}{\hist_{\mathbf{z}, p}}. \end{equation} \end{definition} Note that for $r = n$, \eqref{projectiveOrthogonalAgnosticity} simplifies to \eqref{orthogonalAgnosticity} and projective orthogonal agnosticity becomes equivalent to ordinary orthogonal agnosticity (since the Frobenius-norm is invariant under orthogonal transformations). An even stronger and very intuitive criterion is the following: \begin{definition}[Commuting with orthogonal transformations] We say that a prediction-model $\mathcal{G}$ \textbf{commutes with orthogonal transformations}, if for every $\mathbf{A} \in \orth(n)$ the following holds: \begin{equation} \label{commutingWithOrthogonal} \mathbf{g}_{\mathbf{A}^T \mathbf{z}}^* \quad = \quad \mathbf{A}^T \mathbf{g}_{\mathbf{z}}^*. \end{equation} \end{definition} It is rather obvious that this criterion implies projective and ordinary orthogonal agnosticity. To assure projective orthogonal agnosticity, it is a straightforward procedure to construct models such that they commute with orthogonal transformations. \begin{definition}[Information consistency criterion] We say that a prediction-model $\mathcal{G}$ is \textbf{information-consistent} on $\Omega_t$, if for every $\mathbf{A} \in \orth(n)$, $r \leq n$ the following holds: \begin{equation} \label{informationConsistency} \pfaerrrb{\mathbf{g}_{\mathbf{A}_r^T \mathbf{z}}^*}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}_r^T \hist_{\mathbf{z}, p}} \quad \geq \quad \pfaerrrb{\mathbf{A}_r^T \mathbf{g}_{\mathbf{z}}^*}{\mathbf{A}_r^T \mathbf{z}}{\hist_{\mathbf{z}, p}} \end{equation} \end{definition} An information-consistent model always benefits from more data rather than getting confused by it. Note that for $r = n$, \eqref{informationConsistency} follows from orthogonal agnosticity. \begin{thm} Model $\mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}}$ is projective orthogonal agnostic and information consistent. \end{thm} \begin{proof} Projective orthogonal agnosticity follows, because the model commutes with orthogonal transformations, as $\mathbf{B}_{\mathbf{z}}(\mathbf{A}) = \mathbf{A}^T\mathbf{B}_{\mathbf{z}}(\mathbf{I})\underline{\mathbf{A}}$. To show that $\mathcal{G}_{\eqref{pfa-criterionARGeneralMatrix}}$ is information consistent, we need the solution of the following optimization problem: \begin{equation} \label{proj_fitting_g} \optmin{\mathbf{B} \in \mathbb{R}^{n \times np}} \pfaerrr{\mathbf{I}_r^T\mathbf{g}_{\mathbf{B}}}{\mathbf{A}_r^T\mathbf{z}}{\mathbf{A}^T \hist_{\mathbf{z}, p}} \end{equation} Analytically we find a (not unique) solution to be $\mathbf{B}_{\mathbf{z}}(\mathbf{A})$. Note that $\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r) \in \mathbb{R}^{r \times rp}$. We extend each $(r \times r)$-block at the bottom and right with zeroes to get $(n \times n)$-blocks and overall get an $(n \times np)$-matrix $\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r)^{(n \times np)}$, which can be seen as a candidate to solve \eqref{proj_fitting_g}. Thus we have \begin{equation} \label{W-rxrpTonxnp} \mathbf{B}_{\mathbf{z}}(\mathbf{A}_r) \quad = \quad \mathbf{I}_r^T \mathbf{B}_{\mathbf{z}}(\mathbf{A}_r)^{(n \times np)} \underline{\mathbf{I}_r} \end{equation} which implies that \begin{equation} \label{W-rxrpTonxnp-implication} \pfaerrr{\mathbf{g}_{\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r)}}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}_r^T \hist_{\mathbf{z}, p}} \quad = \quad \pfaerrr{\mathbf{I}_r^T \mathbf{g}_{\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r)^{(n \times np)}}}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}^T \hist_{\mathbf{z}, p}} \end{equation} Since we know that $\mathbf{B}_{\mathbf{z}}(\mathbf{A})$ is an optimal solution of \eqref{proj_fitting_g}, we have \begin{equation} \label{W-rxrpTonxnp-implication2} \pfaerrr{\mathbf{I}_r^T \mathbf{g}_{\mathbf{B}_{\mathbf{z}}(\mathbf{A}_r)^{(n \times np)}}}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}^T \hist_{\mathbf{z}, p}} \quad \geq \quad \pfaerrr{\mathbf{I}_r^T \mathbf{g}_{\mathbf{B}_{\mathbf{z}}(\mathbf{A})}}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}^T \hist_{\mathbf{z}, p}} \end{equation} and thus \begin{equation} \label{W-rxrpTonxnp-implication3} \pfaerrrb{\mathbf{g}_{\mathbf{A}_r^T \mathbf{z}}^*}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}_r^T \hist_{\mathbf{z}, p}} \quad \geq \quad \pfaerrrb{\mathbf{I}_r^T \mathbf{g}_{\mathbf{A}^T \mathbf{z}}^*}{\mathbf{A}_r^T \mathbf{z}}{\mathbf{A}^T \hist_{\mathbf{z}, p}} \end{equation} With the projective orthogonal agnosticity criterion \eqref{projectiveOrthogonalAgnosticity}, we can transform the right side of \eqref{W-rxrpTonxnp-implication3} into the right side of \eqref{informationConsistency} and have formally shown information consistency for model \eqref{pfa-criterionARGeneralMatrix}. \end{proof} \subsection{General formulation of PFA} \label{sec:generalPFA} The criteria in section \ref{sec:criteria} assure that a problem analog to \eqref{pfaAutoRegressiveWhitened} can be relaxed to a tractable problem like \eqref{pfaNonAutoRegressiveWhitened} and that it can be solved like in the corresponding section. Additionally they assure that the theorem in section \ref{sec:proofOfRelaxationGap} holds and that the procedure from \ref{sec:overfitting} is applicable. For any projective orthogonal agnostic and information consistent prediction model, \eqref{pfaAutoRegressiveWhitened} can be relaxed to \begin{equation} \label{pfaGeneralNonAutoRegressiveWhitened} \optmin{\mathbf{A} \in \orth(n)} \; \bigav{\; \norm{\mathbf{A}_r^T(\mathbf{z}-\mathbf{g}_{\mathbf{z}}^*(\hist_{\mathbf{z}, p}))}^2 \; } \end{equation} which can be solved by diagonalizing $\bigav{\left(\mathbf{z} - \mathbf{g}_{\mathbf{z}}^*(\hist_{\mathbf{z}, p})\right) \left(\mathbf{z} - \mathbf{g}_{\mathbf{z}}^*(\hist_{\mathbf{z}, p})\right)^T}$ and sorting the smallest eigenvalues to the upper left. To extend this generalization to \eqref{pfaNonAutoRegressiveIterated}, a version of \eqref{V} that only uses $\mathbf{g}_{\mathbf{z}}^*$ is needed. However this construction is straight forward, but can generally not be written as a matrix like in \eqref{V}. One uses $\mathbf{g}_{\mathbf{z}}^*$ only for prediction of the first (i.e. the new) components of $\mathbf{\zeta}(t+1)$, while the other components can be copied from $\mathbf{\zeta}(t)$. \section{Relaxation Gap Theorem} \label{sec:proofOfRelaxationGap} \begin{thm} \label{sec:thm2} For any prediction model class $\mathcal{G}$ that is projective orthogonal agnostic and information consistent, the following holds: \begin{align} &\text{If} \hphantom{\text{thenad}} \exists \; r \leq n, \hphantom{r,} \!\!\! \mathbf{A}^* \in \orth(n) \colon \err({\mathbf{A}_r^*}^T \mathbf{z}) \hphantom{{\mathbf{A}_r^{(0)}}} \!\! = \quad 0 \label{thm2Prerequisites}\\ &\text{and} \hphantom{\text{Ifthe}} \nexists \; \tilde{r} > r, \hphantom{n,} \!\!\! \mathbf{A}^* \in \orth(n) \colon \err({\mathbf{A}_{\tilde{r}}^*}^T \mathbf{z}) \hphantom{{\mathbf{A}_r^{(0)}}} \vphantom{{\mathbf{A}_r^{(0)}}^T} \!\! = \quad 0 \label{thm2Prerequisites2}\\ &\text{then} \hphantom{\text{Ifad}} \hphantom{\nexists \; \tilde{r} > r, \hphantom{n,} \mathbf{A}^* \in \orth(n) \colon} \!\!\! \err({\mathbf{A}_r^{(0)}}^T \mathbf{z}) \hphantom{{\mathbf{A}_r^*}} \!\! = \quad 0\label{thm2} \end{align} \end{thm} Line \eqref{thm2Prerequisites2} has the purpose to ensure that the maximal $r$ holding \eqref{thm2Prerequisites} is used in line \eqref{thm2}. To prove the theorem, we need to setup some lemmas and the following definition: \begin{definition}[Space-partition preserving orthogonal transformations] \begin{equation} \orth(r, s) \quad \coloneq \quad \diag(\orth(r), \orth(s)) \end{equation} \end{definition} This means that every $\tilde{\mathbf{A}} \in \orth(r, s) \subset \orth(r+s)$ has the form $\left( \begin{smallmatrix} \mathbf{A}_{rr} & \mathbf{0} \\ \mathbf{0} & \mathbf{A}_{ss} \end{smallmatrix} \right)$ with $\mathbf{A}_{rr} \in \orth(r)$ and $\mathbf{A}_{ss} \in \orth(s)$. Now we can elegantly formulate the following lemma, which deals with the non-uniqueness of $\mathbf{A}^{(0)}$: \begin{lem} \label{sec:lem1} Let $\mathbf{A}^{(0)}$ and $\mathbf{B}^{(0)}$ be two different global solutions of \eqref{pfaGeneralNonAutoRegressiveWhitened} and assume that the best $r$ components are well defined (i.e. component $r+1$ has worse error than component $r$). \begin{equation} \exists \; \tilde{\mathbf{A}} \in \orth(r, n-r) \colon \quad \mathbf{A}^{(0)} = \mathbf{B}^{(0)} \tilde{\mathbf{A}} \end{equation} \end{lem} A problem would arise, if for instance the $r$th-worst component and the $(r+1)$th-worst component had equal error. In that case it would not be well defined, which signal space to extract and the lemma would not hold. \begin{proof}[Proof of Lemma \ref{sec:lem1}] As mentioned earlier, we can obtain one global solution by diagonalizing $\bigav{\left(\mathbf{z} - \mathbf{g}_{\mathbf{z}}^*(\mathbf{Z})\right) \left(\mathbf{z} - \mathbf{g}_{\mathbf{z}}^*(\mathbf{Z})\right)^T}$ and sorting the $r$ smallest eigenvalues to the upper left. Since lemma~\ref{sec:lem1} requires to have a unique choice of $r$ best components, every optimal solution must have the same set of eigenvalues in the upper left $(r \times r)$-sub-matrix. Thus we can create every solution by orthogonally transforming the eigenspace of the $r$ smallest eigenvalues in itself. In an analog way, the eigenspace of the $n-r$ largest eigenvalues may be transformed. The set of partition preserving orthogonal transformations $\orth(r, n-r)$ is exactly defined to consist of the transformations performing this. \end{proof} \begin{lem} \label{sec:lem2} $\forall \; r \leq n, \mathbf{A} \in \orth(n), \tilde{\mathbf{A}} \in \orth(r, n-r) \colon$ \begin{equation} \label{lem2.eq} \err(\; \mathbf{I}_r^T\mathbf{A}^T \mathbf{z} \;) \quad =\quad \err(\; \mathbf{I}_r^T(\mathbf{A}\tilde{\mathbf{A}})^T \mathbf{z} \;) \end{equation} \end{lem} Lemma \ref{sec:lem2} implies that solutions of \eqref{pfaAutoRegressiveWhitened} stay solutions, if transformed by any $\tilde{\mathbf{A}}~\in~\orth(r, n-r)$. \begin{proof}[Proof of Lemma \ref{sec:lem2}] First observe the following fact: \begin{equation} \label{lem2Proof.1} \forall \; \tilde{\mathbf{A}} \in \orth(r, n-r) \colon \quad \exists \; \mathbf{A}_{rr} \in \orth(r) \colon \; \mathbf{I}_r^T \tilde{\mathbf{A}}^T = \mathbf{A}_{rr}^T \mathbf{I}_r^T \end{equation} With \eqref{lem2Proof.1} and the orthogonal agnosticity criterion it is straight forward to transform the right side of \eqref{lem2.eq} into the left: \begin{equation} \vphantom{\eq_{\substack{\eqref{lem2Proof.1}\\ \hphantom{\text{orth.}}}}} \err(\mathbf{I}_r^T \tilde{\mathbf{A}}^T \mathbf{A}^T \mathbf{z}) \quad \eq_{\substack{\eqref{lem2Proof.1}\\ \hphantom{\text{orth.}}}} \quad \err(\mathbf{A}_{rr}^T \mathbf{I}_r^T \mathbf{A}^T \mathbf{z}) \quad \eq_{\substack{\text{orth.}\\ \text{agn.}}} \quad \err(\mathbf{I}_r^T \mathbf{A}^T \mathbf{z}) \end{equation} \end{proof} Now we are ready to assemble the proof of the relaxation gap theorem: \begin{proof}[Proof of the relaxation gap theorem] By condition \eqref{thm2Prerequisites}, we have \begin{equation} \err({\mathbf{A}_r^*}^T \mathbf{z}) \quad = \quad \pfaerrr{\mathbf{g}_{{\mathbf{A}_r^*}^T \mathbf{z}}^*}{{\mathbf{A}_r^*}^T \mathbf{z}}{{\mathbf{A}_r^*}^T \mathbf{Z}} \hphantom{\mathbf{I}_r^T ._{{.^*}} {.^*} } \!\!\! = \quad 0 \end{equation} By information consistency, it follows that \begin{equation} \hphantom{\err({\mathbf{A}_r^*}^T \mathbf{z}) \quad = \quad} \pfaerrr{\mathbf{I}_r^T \mathbf{g}_{{\mathbf{A}^*}^T \mathbf{z}}^*}{{\mathbf{A}_r^*}^T \mathbf{z}}{{\mathbf{A}^*}^T \mathbf{Z}} \hphantom{._{{._r^*}} {._r^*} } \!\!\! = \quad 0 \end{equation} So $\mathbf{A}^*$ is a common global optimum of \eqref{pfaAutoRegressiveWhitened} and \eqref{pfaGeneralNonAutoRegressiveWhitened}. Since \eqref{thm2Prerequisites2} ensures maximality of $r$, we can apply lemma \ref{sec:lem1} and get: \begin{equation} \exists \; \tilde{\mathbf{A}} \in \orth(r, n-r) \colon \quad \mathbf{A}^{(0)} = \, \mathbf{A}^* \tilde{\mathbf{A}} \end{equation} Finally by lemma \ref{sec:lem2} we conclude that $\mathbf{A}^{(0)}$ must also be an optimum of \eqref{pfaAutoRegressiveWhitened}. \end{proof} By continuity arguments, the implication of theorem \ref{sec:thm2} extends to signals with components of low error -- the lower the error is, the more precise we can find an optimum of \eqref{pfaAutoRegressiveWhitened} by solving \eqref{pfaGeneralNonAutoRegressiveWhitened}. Providing bounds for the steepness of this continuous relationship is still an open problem and may be subject of future work. \section{Future Work} An important aspect of our future work will be the application of PFA to real world problems. We plan to approach scenarios where SFA is known to produce good results, so we can compare PFA and SFA and get a clearer notion for the differences between the paradigms. On the other hand, we have new scenarios in mind, that specifically would benefit from predictable features. For instance we are working on applications related to robotic navigation and the lane-keeping problem of a simulated car. As mentioned in some sections of this document, another fork of our future work will be to extend the analytic understanding of the heuristic aspects of the algorithm. This way we also aim to improve our methods to avoid overfitting. \section*{Acknowledgments} This work is funded by a grant from the German Research Foundation (Deutsche Forschungsgemeinschaft, DFG) to L. Wiskott (SFB 874, TP B3) and supported by the German Federal Ministry of Education and Research within the National Network Computational Neuroscience - Bernstein Fokus: “Learning behavioral models: From human experiment to technical assistance”, grant FKZ 01GQ0951. \pagebreak
\section{Introduction} One of the most remarkable results in probabilistic symmetries is the de Finetti's theorem~\cite{de1931funzione}, which states that the law of any exchangeable sequence valued in a finite state space is in fact a mixture of i.i.d. sequences. This theorem has a geometrical interpretation via Choquet's theorem. More precisely, the subspace of exchangeable probabilities forms a convex, and those probabilities given by i.i.d. sequences are exactly the extreme points of the convex. In the 1920s, W.E. Johnson conjectured that, under some technical conditions, if a process \(X_n\) is exchangeable and \(\mathbb{ P}(X_{n+1}=i\vert X_0,\cdots,X_n)\) depends only on the number of times \(i\) occurs and the total steps \(n\), then \(X_n\) is nothing but the famous Polya urn: drawing balls uniformly from an urn and put back one additional ball with same color as the drawn one. This is a process with linear reinforcement. In term of random walk the natural counterpart of Polya urn is the edge reinforced random walk (ERRW): Diaconis conjectured that this process have the same characterization as Polya urn. In~\cite{rolles2003edge} S.W.W.Rolles have shown that both conjectures are true under technical conditions. The vertex reinforced jump process (VRJP) is a linearly reinforced process in continuous time. In a recent paper, Sabot and Tarres~\cite{sabot2011edge} have shown that ERRW is a mixture of VRJP, which indicates that the VRJP are building blocks of ERRW, thus should share a similar characterization. We prove that this is true in the sense that we give a counterpart of the characterization given by Rolles. \section{Definitions and results} Let \(G\) be a connected graph such that each vertex have finite degree, define its vertex set \(V\) and edge set \(E\). Denote \(i\sim j\) if \(\{i,j\}\in E\), assume that \(G\) contains no loops (edges with one endpoint). \begin{defi} \label{defi1} We call \((X_t)_{t\geq 0}\) a nearest neighbor jump processes on \(G\), if it is a random process which is right continuous without explosion, and each jump is from some vertex \(i\) to one of its neighbors \(j\) (i.e. \(i\sim j\)). \end{defi} \begin{defi} A nearest neighbor jump process \(X_t\) is a unique mixture of Markov jump process if there exists a unique probability measure \(\mu\) on Markov jump processes such that \(\mathcal{L}(X_t)=\int \mathcal{L}(Y_t) \mu(dY)\), where \(\mathcal{L}\) denotes the law of respective processes. If for \(\mu\) a.e. the Markov process is reversible, then the process is a unique mixture of reversible Markov process. \end{defi} Freedman introduced the notion of partial exchangeability in continuous time in~\cite{freedman1996finetti}. Define the transition count from \(i\) to \(j\) of a finite string of states of length \(n\): \(\xi=(\xi_0,\xi_1,\cdots,\xi_n)\) to be \[N_{i,j}(\xi)=\# \{k,\ 0\leq k\leq n-1, \ \xi_k=i,\xi_{k+1}=j\}.\] Define two finite strings of states \(\xi,\eta\) to be equivalent and denoted \(\xi \sim \eta\), if \(\xi\) and \(\eta\) start at the same state and the transition count from \(i\) to \(j\) of any pair \((i,j)\) are equal for \(\xi\) and \(\eta\), i.e. \(N_{i,j}(\xi)=N_{i,j}(\eta)\) for all \((i,j)\). \begin{defi}[Freedman] \label{defFreedman} A continuous process \(X_t\) is partially exchangeable if for each \(h>0\), the law of \(\{ X_{nh}:\ n=1,2,\cdots\}\) satisfies the following property: for any \( \xi\sim \eta\ \) of length \(l\), \[ \mathbb{ P}(X_0=\xi_0,\cdots,X_{lh}=\xi_l)=\mathbb{ P}(X_0=\eta_0,\cdots,X_{lh}=\eta_l).\] \end{defi} We recall the de Finetti's theorem for continuous time Markov chain introduced by Freedman~\cite{freedman1996finetti} here, \begin{thm} Let \(X_t\) be a continued time process starting at \(i_0\in G\), \(X_t\) is mixture of Markov jump process starting at \(i_0\) if \begin{enumerate} \item \(X_t\) has no fixed points of discontinuity, more precisely, for every \(t\), if \(t_n\rightarrow t\), then \(\mathbb{ P}(X_{t_n}\rightarrow X_t)=1\); \item \(X_t\) is recurrent; \item \(X_t\) is partially exchangeable. \end{enumerate} \end{thm} Next, we define the vertex reinforced jump process \(X_t\). Let \((W_e)_{e\in E}\) be weights on edges, the process \(X_t\) starts at time \(0\) at some vertex \(i_0\), and if \(X\) is at vertex \(i\in V\) at time \(t\), then, conditioned on the past, the process jumps to a neighbor \(j\) of \(i\) with rate \(W_{i,j}(1+l_j(t))\), where for \(e=\{i,j\}\), \(W_{i,j}=W_e\) and \(l_j(t)\) is the local time of vertex \(j\) at time \(t\): \[l_j(t):=\int_0^t \mathds{1}_{X_s=j}ds.\] This process turns out to be partially exchangeable within a time scale: let \[D(s)=\sum_{i\in V}(l_i(s)^2+2l_i(s)),\] then the process \(Y_t=X_{D^{-1}(t)}\) is a mixture of Markov chains, c.f.~\cite{sabot2011edge} Theorem 2. Now we can state our main theorem. \begin{thm} \label{thm1} Let \(X_t\) be a nearest neighbor jump process on \(G\) satisfying the following assumptions: \begin{enumerate} \item For all \(i\in V\), there exists \(\mathbb{R}^+\) homeomorphisms \(h_i\) such that \(X\) is partially exchangeable within the time scale \(D(s)=\sum_{i\in V}h_i(l_i(s))\); \item \(G\) is strongly connected (i.e. any two adjacent vertices are in a cycle); \item The process, at vertex \(i\) at time \(t\), jumps to a neighbor \(j\) of \(i\) with rate \(f_{i,j}(l_j(t))\) for some continuous functions \(f_{i,j}\) \end{enumerate} Then \(X\) is a vertex reinforced jump process within time scale, i.e. there exists another time scale \(\tilde{D}\) such that \(X_{\tilde{D}^{-1}(t)}\) is a vertex reinforced jump process. \end{thm} \begin{rmk} Note that we do not a priori require \(f_{i,j}=f_{j,i}\), i.e. there is no assumption of reversibility for \(X_t\); however the VRJP is a mixture of reversible Markov jump process within time change. \end{rmk} \begin{rmk} Concerning the third assumption, we cannot prove the result with rate \(f_{i,j}(l_i,l_j)\), but the case where \(f_{i,j}(l_i,l_j)=f_i(l_i)f_j(l_j)\) can be treated. In fact, by applying a time change, the process with rate function of the form \(f_i(l_i)f_j(l_j)\) can be reduced to our theorem. \end{rmk} In section 3, we introduce an equivalent notion of partial exchangeability and, as an example, we give a different proof of partial exchangeability of VRJP within a time scale. Section 4 contains the proof of Theorem \ref{thm1}. \section{The two notions of partial exchangeability} \subsection{Partial exchangeability, infinitesimal point of view} Consider a nearest neighbor jump process on \(G\) satisfying the third assumption of Theorem \ref{thm1}. As we have assumed regularity on the path of the process (c.f. Definition \ref{defi1}), to describe the law of our process, it is enough to describe the probability of the following events: \[ \sigma = \{ X_{[0,t_1[}=i_0,X_{[t_1,t_2[}=i_1,X_{[t_2,t_3[}=i_2,\cdots,X_{[t_{n-1},t_n[}=i_{n-1},X_{[t_n,t]}=i_n\},\] which is also denoted \[\sigma:\ i_0\xrightarrow[]{t_1}i_1\xrightarrow[]{t_2-t_1}i_2\cdots i_{n-1}\xrightarrow[]{t_n-t_{n-1}}i_n\xrightarrow[]{t-t_n}\] in the sequel and we call such an event {\it trajectory}. It turns out that when the jump rate is a continuous function of local times, the law of our process can be characterized by some function, which will be called {\it density} in the sequel. More precisely: \begin{defi} We say that \(X_t\) admits a density if, for all \(t\) and for all trajectories \[\sigma =i_0\xrightarrow[]{t_1}i_1\xrightarrow[]{t_2-t_1}i_2\cdots i_{n-1}\xrightarrow[]{t_n-t_{n-1}}i_n\xrightarrow[]{t-t_n},\] there exists an explicit function \(d_{\sigma}\), which is a function of \(t_k\) and \(i_k\), such that for all bounded measurable test function \(\Psi\) defined on the trajectories up to time \(t\), \[\mathbb{ E}(\Psi( X_u,\ u\leq t))=\sum_{n\geq 1}\sum_{i_0,\cdots,i_n}\int d_{\sigma} \Psi(\sigma)dt_1\cdots dt_n+d_{i_0\xrightarrow[]{t}}\Psi(i_0\xrightarrow[]{t})\] and obviously we have \(d_{i_0\xrightarrow[]{t}}=\mathbb{ P}(X_s=i_0,\ 0\leq s\leq t)\) for the trajectory with no jump. We will call such a function \(d_{\sigma}\) the {\it density} of \(\sigma\). \end{defi} Let us now give a notion of partial exchangeability for continuous time process in terms of density. Define two trajectories \(\sigma\) and \(\tau\) to be equivalent and denoted \(\sigma \sim \tau\), if their discrete chain strings are equivalent and the local times are equal at each vertex. Formally, \begin{defi} Let \[\sigma =i_0\xrightarrow[]{t_1}i_1\xrightarrow[]{t_2-t_1}i_2\cdots i_{n-1}\xrightarrow[]{t_n-t_{n-1}}i_n\xrightarrow[]{t-t_n},\] \[\tau =j_0\xrightarrow[]{s_1}j_1\xrightarrow[]{s_2-s_1}j_2\cdots j_{n-1}\xrightarrow[]{s_n-s_{n-1}}j_n\xrightarrow[]{t-s_n}.\] Then \(\sigma\) and \(\tau\) are equivalent if and only if \[\begin{cases} \forall i\in V, \ l_i^{\sigma}(t)=l_i^{\tau}(t)\\ \forall i,j\ N_{i,j}(\sigma)=N_{i,j}(\tau). \end{cases}\] where \(N_{i,j}(\sigma)\) denotes the number of jumps from \(i\) to \(j\) in \(\sigma\), i.e. \(N_{i,j}(\sigma)=N_{i,j}((i_0,\cdots,i_n))\), and \(l_i^{\sigma}(t)=\int_0^t \mathds{1}_{\sigma_s=i}ds\) denotes the local time. \end{defi} \begin{defi} \label{defdensity} A continuous time nearest neighbor jump process is said to be partially exchangeable in density if the densities are equal for any two equivalent trajectories. More precisely, the density depends only on the local times and the transition counts. \end{defi} \iffalse Although this is not the original definition of partial exchangeability in \cite{diaconis1980finetti}, we will show that when the process admits a density, this notion is equivalent to Diaconis and Freedman's.\fi \subsection{Equivalence of the two notions} It turns out that in the case of nearest neighbor jump process with continuous rate functions, the notion of partial exchangeability in Definition \ref{defFreedman} and in Definition \ref{defdensity} are equivalent. \begin{prop} \label{prop1} If a continuous time nearest neighbor jump process is partially exchangeable in the sense of Definition \ref{defdensity}, then it is partially exchangeable in Freedman's definition. \end{prop} \begin{proof} Suppose that the process \(X_t\) is partially exchangeable in density, let \(h>0,\) consider the event \(I= \{X_0=i_0,X_h=i_1,\cdots,X_{nh}=i_n\}\), let \((j_0=i_0,j_1,\cdots,j_n)\) be an equivalent string of \((i_0,\cdots,i_n)\), and \(J=\{X_0=j_0,X_h=j_1,\cdots,X_{nh}=j_n\}\). We can construct an application \(T\) which maps one continuous trajectory to another in such a way that \(T\) maps bijectively from \(I\) to \(J\). More precisely, as these two trajectories are equivalent, for any pair of neighbors \((i,j)\), there are exactly the same number of transition counts from \(i\) to \(j\). Let us define \(T\) to be the transformation which is a permutation of the time segmentations \([lh,(l+1)h)\) of size \(h\); which, for any \(k\), moves the \(k\)th transition \(i\xrightarrow[]{{k\text{th}}} j\) of \(I\) to the \(k\)th transition \(i\xrightarrow[]{k\text{th}} j\) of \(J\), and leaving the last time segmentation \([nh,\infty)\) invariant. Figure \ref{fig:1} illustrates an example of such application. \begin{figure}[!h] \centering \includegraphics[width=.7\textwidth]{transform_T.pdf} \caption{The transformation \(T\) for \(I=\{ X_0=0,X_h=1,X_{2h}=0,X_{3h}=2,X_{4h}=1\} \) and \(J=\{ X_0=0,X_h=2,X_{2h}=1,X_{3h}=0,X_{4h}=1\} \).} \label{fig:1} \end{figure} \noindent Let \[\sigma=k_0\xrightarrow[]{s_1}k_1\xrightarrow[]{s_2}k_2\cdots k_{N-1}\xrightarrow[]{s_N}k_N\xrightarrow[]{s_{N+1}}\] be one trajectory of the event \(I\), it is not hard to check that \[T(\sigma)=k'_0\xrightarrow[]{s'_1}k'_1\xrightarrow[]{s'_2}k'_2\cdots k'_{N-1}\xrightarrow[]{s'_N}k'_N\xrightarrow[]{s'_{N+1}}\] is a trajectory of the event \(J\), and that \(T\) is one-one and on-to (c.f. Figure \ref{fig:1.5}). If we fix the total number of jumps \(N\) and the discrete trajectory \((k_0,k_1,\cdots,k_N)\), then \(T\) can be though as a substitution of integration. Thus \begin{figure}[!h] \centering \includegraphics[width=.7\textwidth]{transform_T2.pdf} \caption{An example of \(\sigma\) and \(T(\sigma)\).} \label{fig:1.5} \end{figure} \begin{align*} \mathbb{ P}(I)&=\sum_N\sum_{k_0,k_1,\cdots k_N}\int \mathds{1}_{s_1,\cdots,s_{N+1}\in I(N,k_0,\cdots,k_N)}d_{\sigma}ds_1\cdots ds_{N+1} \\ &=\sum_N\sum_{k'_0,k'_1,\cdots,k'_N}\int \mathds{1}_{s_1,\cdots,s_N\in I'(N,k'_0,\cdots,k'_N)}d_{T(\sigma)} ds_1\cdots ds_{N+1} =\mathbb{ P}(J), \end{align*} where \(I(N,k_0,\cdots,k_N)\) is the subset of \(\mathbb{R}^{N+1}\) defined as the set of \((s_1,\cdots,s_{N+1})\) such that the event \(k_0\xrightarrow[]{s_1}k_1\xrightarrow[]{s_2}\cdots k_{N}\xrightarrow[]{s_{N+1}}\) is in \( I\); and \(I'(N,k_0,\cdots,k_N)\) is its image by applying \(T\); see Figure 2 for a concrete example. As \(T\) preserves local times and the numbers of transition counts, these two integrals are whence equal. \end{proof} \begin{prop} If a jump process is partially exchangeable in Freedman's sense, and its jump rate is a continuous function of local times (in which case the density can be written down explicitly), then it is also partially exchangeable in density. \end{prop} \begin{proof} Let \(X_t\) denote this process, for \(h>0\), consider the \(\sigma\)-algebra \(\mathcal{F}_h=\sigma(X_{nh}, n\geq 0)\), let \[\mathcal{F}_0=\sigma(\cup_{h>0}\mathcal{F}_h)\] and \[\mathcal{F}=\sigma(X_t,t\geq 0).\] As in \cite{freedman1996finetti}, we only consider \(h\) running through the binary rationals. Note that \(\mathcal{F}_0=\mathcal{F}\) thanks to the right continuity of the trajectories. Let \(\sigma=i_0\xrightarrow[]{t_1}i_1\xrightarrow[]{t_2-t_1}i_2\cdots i_n\xrightarrow[]{t-t_n}\) be a trajectory with \(n\) jumps (say \(n\geq 1\) to avoid triviality). Let \(\{X^{(h)}\sim \sigma/h\}\) denotes the event \[\{ X_0=\sigma_0, \ X_h=\sigma_h,\cdots,X_{Nh}=\sigma_{Nh},\ \text{ with }N=\lfloor t/h\rfloor\}.\] It turns out that \[d_{\sigma}=\lim_{h\rightarrow 0}\mathbb{ P}(X^{(h)}\sim \sigma/h)h^{-n}.\] In fact, let \(\Psi=\mathds{1}_{X^{(h)}\sim \sigma/h}\), by definition of \(d_{\sigma}\), \begin{equation}\label{den} \mathbb{ E}(\Psi( X_u,\ u\leq t))=\mathbb{ P}(X^{(h)}\sim \sigma/h)=\sum_{k\geq 1}\sum_{i_1,\cdots,i_k}\int d_{\tau} \Psi(\tau)dt_1\cdots dt_k\end{equation} where \[\tau =i_0\xrightarrow[]{t_1}i_1\xrightarrow[]{t_2-t_1}i_2\cdots i_{k-1}\xrightarrow[]{t_k-t_{k-1}}i_n\xrightarrow[]{t-t_n}.\] When \(h\) is small enough, the sum in \eqref{den} must be over \(k\geq n\), and we have \begin{align*} \mathbb{ P}(X^{(h)}\sim \sigma/h)=\mathbb{ P}_1+\mathbb{ P}_2. \end{align*} where for some \(p_k,k=1,\cdots, n\) depending on \(h\) \begin{align*}&\mathbb{ P}_1=\mathbb{ P}(\ (X_u)_{0\leq u\leq t} \text{ makes }n \text{ jumps at times }s_1,\cdots,s_n\text{ with }s_k\in (p_kh,(p_k+1)h] \ )\\ &\mathbb{ P}_2=\mathbb{ P}(\ (X_u)_{0\leq u\leq t} \text{ makes more than }n+1 \text{ jumps and } X^{(h)}\sim \sigma/h)\end{align*} Note that the jump rates are bounded from both below and above, by considering Poisson processes of jump rates \(C\), where \(C\) denotes the upper bound of the rate, we have \begin{align*} \mathbb{ P}_2 \leq \sum_{l\geq 1} \frac{(C(n+[l/2])h)^{n+l}}{(n+l)!e^{C(n+[l/2])h}}[t/h]^l . \end{align*} Applying Stirling's formula gives \(P_2\leq O(h^{n+1})\), thus \(P_2\) can be dropped when taking the limit. In addition, \[\mathbb{ P}_1=\int_{p_nh}^{(p_n+1)h}\cdots \int_{p_1h}^{(p_1+1)h} d_{\sigma}\ dt_1\cdots dt_n,\] note that here \(d_{\sigma}\) depends only on \(t_1,\cdots,t_n\) and it is an absolutely integrable function, by Lebesgue differentiation theorem (Theorem 1.6.19 \cite{tao2011introduction}) \(\lim_{h\rightarrow 0}\mathbb{ P}_1/h^{n}=d_{\sigma}\). \begin{figure}[!h] \centering \includegraphics[width=.5\textwidth]{diagram.pdf} \end{figure} Now let \(\sigma\sim \tau\), when \(h\) is sufficiently small, proceeding as in the diagram shows that \(d_{\sigma}=d_{\tau}\). \end{proof} \subsection{Example: VRJP is partially exchangeable within a time change} Recall that \(Y_s=X_{D^{-1}(s)}\), we can write down the density of the trajectory \(\sigma\) of the (time changed) VRJP process \(Y\) (For convenience, write \(s_{n+1}\) for \(s\) in the sequel), where \[\sigma:= \ i_0\xrightarrow[]{s_1}i_1\xrightarrow[]{s_2-s_1}i_2\cdots i_{n-1}\xrightarrow[]{s_n-s_{n-1}}i_n\xrightarrow[]{s-s_n}\] its density is (c.f.~\cite{sabot2013ray}) \begin{equation} \label{densityST} \begin{aligned} d_{\sigma}=&(\frac{1}{2})^n\prod_{k=1}^nW_{i_{k-1},i_k}\prod_{i\in V, i\neq i_n}\frac{1}{\sqrt{1+S_i(s)}}\prod_{k=1}^{n+1}ds_k\\ &\cdot\exp{(- \sum_{i \sim j} \frac{W_{i,j}}{2}(\sqrt{(S_i(s)+1)(S_j(s)+1)}-1))}, \end{aligned} \end{equation} which depends only on final local times and transition counts, thus by Proposition \ref{prop1}, \(Y\) is partially exchangeable. On finite graph it is rather easy to prove that the VRJP is recurrent (for example, using a representation of VRJP by time changed Poisson point process as in \cite{sabot2011edge}, and then use an argument as in~\cite{davis1990reinforced} or~\cite{sellke1994reinforced}). Therefore, \(Y\) is a mixture of Markov jump process; in addition, Sabot and Tarres have computed the mixing measure in \cite{sabot2011edge}. For convenient, we include a proof of this in the sequel, as a corollary of the Proposition 3, since the mechanisms of this proof enlights the proof of the main theorem. \section{Proof of theorem 1} \subsection{Computation of densities} Let \(X\) be a nearest neighbor jump process on \(G\) satisfying the assumptions of Theorem \ref{thm1}, in particular, recall the time scale \begin{equation} \label{timescale} D(s)=\sum_{i\in V}h_i(l_i(s)).\end{equation} Let \(l_i(t)\) be the local time of the process \(X\) at vertex \(i\) at time \(t\). Let us denote the process after time change to be \begin{equation}\label{processY} Y_t=X_{D^{-1}(t)},\end{equation} let \begin{equation} \label{localtimeY}S_i(s)=\int_0^s \mathds{1}_{Y_u=i}du\end{equation} denote the local time of \(Y\). Consider the trajectory \begin{equation}\label{trajY}\sigma: \ i_0\xrightarrow[]{t_1}i_1\xrightarrow[]{t_2-t_1}i_2\cdots i_{n-1}\xrightarrow[]{t_n-t_{n-1}}i_n\xrightarrow[]{t-t_n}\end{equation} where \(0<t_1<\cdots <t_n<t\), by applying the time change, the corresponding trajectory for \(Y\) is \[\sigma_Y: \ i_0\xrightarrow[]{s_1}i_1\xrightarrow[]{s_2-s_1}i_2\cdots i_{n-1}\xrightarrow[]{s_n-s_{n-1}}i_n\xrightarrow[]{s-s_n}\] where \(s_k=D(t_k)\). \begin{prop} \label{prop3} With the same settings as in equations \eqref{timescale} \eqref{processY} \eqref{localtimeY} \eqref{trajY}, the density of the trajectory \(\sigma_Y\) for \(Y\) is \[d_{\sigma}^Y=\exp{\left(-\int_0^{s}\sum_{j\sim Y_v}\frac{f_{Y_v,j}(h_j^{-1}(S_j(v)))}{h'_{Y_v}(h_{Y_v}^{-1}(S_{Y_v}(v)))}dv\right)}\prod_{k=1}^n \frac{f_{i_{k-1},i_k}(h_{i_k}^{-1}(S_{i_k}(s_{k-1})))}{h'_{i_{k-1}}(h^{-1}_{i_{k-1}}(S_{i_{k-1}}(s_k)))}.\] \end{prop} \begin{proof} Remark that if at time \(t_k\) the process just jumps to \(i_k\), then during the time interval \([t_k,t_{k+1})\), the jump rate to any adjacent vertex \(j\) is \(f_{i_k,j}(l_j(t))\), which is constant equal to \(f_{i_k,j}(l_j(t_k))\). Therefore, the holding time \(\tau_{k+1}:=t_{k+1}-t_k\) is exponentially distributed of rate \(\sum_{j\sim i_k}f_{i_k,j}(l_j(t_k))\). On the other hand, using an elementary property of exponential variable, the probability that at time \(t_{k+1}\) the process jumps to \(i_{k+1}\) is \[\frac{f_{i_k,i_{k+1}}(l_{i_{k+1}}(t_k))}{\sum_{j\sim i_k}f_{i_k,j}(l_j(t_k))}.\] Combining these and using substitution \(t_k=\tau_1+\cdots+\tau_k\), the density \(d_{\sigma}\) for the process \(X\) admits the following explicit form: \[d_{\sigma}=\exp{\left(-\int_0^{t}\sum_{j\sim X_u} f_{X_u,j}(l_j(u))du\right)}\prod_{k=1}^n f_{i_{k-1},i_k}(l_{i_k}(t_{k-1})).\] Recall that in \eqref{timescale} we assumed that \(h_i: \mathbb{R^+}\rightarrow \mathbb{R^+}\) are diffeomorphisms satisfying \(h_i(0)=0\). Next we compute the same density but for the process \(Y_s=X_{D^{-1}(s)}\), as we have \(S_i(D(s))=h_i(l_i(s))\), derivation leads to \[S_i(D(s))'=D'(s)\mathds{1}_{Y_{D(s)}=i}=h_i'(l_i(s))\mathds{1}_{X_s=i}.\] Hence \[(D^{-1}(t))'=\frac{1}{D'(D^{-1}(t))}=\frac{1}{h'_{Y_{t^-}}\circ h_{Y_{t^-}}^{-1}(S_{Y_{t^-}}(t))}\] \[l_{i_k}(t_{k-1})=h_{i_k}^{-1}(S_{i_k}(D(t_{k-1})))=h^{-1}_{Y_{S_k}}(S_{Y_{S_k}}(s_{k-1})).\] By substitution \(s=D^{-1}(t)\), we have \begin{align*}d^Y_{\sigma}= \exp{(-\int_0^{s}\sum_{j\sim Y_v}\frac{f_{Y_v,j}(h_j^{-1}(S_j(v)))}{h'_{Y_v}(h_{Y_v}^{-1}(S_{Y_v}(v)))}dv)}\prod_{k=1}^n \frac{f_{i_{k-1},i_k}(h_{i_k}^{-1}(S_{i_k}(s_{k-1})))}{h'_{i_{k-1}}(h^{-1}_{i_{k-1}}(S_{i_{k-1}}(s_k)))}. \end{align*} \end{proof} {\bf Back to the partial exchangeability of VRJP} \begin{proof} Apply the previous proposition to VRJP, where \(f_{i,j}(l_j)=W_{i,j}(1+l_j)\) and \(h_i(l_i)=l_i^2+2l_i\). The density \(d_{\sigma}^Y\) is \[\frac{1}{2^n}\exp\left ( -\int_0^{s}\sum_{j\sim Y_u} \frac{W_{Y_u,j}\sqrt{S_j(u)+1}}{2\sqrt{S_{Y_u}(u)+1}}du\right) \prod_{k=1}^n\left(W_{i_{k-1},i_k} \frac{\sqrt{S_{i_k}(s_{k-1})+1}}{\sqrt{S_{i_{k-1}}(s_k)+1}}\right).\] As our path is left continuous without explosion, starting at \(i_0\), if we calculate the product through the path, by telescopic simplification, it results that the product reduces to \[\prod_{i\in V\ i\neq i_n}\frac{1}{\sqrt{S_i(s)+1}} \prod_{k=1}^n W_{i_{k-1},i_k}.\] While the integral inside the exponential becomes \begin{align*} &\int_0^s \sum_{j\sim Y_u}\frac{W_{Y_u,j}\sqrt{S_j(u)+1}}{2\sqrt{S_{Y_u}(u)+1}}du\\ &=\sum_{k=1}^{n+1}\int_{s_{k-1}}^{s_k}\sum_{j\sim Y_u}\frac{W_{Y_u,j}\sqrt{S_j(u)+1}}{2\sqrt{S_{Y_u}(u)+1}}du \\ &=\sum_{k=1}^{n+1}\sum_{j\sim i_{k-1}}\frac{W_{i_k,j}}{2}\sqrt{S_j(s_{k-1})+1}\left(\sqrt{S_{i_{k-1}}(s_k)+1}-\sqrt{S_{i_{k-1}}(s_{k-1})+1}\right)\\ &=\sum_{i\sim j}\frac{W_{i,j}}{2}\sum_{k=1}^{n+1}\mathds{1}_{i_{k-1}=i}\left(\sqrt{(S_j(s_{k-1})+1)(S_i(s_k)+1)}-\sqrt{(S_j(s_{k-1})+1)(S_i(s_{k-1})+1)}\right). \end{align*} summing through the path, for every pair \(i,j\), there goes another telescopic simplification, which gives \eqref{densityST}, and expression \eqref{densityST} depend only on final local times and transition counts, the result hence follows. \end{proof} \subsection{Determination of time change \(h\)} In the sequel we work with the time changed process \(Y\), to simplify notations, we will write \(d_{\sigma}\) for \(d^Y_{\sigma}\) when it does not lead to any confusion. By Proposition \ref{prop3}, the density of certain trajectory contains an exponential term and a product term, let us denote \[d_{\sigma}=\exp(-\int\sigma) \cdot \prod \sigma,\] with \[\begin{cases} \int \sigma=\int_0^{s}\sum_{j\sim Y_v}\frac{f_{Y_v,j}(h_j^{-1}(S_j(v)))}{h'_{Y_v}(h_{Y_v}^{-1}(S_{Y_v}(v)))}dv\\ \prod \sigma =\prod_{k=1}^n \frac{f_{i_{k-1},i_k}(h_{i_k}^{-1}(S_{Y_{s_k}}(s_{k-1})))}{h'_{i_{k-1}}(h^{-1}_{i_{k-1}}(S_{Y_{s_{k-1}}}(s_k)))} \end{cases}\] Where the exponential term stems from those exponential waiting times, and the product term corresponds to the probability of the discrete chain. The heuristics of this subsection is the following: as we assumed partial exchangeability, if we consider two equivalent trajectories, then their densities share the same expression, by comparing them we can hence deduce certain equalities involving \(f_{i,j}\) and \(h_i\) etc. It turns out that these equalities determine \(h_i\)s then \(f_{i,j}\)s. The following fact is simple but important, suppose that at time \(s\), the random walker arrives at \(i_0\), each vertex \(i\) has accumulated local time \(l_i:=S_i(s)\); then it jumps to \(i_1\) after an amount of time \(t\), by Proposition \ref{prop3}, the density has acquired a multiplicative factor \begin{equation} \label{factor} \exp\left(-\int_s^{s+t}\sum_{j\sim i_0}\frac{f_{i_0,j}\circ h_j^{-1}(l_j)}{h'_{i_0}\circ h_{i_0}^{-1}(l_{i_0}+v)}dv\right)\cdot\frac{f_{i_0,i_1}\circ h_{i_1}^{-1}(l_{i_1})}{h'_{i_0}\circ h_{i_0}^{-1}(l_{i_0}+t)}. \end{equation} This fact is in constant use in the sequel, when we explicit the density of certain trajectory. \begin{lem} \label{lem0} Let \(\sigma=i_0\xrightarrow[]{s_1}i_1\xrightarrow[]{s_2-s_1}i_2\cdots i_{n-1}\xrightarrow[]{s_n-s_{n-1}}i_n\xrightarrow[]{s-s_n} \) be a trajectory, then \(\displaystyle \int \sigma=\int \tilde{\sigma} + \int \hat{\sigma}\) where \[\int \tilde{\sigma}= \int_0^s \sum_{j\in \sigma, j\sim Y_v}\frac{f_{Y_v,j}(h_j^{-1}(S_j(v)))}{h'_{Y_v}(h_{Y_v}^{-1}(S_{Y_v}(v)))}dv,\ \int\hat{\sigma}=\int_0^s \sum_{j\notin \sigma, j\sim Y_v}\frac{f_{Y_v,j}(h_j^{-1}(S_j(v)))}{h'_{Y_v}(h_{Y_v}^{-1}(S_{Y_v}(v)))}dv \] and if \(\tau\) is such that \(\tau\sim \sigma\), then \(\displaystyle \int \hat{\sigma}=\int \hat{\tau}.\) \end{lem} \begin{proof} Note that for \(j\notin \sigma\), \(S_j(u)=0\) for all \(u\leq s\). Let \(\hat{H}_i\) be a primitive of \(\displaystyle \frac{1}{h_i'\circ h_i^{-1}}\), \begin{align*} \int \hat{\sigma}&=\sum_{j\notin \sigma}\int_0^s \mathds{1}_{Y_v\sim j} \frac{f_{Y_v,j}(0)}{h'_{Y_v}(h_{Y_v}^{-1}(S_{Y_v}(v)))}dv \\ &=\sum_{j\notin \sigma,i\in \sigma,j\sim i}f_{i,j}(0) \int_0^s \frac{\mathds{1}_{Y_v=i}}{h'_i(h_i^{-1}(S_i(v)))}dv \\ &=\sum_{j\notin \sigma,i\in \sigma,j\sim i}f_{i,j}(0) (\hat{H}(S_i(s))-\hat{H}(0)) \end{align*} which depends only on final local times, thus if \(\tau\sim \sigma\), then \(\displaystyle \int\hat{\tau}=\int\hat{\sigma}\). \end{proof} In the sequel \(cst\) denotes some constant, which can vary from line to line. \begin{lem} \label{lem1} If the process \(X\) admits such a time change \(D\) which makes it partially exchangeable in density, then for any \(i\sim j\), there exists some constants \(\lambda_{i,j}\) such that \begin{equation} \label{eq:1} f_{i,j}(x)=\lambda_{i,j}h_j'(x),\ \forall x\geq 0. \end{equation} \end{lem} \begin{proof} Let \(\epsilon>0\), consider the following two trajectories for the process \(Y\): \[ \sigma = i\xrightarrow[]{\epsilon}j\xrightarrow[]{\epsilon}i\xrightarrow[]{t}j\xrightarrow[]{s}i \xrightarrow[]{\cdot}\] \[ \tau = i\xrightarrow[]{t}j\xrightarrow[]{s}i\xrightarrow[]{\epsilon}j\xrightarrow[]{\epsilon}i\xrightarrow[]{\cdot}\] Note that \(\sigma\) and \(\tau\) have the same transition counts and the final local times on vertex \(i,j\) are respectively equal. Thus the densities of these trajectories are a.s. equal by partial exchangeability. By Lemma \ref{lem0}, \begin{align*} d_{\sigma}= \prod \sigma \cdot \exp(\int \tilde{\sigma}+\int\hat{\sigma}), \end{align*} with \[ \begin{cases} \prod \sigma =\frac{f_{i,j}\circ h_j^{-1}(0)}{h_i'\circ h_i^{-1}(\epsilon)}\cdot\frac{f_{j,i}\circ h_i^{-1}(\epsilon)}{h_j'\circ h_j^{-1}(\epsilon)}\cdot \frac{f_{i,j}\circ h_j^{-1}(\epsilon)}{h_i'\circ h_i^{-1}(\epsilon+t)}\cdot\frac{f_{j,i}\circ h_i^{-1}(\epsilon+t)}{h_j'\circ h_j^{-1}(\epsilon+s)}\\ \int \tilde{\sigma}= \int_0^{\epsilon}\frac{f_{i,j}\circ h_j^{-1}(0)}{h_i'\circ h_i^{-1}(v)}dv+ \int_0^{\epsilon}\frac{f_{j,i}\circ h_i^{-1}(\epsilon)}{h_j'\circ h_j^{-1}(v)}dv+ \int_0^{t}\frac{f_{i,j}\circ h_j^{-1}(\epsilon)}{h_i'\circ h_i^{-1}(\epsilon+v)}dv+ \int_0^{s}\frac{f_{j,i}\circ h_i^{-1}(\epsilon+t)}{h_j'\circ h_j^{-1}(\epsilon+v)}dv. \end{cases}\] \begin{align*} d_{\tau}=\prod \tau \cdot \exp(\int \tilde{\tau}+\int\hat{\tau}), \end{align*} with \[\begin{cases} \prod \tau= \frac{f_{i,j}\circ h_j^{-1}(0)}{h_i'\circ h_i^{-1}(t)}\cdot\frac{f_{j,i}\circ h_i^{-1}(t)}{h_j'\circ h_j^{-1}(s)}\cdot \frac{f_{i,j}\circ h_j^{-1}(s)}{h_i'\circ h_i^{-1}(t+\epsilon)}\cdot\frac{f_{j,i}\circ h_i^{-1}(\epsilon+t)}{h_j'\circ h_j^{-1}(\epsilon+s)} \\ \int \tilde{\tau}= \int_0^{t}\frac{f_{i,j}\circ h_j^{-1}(0)}{h_i'\circ h_i^{-1}(v)}dv+ \int_0^{s}\frac{f_{j,i}\circ h_i^{-1}(t)}{h_j'\circ h_j^{-1}(v)}dv+ \int_0^{\epsilon}\frac{f_{i,j}\circ h_j^{-1}(s)}{h_i'\circ h_i^{-1}(t+v)}dv+ \int_0^{\epsilon}\frac{f_{j,i}\circ h_i^{-1}(\epsilon+t)}{h_j'\circ h_j^{-1}(s+v)}dv;\end{cases}\] where we do not explicte \(\int \hat{\sigma}\) and \(\int\hat{\tau}\) as they cancel when we compare these expressions (c.f. Lemma \ref{lem0}). Letting \(\epsilon\rightarrow 0\) yields that \(\exp(\int \tilde{\sigma})=\exp(\int \tilde{\tau})\); therefore \(\prod \sigma=\prod \tau\), i.e. \[\forall s,t,\ \frac{f_{i,j}\circ h_j^{-1}(s)}{h_j'\circ h_j^{-1}(s)} \cdot \frac{f_{j,i} \circ h_i^{-1}(t)}{h_i'\circ h_i^{-1}(t)}=cst.\] Now fix \(t\), let \(s\) vary, whence \[\forall s,\ f_{i,j}\circ h_j^{-1}(s)=cst\cdot h_j'\circ h_j^{-1}(s),\] and let \(\lambda_{i,j}\) denotes this constant, as \(h_j^{-1}\) is a diffeomorphism, its range is \(\mathbb{R}^+\), which allows us to conclude. \end{proof} The next lemma states in some sense that the exponential part and the product part appearing in the density of a trajectory can be treated separately. \begin{lem} \label{lem2} Let \(\sigma,\tau\) be two trajectories, and denote \[d_{\sigma}=\exp(\int \sigma) \cdot \prod \sigma,\hspace{.5cm} d_{\tau}=\exp(\int \tau)\cdot \prod \tau,\] if \(\sigma \sim \tau\), then \(\prod \sigma =\prod \tau\). \end{lem} \begin{proof} We have \(S_{Y_{s_k}}(s_k)=S_{Y_{s_k}}(s_{k-1})\), thus Lemma \ref{lem1} yields that \(f_{i_{k-1},i_k}\circ h_{i_k}^{-1}(S_{Y_{s_k}}(s_{k-1}))=\lambda_{i_{k-1},i_k}h_{i_k}'\circ h_{i_k}^{-1}(S_{Y_{s_k}}(s_{k}))\). Whence the product part is \[\prod \sigma = \prod_{k=1}^n \frac{f_{i_{k-1},i_k}(h_{i_k}^{-1}(S_{Y_{s_k}}(s_{k-1})))}{h'_{i_{k-1}}(h^{-1}_{i_{k-1}}(S_{Y_{s_{k-1}}}(s_k)))}ds_k=\prod_{k=1}^n \lambda_{i_{k-1},i_k}\prod_{i\in V}\frac{1}{h_i'\circ h_i^{-1}(S_i(t_n))},\] and the last term depends only on the transition counts and final local times. \end{proof} \begin{lem} \label{lem3} Let \(H_i=h_i'\circ h_i^{-1}\), then for some constant \(A_i\), \[ (H_i^2)'=A_i \text{ and if }i\sim j, \text{ then } \lambda_{i,j}A_j=\lambda_{j,i}A_i.\] \end{lem} \begin{rmk} The latest equality tells that the process is reversible. However, we did not assume the reversibility of the process, but vertex reinforced jump processes are reversible (as a mixture of reversible Markov jump process), so are the edge reinforced random walks. In contrast, directed edge reinforced random walks are mixtures of non reversible Markov chains, with independent Dirichlet environments. We can hence expect that the reversibility is a consequence of a non oriented linear reinforcement (where linearity corresponds to partial exchangeability). \end{rmk} \begin{proof} Recall that we have assumed that the graph is strongly connected, i.e. if \(i,j\) are two adjacent vertices, there exists a shortest cycle \(i_1\sim i_2\sim i_3\cdots \sim i_n\sim i_1\) with \(i_1=i,i_n=j\) and the \(i_k\)s are distinct. \begin{figure}[!h] \centering \includegraphics[width=.7\textwidth]{sigma_tau.pdf} \caption{the trajectories \(\sigma\) and \(\tau\) in Lemma \ref{lem3}.} \label{fig:st} \end{figure} Let \((i_1=i,i_2,i_3,\cdots,i_n=j)\) be a cycle as described, consider the trajectories (c.f. Figure \ref{fig:st}) \[\sigma=i_1\xrightarrow[]{r_1}i_n\xrightarrow[]{r_2}i_1\xrightarrow[]{s_1}i_2\xrightarrow[]{s_2}i_{3}\cdots i_{n-2}\xrightarrow[]{s_{n-2}}i_{n-1}\xrightarrow[]{s_{n-1}}i_n \] \[\tau=i_1\xrightarrow[]{r_1}i_2\xrightarrow[]{s_2}i_{3} \cdots i_{n-2}\xrightarrow[]{s_{n-2}}i_{n-1}\xrightarrow[]{s_{n-1}}i_n\xrightarrow[]{r_2}i_1\xrightarrow[]{s_1}i_n.\] As \(\sigma\sim \tau\), by Lemma \ref{lem2} and Lemma \ref{lem0}, \(\int \tilde{\sigma}=\int \tilde{\tau}\). Also let \[\sigma'=i_1\xrightarrow[]{r_1}i_n\xrightarrow[]{r_2}i_1\xrightarrow[]{s_1}i_2\xrightarrow[]{s_2}i_1\] \[\tau'=i_1\xrightarrow[]{r_1}i_2\xrightarrow[]{s_2}i_1\xrightarrow[]{s_1}i_n\xrightarrow[]{r_2}i_1,\] thus \(\int \tilde{\sigma'}=\int \tilde{\tau'}\). We are going to compute explicitly \(\int \tilde{\sigma},\ \int\tilde{\tau}\) etc, using \eqref{factor}, let \(s=r_1+r_2+s_1+\cdots+s_{n-1}\) and recall that \(\hat{H}_i\) is a primitive of \(\displaystyle \frac{1}{h_i'\circ h_i^{-1}}\). \begin{align*} \int\tilde{\sigma}=&\sum_{(i,j)\in\sigma^2,i\sim j} \lambda_{i,j}\int_0^{s} \mathds{1}_{Y_v=i}\frac{h_j'\circ h_j^{-1}(v)}{h_i'\circ h_i^{-1}(v)}dv \\ &=\lambda_{i_1,i_2}H_{i_2}(0)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(0)) + \lambda_{i_2,i_1}H_{i_1}(r_1+s_1)(\hat{H}_{i_2}(s_2)-\hat{H}_{i_2}(0))\\ &+\lambda_{i_1,i_n}\left(H_{i_n}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_n}(r_2)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\right)\\ &+\lambda_{i_n,i_1}H_{i_1}(r_1)(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0))+\lambda_{i_n,i_{n-1}}H_{i_{n-1}}(0)(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0))\\ &+\lambda_{i_{n-1},i_n}H_{i_n}(r_2)(\hat{H}_{i_{n-1}}(s_{n-1})-\hat{H}_{i_{n-1}}(0)) + \Delta \end{align*} where \(\Delta\) is defined as follows: let \(Q_k:=H_{i_{k}}(0)(\hat{H}_{i_{k-1}}(s_{i_{k-1}})-\hat{H}_{i_{k-1}}(0))\) and \(Q_k':=H_{i_k}(s_k)(\hat{H}_{i_{k+1}}(s_{i_{k+1}})-\hat{H}_{i_{k+1}}(0))\), \begin{align*} \Delta=\sum_{k=3}^{n-1}\lambda_{i_{k-1},i_k}Q_k+\lambda_{i_k,i_{k-1}}Q'_{k-1}. \end{align*} For \(\tilde{\tau}\) we have: \begin{align*} \int\tilde{\tau}=&\sum_{(i,j)\in\tau^2,i\sim j} \lambda_{i,j}\int_0^{s} \mathds{1}_{Y_v=i}\frac{h_j'\circ h_j^{-1}(v)}{h_i'\circ h_i^{-1}(v)}dv \\ &=\lambda_{i_1,i_2}H_{i_2}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_2}(s_2)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1)) \\ &+\lambda_{i_2,i_1}H_{i_1}(r_1)(\hat{H}_{i_2}(s_2)-\hat{H}_{i_2}(0))\\ &+\lambda_{i_1,i_n}\left(H_{i_n}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_n}(r_2)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\right)\\ &+\lambda_{i_n,i_1}H_{i_1}(r_1)(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0))+\lambda_{i_n,i_{n-1}}H_{i_{n-1}}(s_{n-1})(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0))\\ &+\lambda_{i_{n-1},i_n}H_{i_n}(0)(\hat{H}_{i_{n-1}}(s_{n-1})-\hat{H}_{i_{n-1}}(0)) + \Delta \end{align*} with the same \(\Delta\). Also \begin{align*} \int\tilde{\sigma}'&=\lambda_{i_1,i_2}\left(H_{i_2}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_2}(0)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\right) \\ &+\lambda_{i_2,i_1}H_{i_1}(r_1+s_1)(\hat{H}_{i_2}(s_2)-\hat{H}_{i_2}(0))\\ &+\lambda_{i_1,i_n}\left(H_{i_n}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_n}(r_2)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\right)\\ &+\lambda_{i_n,i_1}H_{i_1}(r_1)(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0)) \end{align*} \begin{align*} \int\tilde{\tau}'&=\lambda_{i_1,i_2}\left(H_{i_2}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_2}(s_2)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\right) \\ &+\lambda_{i_2,i_1}H_{i_1}(r_1)(\hat{H}_{i_2}(s_2)-\hat{H}_{i_2}(0))\\ &+\lambda_{i_1,i_n}\left(H_{i_n}(0)(\hat{H}_{i_1}(r_1)-\hat{H}_{i_1}(0))+H_{i_n}(0)(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\right)\\ &+\lambda_{i_n,i_1}H_{i_1}(r_1+s_1)(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0)). \end{align*} Recall that \(\int\sigma-\int\sigma'=\int\tau-\int\tau'\), which leads to \begin{align*} &\lambda_{i_n,i_{n-1}}H_{i_{n-1}}(0)(\tilde{H}_{i_n}(r_2)-\tilde{H}_{i_n}(0))+\lambda_{i_{n-1},i_n}H_{i_n}(r_2)(\hat{H}_{i_{n-1}}(s_{n-1})-\hat{H}_{i_{n-1}}(0))\\ &=\lambda_{i_1,i_n}(H_{i_n}(r_2)-H_{i_n}(0))(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))\\ &+\lambda_{i_n,i_1}(H_{i_1}(r_1)-H_{i_1}(r_1+s_1))(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0))\\ &+\lambda_{i_n,i_{n-1}}H_{i_{n-1}}(s_{n-1})(\tilde{H}_{i_n}(r_2)-\tilde{H}_{i_n}(0))+\lambda_{i_{n-1},i_n}H_{i_n}(0)(\hat{H}_{i_{n-1}}(s_{n-1})-\hat{H}_{i_{n-1}}(0)) \end{align*} letting \(s_{n-1}\rightarrow 0\) leads to \begin{align*} & \lambda_{i_1,i_n}(H_{i_n}(r_2)-H_{i_n}(0))(\hat{H}_{i_1}(r_1+s_1)-\hat{H}_{i_1}(r_1))=\\ &\lambda_{i_n,i_1}(H_{i_1}(r_1+s_1)-H_{i_1}(r_1))(\hat{H}_{i_n}(r_2)-\hat{H}_{i_n}(0)) \end{align*} as \(i_1,i_n,r_2,s_1,r_1\) are arbitrary, divide the formula by \(r_2s_1\) and let \(r_2,s_1\) go to zero leads to \[\lambda_{i_1,i_n}H'_{i_n}(0)\hat{H}'_{i_1}(r_1)=\lambda_{i_n,i_1}H'_{i_1}(r_1)\hat{H}'_{i_n}(0),\] finally note that \(\hat{H}'_i=1/H_i\), thus \(\lambda_{i_1,i_n}(H_{i_n}^2)'(0)=\lambda_{i_n,i_1}(H_{i_1}^2)'(r_1)\). \end{proof} \begin{lem} For all \(i\sim j\), let \(W_{i,j}=\lambda_{i,j}A_j/2=\lambda_{j,i}A_i/2\), there exists constant \(D_j\) depends only on \(j\), such that \(f_{i,j}(x)=W_{i,j}x+D_j\). \end{lem} \begin{proof} As \((H_j^2(s))'=A_j\), there exists \(B_j\) such that \(H_j^2(s)=A_js+B_j\), therefore \[f_{i,j}\circ h_j^{-1}(s)=\lambda_{i,j}H_j(s)=\lambda_{i,j}\sqrt{A_js+B_j}.\] On the other hand, \((h^{-1}_j)'(s)=\frac{1}{\sqrt{A_js+B_j}}\), thus for some \(C_j\), \[h_j^{-1}(s)=\frac{2}{A_j}\sqrt{A_js+B_j}+C_j.\] \(f_{i,j}(h_j^{-1}(s))=f_{i,j}(\frac{2}{A_j}\sqrt{A_js+B_j}+C_j)=\lambda_{i,j}\sqrt{A_js+B_j}\), which leads to \[f_{i,j}(x)=W_{i,j}x+D_j,\] where \(D_j\) is some constant depends only on \(j\). Applying the time change \[D(s)=\sum_i \frac{l_i(s)-D_i}{D_i},\] the resulting process will be of jump rate \[W_{i,j}D_iD_j(1+T_j(t))\] where \(T_j(t)\) is the local time for the time changed process \(Z_t=X_{D^{-1}(t)}\). \end{proof} \noindent{\bf Acknowledgments:} I would like to thank Christophe Sabot for his constant support in this project. \nocite{*}
\section{Introduction} \label{s:intro} Between $z\sim11$ and $z\sim6$, the universe experienced a change of state known as cosmic reionization \citep[e.g.,][]{b:Shull_apj12,b:Fan_araa06,b:Fan_aj01}. The source for cosmic reionization is still unclear; the primary contenders are star-forming galaxies and quasars. However, the space density of quasars peaks at $z\sim3$ and drops below the threshold needed to supply the required amount of ionizing radiation at $z\gtrsim 6$ \citep{b:Madau_apj99,b:Fan_aj01,b:Meiksin_mnras05, b:Shankar_apj07}. The massive stars found in high-redshift star-forming galaxies have the potential to provide enough ionizing radiation, but whether they contribute significantly to the intergalactic ionizing emissivity is not yet established \citep[e.g.,][]{b:Yajima_arXiv12}. In order to account for cosmic reionization, the mean fraction of ionizing radiation that escapes the galaxy population $<\mbox{$f_{\rm esc}$}>$, needs to be at least 20\% \citep[e.g.,][]{b:Bouwens_apj10}. In theory, starburst galaxies should be porous to ionizing radiation \citep[e.g.,][]{b:Clarke_mnras02,b:Paardekooper_aap11}. Stellar winds and supernovae, in conjunction with radiation pressure, will punch holes in the interstellar medium (ISM) \citep[e.g.,][]{b:MacLow_apj88}, and thereby create low density passageways through which ionizing radiation may escape \citep[e.g.,][]{b:Dove_apj00}. Feedback from massive stars is strongest in galaxies with high star-formation rates, making starbursts attractive candidates for the source of cosmic reionization. The observational evidence for significant \mbox{$f_{\rm esc}$}\ at high redshift is inconclusive. Using a sample of 29 Lyman break galaxies (LBG), \citet{b:Steidel_apj01} showed that at least some $z\sim3$ starbursts have large escape fractions. However, Lyman continuum narrow-band imaging studies with larger galaxy samples found escaping Lyman continuum in only 10\% of the galaxies observed \citep[e.g.,][]{b:Shapley_apj06,b:Iwata_apj09}. Observations by \citet{b:Nestor_apj11}, put the mean $z\sim 3$ escape fraction at $<\mbox{$f_{\rm esc}$}>\sim 0.12$. Is it important to note that \citet{b:Vanzella_apj12} recently illustrated that a majority of narrowband Lyman continuum detections could actually be low $z$ interlopers, which would drive the observational \mbox{$f_{\rm esc}$}\ even lower. At intermediate to low redshift, some evidence exists for significant \mbox{$f_{\rm esc}$}\ from LGB analogs. A subset of these galaxies with extreme feedback have favorable conditions for escaping Lyman continuum, such as low opacity in the ISM \citep{b:Heckman_apj11}. However, the fraction of galaxies with these conditions is still small, indicating a low mean \mbox{$f_{\rm esc}$}\ \citep[e.g.,][]{b:Siana_apj10}. Looking at $z\sim0$ galaxies, the detection rate drops further, with only two examples of local starbursts that have measurable \mbox{$f_{\rm esc}$}\ \citep{b:Leitet_aap13, b:Leitet_aap11}. Most other studies find no direct evidence for escaping Lyman continuum in local starbursts \citep{b:Bridge_apj10,b:Grimes_apjs09,b:Heckman_apj01,b:Leitherer_apj95}. In order to determine the mean galactic \mbox{$f_{\rm esc}$}\ from observed escape fractions, we need to disentangle the effects of observational biases from trends with intrinsic galaxy properties. In the massive-star feedback-driven model, the escape fraction will depend on the star formation rate, the ISM morphology and density \citep{b:Fernandez_apj11,b:Benson_arXiv12}, and distribution of star formation. However, the winds that clear the ISM are not isotropic, which leads to a preferred direction for escape and finite opening angles \citep{b:Veilleux_araa05}. These effects create observational dependencies with the orientation of the host galaxy. Finally, measuring the Lyman continuum at any redshift is a challenge. Most of the flux is absorbed by the intergalactic medium, thus the signal is low, which requires long integration times. For targets at higher redshift, low-redshift interlopers confuse the statistics \citep{b:Vanzella_apj12}. Some studies measure the optical depth to UV lines in the ISM in lieu of direct measurements of the Lyman continuum. They obtain the escape fraction after making assumptions about the gas distribution and the relative optical depths in the Lyman continuum and the observed ion \citep[e.g.,][]{b:Heckman_apj11}. In this study we use a method that is complementary to other established approaches for investigating the factors that affect the escape of ionizing radiation from starbursts. We map the ionization structure of extended ionized gas in six local starbursts using emission-line ratio images. Spatial changes in emission-line ratios reveal the passage of ionizing radiation through the galaxy \citep{b:Pellegrini_apj12}. Thus, by studying the ionization structure of extended ionized gas in our sample, we can determine the optical depth of the galaxies and evaluate the likelihood of escaping Lyman continuum. \section{Method and Data} \label{s:methdata} \subsection{Method}\label{s:method} Determining the fate of ionizing radiation in starburst galaxies is a challenging observational problem. Here, we present an approach that uses the technique of ionization parameter mapping \citep[IPM;][]{b:Pellegrini_apj12}. Spatial changes to the ionization parameter, which evaluates the ionizing photon density relative to the gas density, trace the passage of ionizing radiation through the ISM. In order to track these spatial changes, we create maps of the starbursts using the [S{\sc iii}]/[S{\sc ii}]\ line ratio, which acts as a proxy for the ionization parameter. Optically thin and thick regions exhibit different ionization parameter morphologies. In regions that are optically thick, there is a transition from high to low ionization states at the interface of the ionized region and the neutral environment, whereas optically thin regions lack this transition zone \citep{b:Pellegrini_apj12}. Thus, by using ratio maps of two ions, one can distinguish between optically thick and thin regions \citep{b:Pellegrini_apj12}. Across an entire galaxy, one can thus use the ionization structure to evaluate whether and in what manner ionizing radiation escapes the galaxy \citep[e.g.,][]{b:Zastrow_apj11,b:Pellegrini_apj12}. This method has the advantage that it can be accomplished with narrow-band ground-based imaging, which makes it more observationally straightforward than other methods. \subsection{Data}\label{s:data} The galaxies selected for this study are all nearby dwarf starbursts. They have star formation rates (SFR) of 0.05--5 \mbox{${\cal M}_\odot$}/yr and stellar masses log(M$_*$) $\sim 8-10$. Since massive star feedback is thought to play a key role in the escape of Lyman continuum (LyC) photons \citep{b:Heckman_apj11}, we select galaxies that have evidence in the literature for winds and expanding bubbles. Table 1 shows the properties of our sample. We calculate the stellar masses (Column 2) using the $K$-band magnitudes from \citet{b:Skrutskie_cat03} and the mass to light ($M$/$L$) ratios following \citet{b:Bell_apjs03}, which use the galaxies' $B-V$ or $B-R$ colors to derive $M/L$. We restrict this study to starbursts within 45 Mpc, which corresponds to spatial scales of $\sim 200$ pc/arcsec, to ensure high enough spatial resolution to trace the ionization structure of the extended gas. We obtained narrow-band emission-line imaging in \Hline{\mbox{$\alpha$}}, [S{\sc ii}]$\lambda6716$, and [S{\sc iii}]$\lambda9069$\ on the nights of 2009 July 7--11 and 2010 February 12--14. The wavelengths quoted above are the rest frame wavelengths of the emission lines. We imaged the galaxies using the Maryland-Magellan Tunable Filter \citep[MMTF;][]{b:Veilleux_aj10}, which is mounted on the Inamori-Magellan Areal Camera and Spectrograph (IMACS) on the Magellan Baade telescope at Las Campanas Observatory. For every emission-line image, we obtained paired continuum exposures of the galaxies. Our observing log is presented in Table \ref{t:obs}. For each galaxy (Column 1), we list the targeted emission line and central bandpass (Columns 2, 3), exposure times (Columns 4), and the 3$\sigma$ surface brightness limit in the reduced image (Columns 5). We reduced the MMTF data using version 1.4 of the MMTF data reduction pipeline\footnote{http://www.astro.umd.edu/\textasciitilde veilleux/mmtf/datared.html}. This pipeline performs bias and flatfield corrections and then subtracts the sky using an azimuthal averaging. Next, it corrects cosmic rays and bad CCD pixels before mosaicking the different CCD chips into one image. Finally, it averages the individual emission-line and continuum images. We flux calibrated the images using standard stars LTT 1020, 6248 and 7987 \citep{b:Hamuy_pasp94}. We note that for the February 2010 observing run, our measured \Hline{\mbox{$\alpha$}}\ fluxes are a factor of 2--3 lower than the \Hline{\mbox{$\alpha$}}\ fluxes reported in the literature. A combination of shallower exposure times, combined with errors in the velocity offsets of our central bandpass, wind activity that shifts flux out the narrow MMTF bandpass, and uncertainty in the [N{\sc ii}]\ correction factors for the literature values, lead to this offset. However, a comparison with long-slit spectra confirms that the line ratios are sound. In order to cleanly subtract the continuum from the emission-line images, we first aligned the emission-line and continuum frames using {\sl \rm IRAF\/}\footnote{{\sl \rm IRAF\/}\ is distributed by NOAO, which is operated by AURA, Inc., under cooperative agreement with the National Science Foundation.} tasks {\tt wregister} and {\tt imalign}. Next, we used {\tt psfmatch} to convolve the image with the better seeing with the psf from the other. Finally, we used several bright stars in the field and the task {\tt mscimatch} to determine the proper scaling to match counts in the stars between the images. This last step was performed iteratively until we obtained a clean continuum subtraction. We present the final continuum-subtracted \Hline{\mbox{$\alpha$}}\ images and three-color composite images of the galaxies in our sample in Section \ref{s:results}. Scattered light in the instrument causes an optical artifact that is not easily corrected for and that stretches across CCD chips 3 and 5. For the most part, we avoid this issue by placing our targets on the other chips. However, for NGC 178, most of our observations have the galaxy sitting on chips 3 or 5. To mitigate the artifact, we mask the affected region in the flat field. While this increases the flux error on this galaxy, this effect is much less significant than the alteration in flux caused by the artifact on the flat. We also note that the data from the 2009 July observing run are bright time observations. As a result of scattered light in the f/2 camera and the MMTF instrument, these observations are challenging to reduce. The emission-line images of NGC~178 and NGC~7126 show many background sky features that are caused by these effects (Figure \ref{f:img_n178} and \ref{f:img_n7126}). { \begin{deluxetable}{l c c c c } \tablewidth{0pt} \tabletypesize{\footnotesize} \tablecaption{Sample\label{t:props}} \tablehead{\colhead{Galaxy} & \colhead{log($M_*$)$^a$} & \colhead{Distance} & \colhead{$L(\Hline{\mbox{$\alpha$}})$} & \colhead{Ref$^b$}\\ \colhead{ } & \colhead{[\mbox{${\cal M}_\odot$}]} & \colhead{[Mpc]} & \colhead{[erg s${-1}$]} & \colhead{ }} \startdata NGC~178 & 9.24 & 20.6 & $6.1\times 10^{40}$&(5),(8)\\ NGC~1482& 10.56 & 22.6 & $1.12\times 10^{42}$ &(2)\\ NGC~1705& 8.37 & 5.1 & $8.09\times 10^{39}$ &(3),(7) \\ NGC~3125& 9.09 & 11.5 & $4.49\times 10^{40}$&(1),(7)\\ NGC~5253& 9.05 & 3.8 & $3.71\times 10^{40}$ & (9),(7) \\ NGC~7126& 10.44 & 45.5 & $6.31\times10^{41}$ & (5),(8) \\ He~2-10 & 9.50 & 9 & $6.09\times 10^{40}$&(6),(4) \enddata \tablenotetext{a}{Stellar masses calculated using the $M/L$ ratios derived from \citep{b:Bell_apjs03}, 2MASS $K$ magnitudes \citep{b:Skrutskie_cat03}, and optical colors as follows: B-R colors from \citet{b:GildePaz_apjs03} for NGC~1705 and NGC~3125, B-V colors from \citet{b:Moustakas_apjs10} for NGC~1482, B-V colors from \citet{b:deVaucouleurs_rc391} for NGC~7126 and NGC~178, B-V colors from \citet{b:Taylor_apj05} for NGC 5253, and B-V colors from \citet{b:Ho_apjs11} for He~2-10. } \tablenotetext{b}{References listed for D and $L(\Hline{\mbox{$\alpha$}})$:(1) \citet{b:Schaerer_aaps99} (2) \citet{b:Kennicutt_pasp11} (3) \citet{b:Tosi_aj01} (4) \citet{b:Vacca_apj92} (5) \citet{b:Meurer_apjs06}, (6) \citet{b:Johnson_aj00}, (7) \citet{b:Marlowe_apj95}, (8) \citet{b:Oey_apj07}, (9) \citet{b:Sakai_apj04}} \tablenotetext{c}{Hubble flow distances for which $H_0=70 \rm km s^{-1} Mpc^{-1}$} \end{deluxetable}} { \begin{deluxetable}{l c c c c } \tablewidth{0pt} \tabletypesize{\footnotesize} \tablecaption{Table of Observations \label{t:obs}} \tablehead{\colhead{Galaxy} & \colhead{band} & \colhead{$\lambda$} & \colhead{Exp time} & \colhead{$3\sigma $} \\ \colhead{Name} & \colhead{} & \colhead{\AA} & \colhead{[s]} & \colhead{[erg/s/cm$^{2}$/\arcsec]} } \startdata \\ \multicolumn{5}{c}{$7-11\ \rm July\ 2009$}\\ \\ \hline NGC 178 &\Hline{\mbox{$\alpha$}} &6594 & $3\times1200$ & $8.46\times 10^{-17} $\\ &[S{\sc ii}] &6749 & $4\times1200$ & $1.47\times 10^{-16}$ \\ &[S{\sc iii}]&9113 & $5\times1200$ & $7.36 \times 10^{-17} $\\ & & & $1\times1800$ & \\ NGC 7126&\Hline{\mbox{$\alpha$}} &6628 & $3\times1200$ & $9.98\times 10^{-17}$ \\ &[S{\sc ii}] &6783 & $4\times1200$ & $8.25\times 10^{-17}$ \\ &[S{\sc iii}]&9159 & $4\times1200$ & $4.84\times 10^{-17}$\\ \hline \\ \multicolumn{5}{c}{$12-14\ \rm Feb\ 2010$}\\ \\ \hline NGC 1705&\Hline{\mbox{$\alpha$}} &6572$^a$ & $1\times1200$& $9.97\times 10^{-17}$ \\ &[S{\sc ii}] &6730 & $2\times1200$ & $4.10\times 10^{-17}$ \\ &[S{\sc iii}]&9087 & $3\times1200$ & $6.22\times 10^{-17}$ \\ He 2-10 &\Hline{\mbox{$\alpha$}} &6582 & $1\times1200$& $1.33\times 10^{-17}$ \\ &[S{\sc ii}] &6735 & $3\times1200$ & $7.01\times 10^{-17}$ \\ &[S{\sc iii}]&9095 & $3\times1200$ & $5.82\times 10^{-17}$ \\ NGC 3125&\Hline{\mbox{$\alpha$}} &6587 & $1\times1200$& $1.42\times 10^{-16}$ \\ &[S{\sc ii}] &6740 & $3\times1200$ & $5.02\times 10^{-17}$ \\ &[S{\sc iii}]&9102 & $3\times1200$ & $3.59\times 10^{-17}$ \\ NGC 1482&\Hline{\mbox{$\alpha$}} &6602 & $1\times1200$& $1.32\times 10^{-16}$ \\ &[S{\sc ii}] &6759 & $1\times1200$ & $1.20\times 10^{-16}$\\ &[S{\sc iii}]&9121 & $2\times1200$ & $5.36\times 10^{-17}$ \enddata \tablenotetext{a}{The appropriate wavelength for \Hline{\mbox{$\alpha$}}\ in NGC 1705 should be $\lambda6576$. However, our observation set up had the central $\lambda$ = 6572\AA } \end{deluxetable}} \section{Results}\label{s:results} In this section we present and discuss the ionization parameter maps of the galaxies individually. In section \S \ref{s:discuss}, we discuss the results from the sample as a whole. \subsection{Galaxies with Extended emission}\label{s:detections} \subsubsection{NGC 5253} NGC~5253 is a nearby dwarf galaxy currently undergoing an extreme burst of star formation. Using [S{\sc iii}]/[S{\sc ii}]\ ratio maps, \citet{b:Zastrow_apj11} discovered an optically thin ionization cone in this amorphous elliptical galaxy \citep[Figure 2 of][]{b:Zastrow_apj11}. The ionization cone is narrow with an opening angle $\sim40\degr$ and is detected out to $\sim 800$ pc from the starburst. The measured opening angle implies a solid angle for the cone that spans $\sim 3\%$ of 4$\pi$ steradians on the sky, assuming axisymmetry about the cone axis \citep{b:Zastrow_apj11}. If many starbursts have similar geometry, this suggests that an orientation bias may limit our ability to detect escaping LyC in starburst and high-redshift galaxies. \subsubsection{NGC 3125} NGC 3125 is a dwarf galaxy that has a bursty star formation history with roughly three recent epochs of star formation \citep{b:Vanzi_aap11}. This galaxy has two prominent star forming knots, knot A to the northwest and knot B to the southeast \citep{b:Vacca_apj92}; both contain massive star clusters and are the main hosts for the Wolf-Rayet population in the galaxy. The ionized gas in NGC~3125 is concentrated around the star-forming knots towards the eastern side of the galaxy relative to the overall stellar distribution (Figure \ref{f:img_n3125}). Figure \ref{f:ratios_3125} shows the emission-line ratio maps for NGC~3125. \begin{figure}[h] \centering \includegraphics[width=3.25in]{ha3125_withlines2-eps-converted-to.pdf} \includegraphics[width=3.25in]{3125-eps-converted-to.pdf} \caption{Emission-line images of NGC 3125. Top: \Hline{\mbox{$\alpha$}}. The red lines illustrate the bounds used to measure the opening angle of the cone. Bottom: Three-color composite with rest frame [S{\sc iii}]$\lambda9069$, [S{\sc ii}]$\lambda6716$, and continuum at $\lambda6680$ in red, blue, and green, respectively. At 11.5 Mpc, 10\arcsec = 560 pc. In this figure, N is up and E is to the left. \label{f:img_n3125}} \end{figure} \begin{figure}[h] \subfigure[[S{\sc iii}]/[S{\sc ii}]]{ \includegraphics[width=3in]{N3125_S3dS2_0_2_b3-eps-converted-to.pdf} \label{f:s3ds2_3125}}\\ \subfigure[[S{\sc iii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N3125_S3dha_0_05_b3-eps-converted-to.pdf} \label{f:s3dha_3125}}\\ \subfigure[[S{\sc ii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N3125_S2dha_0_05_b3-eps-converted-to.pdf} \label{f:s2dha_3125}} \caption{Emission-line ratio maps of for NGC 3125. These images have been binned 3x3, resulting in 0.6\arcsec\ pixel sizes. The maps are 50\arcmin\ on a side with an orientation such that North is up and East is to the left.} \label{f:ratios_3125} \end{figure} We discover a bipolar ionization cone in the emission-line gas extending northeast and southwest of knot B. To the northeast this ionization cone has an opening angle of $40\pm5\degr$. We obtain this measurement using the lines parallel to the edges of the cone, which are shown by the red dotted lines on the \Hline{\mbox{$\alpha$}}\ image in Figure \ref{f:img_n3125}. As can be seen in the [S{\sc iii}]/[S{\sc ii}]\ ratio map (Figure \ref{f:s3ds2_3125}), the ionization cone to the northeast of knot B exhibits high [S{\sc iii}]/[S{\sc ii}]\ throughout. The lack of transition to low values of [S{\sc iii}]/[S{\sc ii}]\ at the edge implies that the ionization cone is optically thin. This conclusion is supported by the low [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ observed throughout this portion of the cone (Figure \ref{f:s2dha_3125}), which is consistent with expectations for optically thin gas \citep{b:Pellegrini_apj12}. To the southwest, the line ratios are less conclusive. While they transition to low [S{\sc iii}]/[S{\sc ii}]\ as one moves away from knot B, there seems to be another transition to high [S{\sc iii}]/[S{\sc ii}]\ near the outer edge. However, the line ratios are less certain in this region. In Figure \ref{f:n3125_quant}, we quantitatively compare the line ratios in the ionization cone to those of the rest of the galaxy. Figure \ref{f:n3125_quant} shows the line ratios from the binned ratio maps (Figure \ref{f:ratios_3125}) as a function of $z$, the projected distance above and below the major axis of the galaxy. The red points represent the line ratios from the ionization cone, selected from the region between the white lines shown in the bottom panel of Figure \ref{f:n3125_quant}. The blue points are the mean line ratios from the rest of the galaxy at that $z$. The error bars on the blue points represent either the standard deviation of the line ratios at that $z$ or the error on the mean, whichever is more. All points originate from pixels that have a minimum 2$\sigma$ detection in all three emission-line bands in the 3$\times$3 binned images. From the top and middle plots in Figure \ref{f:n3125_quant}, we see a clear excess in [S{\sc iii}]/[S{\sc ii}]\ and a deficiency of [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ in the NE side of the cone. This behavior is consistent with expectations for optically thin gas. The trends in the line ratios discussed above are robust, even after considering the effects of internal reddening on the ratio maps. The lines shown on the upper right of Figure \ref{f:n3125_quant} indicate how the line ratios would change with different levels of reddening. In order to explain the observed [S{\sc iii}]/[S{\sc ii}]\ excess, the reddening would have to have $A_V \sim 5$. We note that the measured reddening toward knots A and B are $E(B-V) = 0.24$\ and 0.21, respectively \citep{b:Hadfield_mnras06}, which correspond roughly to $A_V \sim 1$. Based on polarimetric observations most of the dust is located in the plane of the galaxy \citep{b:Alton_mnras94}. Thus it is reasonable to expect that the extinction in the cone is much less than that toward knots A and B. Furthermore, while strong reddening would enhance the [S{\sc iii}]/[S{\sc ii}]\ ratio, correcting for it would drive the [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ ratio even lower, which strengthens the argument for optically thin gas. \citet{b:Marlowe_apj95} study the kinematics of the \Hline{\mbox{$\alpha$}}\ gas in NGC~3125 using echelle spectra. They detect a Doppler ellipse along the ionization cone to the northeast, showing that it is a bubble expanding with $v = 50$\ km/s. This detected bubble is consistent with the idea that massive star feedback clears low-density passages in the ISM through which ionizing radiation may escape. To add further support to this picture, low-resolution X-ray observations show that the X-ray distribution is elongated along the minor axis of the galaxy \citep{b:Fabbiano_apjs92}, suggestive of feedback related activity. Furthermore, the ionization cone occurs at the edge of the galaxy, which may make it easier for ionizing radiation to escape \citep{b:Gnedin_apj08,b:Paardekooper_aap11}. The kinematics towards the southwest are less clear, and \citet{b:Marlowe_apj95} do not detect a Doppler ellipse in this region. We also see ionized gas extending to the northeast of knot A. Figure \ref{f:s3ds2_3125} shows that while the gas near the cluster has high ionization parameter, it clearly transitions to a low ionization parameter as the distance from the cluster increases. The transition indicates that the galaxy is optically thick in that direction. The line ratio comparison in Figure \ref{f:n3125_quant} supports this analysis. \begin{figure}[ht] \centering \includegraphics[width=3.45in]{N3125_S3dS2_ratiovsheight_sig2bin3-eps-converted-to.pdf} \includegraphics[width=3.45in]{N3125_S2dHa_ratiovsheight_sig2bin3-eps-converted-to.pdf} \includegraphics[width=2.5in]{3125_withlines-eps-converted-to.pdf} \caption{Line ratios as a function of projected distance, $z$, above and below the major axis of NGC 3125 for the ionization cone (red points) and the mean line ratio at that $z$ from rest of the galaxy (blue points). \textit{Top}: [S{\sc iii}]/[S{\sc ii}]. The lines shown on the upper right indicate how the line ratios would change due to reddening of A$_V$ = 1, 2, 5, and 10. All points represent pixels that have a minimum 2$\sigma$ detection in all three emission-line bands. Negative values are towards the northeast, positive values toward the southwest. \textit{Middle}: Same as the top plot but for [S{\sc ii}]/\Hline{\mbox{$\alpha$}}. \textit{Bottom}: 3-color emission-line image from Figure \ref{f:img_n3125} with white lines delineating the region from which the red points are selected. This Figure demonstrates that there is a clear excess of [S{\sc iii}]/[S{\sc ii}]\ and a deficiency of [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ in the NE side of the cone.\label{f:n3125_quant}} \end{figure} As discussed in \S \ref{s:data}, our \Hline{\mbox{$\alpha$}}\ observations are missing a significant amount of flux due to a combination of wavelength calibration uncertainties, velocity offsets, and shallow integration. For this reason, we use the \Hline{\mbox{$\alpha$}}\ observations of \citet{b:GildePaz_apjs03}, to measure the \Hline{\mbox{$\alpha$}}\ flux in the ionization cone. After correcting for an expected 10\% contribution to the flux from the nearby [N{\sc ii}]\ lines, we obtain $F(\Hline{\mbox{$\alpha$}})= 2.0\ \pm 0.6\ \times 10^{-13}$\ erg s$^{-1}$ cm$^{-2}$, which corresponds to a rate of ionizing photons of log(\mbox{$Q_0$})=51.02. This means that the equivalent of at least 100 ($\pm 30$) O7 stars ionize the cone, assuming \mbox{$Q_0$} = $10^{49}$ for an O7 star at the metallicity of NGC~3125 \citep{b:Smith_mnras02}. Knot B has the equivalent of $\sim 1300$ O7 stars based on \Hline{\mbox{$\alpha$}}\ flux \citep[][adjusted to the same \mbox{$Q_0$}\ per O7 star]{b:Hadfield_mnras06}. Based on these numbers, roughly 8\% of the ionizing radiation generated by the cluster needs to escape into the NE cone. The ionization cone in NGC~3125 is narrow and appears somewhat collimated, with an opening angle of $40\pm5\degr$. This angle corresponds to a solid angle of $3\pm1$\% of 4$\pi$ steradians, assuming axisymmetry about the cone axis. We note that this value is not inconsistent with the 8\% escape fraction estimated from the \Hline{\mbox{$\alpha$}}\ flux, above. Since knot B is smaller than the base of the cone, the covering fraction of the cone relative to the cluster is larger than that of the cone on the sky. We note that the 8\% quoted here only refers to the cluster in knot B. Using the total $F$(\Hline{\mbox{$\alpha$}})=3.4$\times 10^{-12}$ erg s$^{-1}$ cm$^{-2}$ \citep{b:Marlowe_apj95}, the cone represents $\sim6\%$ of the ionizing flux of the galaxy. The presence of an optically thin ionization cone suggests that ionizing radiation is escaping this starburst. At the very least, we see that ionizing radiation is escaping the main body of the galaxy and traveling into the halo. The small implied solid angle, in conjunction with a preferred direction is similar to the geometry for NGC 5253, again supporting the possibility that an orientation bias affects our ability to detect escaping Lyman continuum in starbursts and LBGs. \subsubsection{NGC 1705} NGC 1705 is a nearby starburst galaxy with a stunning starburst-driven galactic wind \citep{b:Meurer_aj92}. NGC 1705 has a combined stellar and gas mass of $\sim2.7\times10^8$\mbox{${\cal M}_\odot$}\ \citep{b:Meurer_aj92}. The metallicity measurements range from 12+log(O/H) = 8.21 to 8.46 \citep{b:Lee_apj04,b:Storchi-Bergmann_apj94,b:Meurer_aj92}. Like similar dwarf starbursts, NGC 1705 has a bursty star formation history with two recent star forming epochs in the last 15 Myr \citep{b:Annibali_aj03,b:Annibali_aj09}. One of these bursts formed the well known super-star cluster (SSC) at the center, which has log($M$)$\sim 10^5$ and an age of $\sim 10$\ Myr \citep{b:Melnick_aap85,b:OConnell_apj94,b:Ho_apj96}. This SSC generates most of the UV luminosity and is consistent with a population dominated by B stars \citep{b:Heckman_aj97}. That same 10--15 Myr old burst of star formation is thought to have launched the galactic wind \citep{b:Annibali_aj03,b:Annibali_aj09}. This burst was followed by a more recent epoch of star formation, $\sim 3$\ Myr ago that may have been triggered by the wind \citep{b:Annibali_aj03,b:Annibali_aj09}. \begin{figure}[h] \centering \includegraphics[width=3.45in]{ha1705-eps-converted-to.pdf} \includegraphics[width=3.25in]{1705-eps-converted-to.pdf} \caption{Emission-line images of NGC 1705. Top: \Hline{\mbox{$\alpha$}}\ Bottom: Three-color composite with rest frame [S{\sc iii}]$\lambda9069$, [S{\sc ii}]$\lambda6716$, and continuum at $\lambda6680$ in red, blue, and green, respectively. At 5.1 Mpc, 10\arcsec = 250 pc. In this figure, N is up and E is to the left. \label{f:img_n1705}} \end{figure} \begin{figure}[h] \subfigure[[S{\sc iii}]/[S{\sc ii}]]{ \includegraphics[width=3in]{N1705_S3dS2_0_2_b3-eps-converted-to.pdf} \label{f:s3ds2_1705}}\\ \subfigure[[S{\sc iii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N1705_S3dha_0_05_b3-eps-converted-to.pdf} \label{f:s3dha_1705}}\\ \subfigure[[S{\sc ii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N1705_S2dha_0_05_b3-eps-converted-to.pdf} \label{f:s2dha_1705}} \caption{Emission-line ratio maps of for NGC 1705. These images have been binned 3x3, resulting in 0.6\arcsec\ pixel sizes. The maps are 62.5\arcmin\ on a side with an orientation such that North is up and East is to the left.} \label{f:ratios_1705} \end{figure} In emission-line gas, NGC~1705 shows filamentary extended substructure as a result of its galactic wind \citep{b:Meurer_aj92}. Based on \Hline{\mbox{$\alpha$}}\ observations, the superbubble has an expansion velocity 50--132 km/s \citep{b:Marlowe_apj95,b:Meurer_aj92}. \citet{b:Heckman_aj97} studied the kinematics using UV interstellar absorption lines and found further evidence for an expanding wind. X-ray observations show that the wind has similar size and morphology to the optically detected gas \citep{b:Strickland_apjs04}. The star-forming nucleus resides on the axis of symmetry for the superbubble, which further supports that the wind is driven by star formation \citep{b:Meurer_aj92}. Shocks contribute only a small amount to the ionization in the wind, as most of the emission-line gas is photo-ionized \citep{b:Marlowe_apj95,b:Veilleux_aj03}. Our emission-line images and ratio maps for NGC~1705 are shown in Figures \ref{f:img_n1705}, and \ref{f:ratios_1705}. The ionized gas associated with the wind is clearly detected in both [S{\sc iii}]\ and [S{\sc ii}]. The spatial changes in the [S{\sc iii}]/[S{\sc ii}]\ ratio map suggest that the wind in NGC~1705 is optically thick. Close to the starburst, the [S{\sc iii}]/[S{\sc ii}]\ ratio is high, as expected from photoionized gas near the ionization source. However, as one looks further into the wind, there is a clear transition towards lower [S{\sc iii}]/[S{\sc ii}]\ ratios, and at the far edge of the shells, the line ratios indicate the [S{\sc ii}]\ emission is dominant. This is further supported by the [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ and [S{\sc iii}]/\Hline{\mbox{$\alpha$}}\ ratio maps in Figures \ref{f:s2dha_1705} and \ref{f:s3dha_1705}. In these maps, the [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ ratio increases at the edges of the filament, and, as expected based on Figure \ref{f:s3ds2_1705}, the [S{\sc iii}]/\Hline{\mbox{$\alpha$}}\ ratio decreases. While this behavior is consistent with optically thick gas, we note that if the radiation field is dominated by hot stars with softer ionizing spectra, as in the case of NGC~1705, the gas may simultaneously exhibit a low-ionization transition zone and be optically thin \citep{b:Pellegrini_apj12}. In addition to evaluating the optical depth in the plane of the sky, the ratio maps can be used to evaluate the optical depth along the line of sight. For gas with abundances similar to that of the LMC, [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ $< 0.05$ is a diagnostic for optically thin photoionized gas along the line of sight. Of particular interest is the region around NGC~1705-1, the super star cluster. \citet{b:Heckman_aj97} estimated a very low absorbing {\sc \rm HI}\ column, $N_{\sc \rm HI} = 1.5 \times 10^{20} \rm cm^{-2}$, toward NGC~1705-1. This column is an order of magnitude smaller than the column measured from 21 cm observations, which suggests that the cluster is either sitting in front of most of the {\sc \rm HI}\ or that we are observing the SSC through a hole in the ISM \citep{b:Heckman_aj97}. Either scenario increases the likelihood for escaping Lyman continuum. However, Figure \ref{f:s2dha_1705} shows that [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ $> 0.05$ over the entire nuclear region, which indicates that the nuclear region is optically thick in the line of sight. The low {\sc \rm HI}\ column combined with optical [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ implies that even though the neutral gas may have large variations in column density, the majority of sight lines are optically thick. It is expected that strong starburst feedback, such as that present in NGC~1705, is conducive to escaping Lyman continuum. However, our observations indicate that the wind in NGC 1705 is optically thick. Both the strength of the radiation field and the amount of gas available to absorb the ionizing radiation will determine whether the galactic wind is optically thick or thin. While there have been two recent epochs of star formation, one 10-15 Myr ago and one $\sim 3$ Myr ago \citep{b:Annibali_aj03,b:Annibali_aj09}, the ionizing population is dominated by B stars \citep{b:Heckman_aj97}. This means that there are relatively fewer ionizing photons being emitted and the radiation field is softer than if there were significant numbers of O stars. In addition to a reduced radiation field, a `spur' of {\sc \rm HI}\ gas is associated with the outflow \citep{b:Meurer_mnras98}. This spur of neutral gas is likely co-spatial with the optically detected galactic wind based on its kinematics and orientation \citep{b:Meurer_mnras98,b:Elson_mnras13}. The feature has an {\sc \rm HI}\ column density of $\sim10^{20}\rm\ cm^{-2}$, which is more than high enough to be optically thick to Lyman continuum. Thus, despite the presence of a strong galactic wind, NGC~1705 is likely optically thick due to a combination of population age and gas morphology. \subsubsection{He 2-10} He 2-10 is one of the prototypical HII galaxies \citep{b:Allen_mnras76}. It is 9 Mpc away \citep{b:Vacca_apj92} and has dynamical mass $\sim3 \times 10^9$ \mbox{${\cal M}_\odot$} \citep{b:Kobulnicky_aj95}. In emission-line gas there are many notable loops and filaments (Figure \ref{f:img_he210}) whose kinematics suggest expanding bubbles and possible outflows \citep{b:Mendez_aap99}. These bubbles are centered on sites of recent star formation. He 2-10 is the host of nearly 80 SSCs \citep{b:Johnson_aj00}. The central starburst contains a handful of clusters in ultra-compact H{\sc ii}\ regions that contain the ionizing equivalent of 700-2600 O7 stars each \citep{b:Johnson_apj03}. In addition to prodigiously forming stars, SFR $\sim$ 0.87 - 2 \mbox{${\cal M}_\odot$}/yr \citep[][]{b:Calzetti_apj10,b:Reines_nature11}, this galaxy was recently found to be the host of an AGN \citep[][]{b:Reines_nature11}. \begin{figure}[h] \centering \includegraphics[width=3.35in]{haHe210_wcircles-eps-converted-to.pdf} \includegraphics[width=3.25in]{He210-eps-converted-to.pdf} \caption{Emission-line images of He 2-10. Top: \Hline{\mbox{$\alpha$}}. The two circles correspond to regions that have [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ $<0.05$. Bottom: Three-color composite with rest frame [S{\sc iii}]$\lambda9069$, [S{\sc ii}]$\lambda6716$, and continuum at $\lambda6680$ in red, blue, and green, respectively. At 9 Mpc, 10\arcsec = 460 pc. In this figure, N is up and E is to the left. \label{f:img_he210}} \end{figure} \begin{figure}[h] \subfigure[[S{\sc iii}]/[S{\sc ii}]]{ \includegraphics[width=3in]{He210_S3dS2_0_2_b3-eps-converted-to.pdf} \label{f:s3ds2_he210}}\\ \subfigure[[S{\sc iii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{He210_S3dha_0_05_b3-eps-converted-to.pdf} \label{f:s3dha_he210}}\\ \subfigure[[S{\sc ii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{He210_S2dha_0_05_b3-eps-converted-to.pdf} \label{f:s2dha_he210}} \caption{Emission-line ratio maps of for He~2-10. These images have been binned 3x3, resulting in 0.6\arcsec\ pixel sizes. The maps are 39.58\arcmin\ on a side with an orientation such that North is up and East is to the left.} \label{f:ratios_he210} \end{figure} In the [S{\sc iii}]/[S{\sc ii}]\ ratio map (Figure \ref{f:s3ds2_he210}), the central star forming region stands out with high [S{\sc iii}]/[S{\sc ii}], as would be expected for a region hosting thousands of O stars. As the radius from the center increases, [S{\sc iii}]/[S{\sc ii}]\ drops considerably. There is a local maximum $\sim 8\arcsec$\ to the east around the location of the second SSC. However, even here the emission-line gas transitions to low [S{\sc iii}]/[S{\sc ii}]\ toward the edge of the expanding bubble. The contrast at the edges of the bubbles is seen more clearly in the [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ ratio map (Figure \ref{f:s2dha_he210}), where a clear enhancement of [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ is present. In contrast the [S{\sc iii}]/\Hline{\mbox{$\alpha$}}\ ratio (Figure \ref{f:s3dha_he210}) falls off as the distance from the ionizing source increases, as would be expected for an optically thick nebula. The spatial changes in the line ratio maps are consistent with expanding superbubbles. High ionization parameter gas is present close to the ionizing source, while the edges of the bubbles have low ionization parameter and possible contribution to [S{\sc ii}]\ from shocks. Studies of the \Hline{\mbox{$\alpha$}}\ kinematics indicate a potential Doppler ellipse in the NE bubble and velocities differences of up to $\pm300$ km/s across the galaxy \citep{b:Mendez_aap99}. Using UV interstellar lines, \citet{b:Johnson_aj00} determined that the bulk motion of the ISM indicates an outflow with $v\sim360$ km/s. The velocities reported by both works are in excess of the escape velocity $\sim160\pm30$\ km/s \citep{b:Johnson_aj00}. However, the closed loops of the outflow seen in the \Hline{\mbox{$\alpha$}}\ images suggest that these bubbles have not yet broken out of the ISM to form a galactic fountain. Bubble walls severely hinder the passage of ionizing radiation, and can delay the escape of Lyman continuum radiation until $Q_0$ has dropped considerably \citep{b:Dove_apj00,b:Fujita_apj03}. He 2-10 is unique in our sample, in that it is the only galaxy that contains a confirmed AGN. This AGN was discovered by \citet{b:Reines_nature11} as a radio and X-ray point source and is located near the star forming region of knot A. The accretion rate onto the AGN is modest, but may have been greater in the past \citep{b:Reines_nature11}. Thus, the presumed tracers for massive-star feedback, such as the large expanding bubbles, may be contaminated by contributions from the AGN as opposed to being driven by star formation. The distribution of gas in the ISM plays a critical role in the passage of ionizing radiation in a galaxy. The {\sc \rm HI}\ observations from \citet{b:Kobulnicky_aj95} show that the neutral ISM extends significantly beyond the optically detected ionized gas. This observation is consistent with the observations of optically thick bubbles. However, if this overall {\sc \rm HI}\ envelope has a clumpy distribution, it is still possible to have small low-density passages through which ionizing radiation may escape the starburst. The [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ ratio map shows two locations that have [S{\sc ii}]/\Hline{\mbox{$\alpha$}} $< 0.05$ and may be optically thin. We note these locations with small circles on Figure \ref{f:img_he210}. One of these regions coincides with the location of the AGN, and the other lies 4\arcsec\ to the SE from the first. While our data cannot establish the presence of an optically thin path from the galaxy conclusively, these observations are suggestive that feedback from the AGN or star formation in the region have cleared out a small hole in the ISM. The variable reddening measurements derived from different wavelengths and positions in the galaxy further suggest an inhomogeneous distribution of gas and dust throughout the galaxy \citep[e.g.,][]{b:Sauvage_aap97}. The presence of small local holes is also supported by the fact that many of the probably young SSCs do not show commensurate \Hline{\mbox{$\alpha$}}\ emission, which suggests that gas has been evacuated in the regions of these clusters \citep{b:Johnson_aj00}. If this is the case, it supports the picture that ionizing radiation escapes galaxies through small holes and that an orientation bias affects our ability to detect escaping Lyman continuum. We note that these line ratios have not been corrected for reddening. However, any reddening correction would drive the [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ line ratio lower, strengthening the argument for optically thin gas along the line of sight. \subsubsection{NGC 1482} NGC 1482 is an SA0 \citep{b:deVaucouleurs_rc391} with a well known galactic wind \citep{b:Hameed_aj99}. At a distance of 22.6 Mpc \citep{b:Kennicutt_pasp11}, this starburst is a highly inclined disk with a prominent dust lane \citep{b:deVaucouleurs_rc391}. The wind was first discovered by \citep{b:Hameed_aj99} in optical emission-lines, with later optical and X-ray observations confirming its nature as a wind \citep[][]{b:Veilleux_apj02,b:Strickland_apjs04}. \begin{figure}[h] \centering \includegraphics[width=3.45in]{ha1482-eps-converted-to.pdf} \includegraphics[width=3.25in]{1482-eps-converted-to.pdf} \caption{Emission-line images of NGC 1482. Top: \Hline{\mbox{$\alpha$}}\ Bottom: Three-color composite with rest frame [S{\sc iii}]$\lambda9069$, [S{\sc ii}]$\lambda6716$, and continuum at $\lambda6680$ in red, blue, and green, respectively. At 22.6 Mpc, 10\arcsec = 1.1 kpc. In this figure, N is up and E is to the left. \label{f:img_n1482}} \end{figure} \begin{figure}[h] \subfigure[[S{\sc iii}]/[S{\sc ii}]]{ \includegraphics[width=3in]{N1482_S3dS2_0_2_b3-eps-converted-to.pdf} \label{f:s3ds2_1482}}\\ \subfigure[[S{\sc iii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N1482_S3dha_0_05_b3-eps-converted-to.pdf} \label{f:s3dha_1482}}\\ \subfigure[[S{\sc ii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N1482_S2dha_0_05_b3-eps-converted-to.pdf} \label{f:s2dha_1482}} \caption{Emission-line ratio maps of for NGC~1482. These images have been binned 3x3, resulting in 0.6\arcsec\ pixel sizes. The maps are 41.67\arcmin\ on a side with an orientation such that North is up and East is to the left.} \label{f:ratios_1482} \end{figure} In our emission-line images (Figure \ref{f:img_n1482}) and ratio maps (Figure \ref{f:ratios_1482}), we primarily detect the wind in [S{\sc ii}]\ and \Hline{\mbox{$\alpha$}}. The [S{\sc iii}]\ emission is confined to the star forming region in the inner disk. The ionization parameter map shows strong evidence for radiation bounding in NGC~1482. High [S{\sc iii}]/[S{\sc ii}]\ traces the disk star forming region, and the line ratio drops precipitously as one moves further into the wind (Figure \ref{f:s3ds2_1482}). \citet{b:Veilleux_aj03} found the wind to be dominated by shock ionization based on [N{\sc ii}]/\Hline{\mbox{$\alpha$}}\ observations. Our [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ map (Figure \ref{f:s2dha_1482}) confirms this picture with [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ $\geq 0.50$ in the wind. In reality, the emission observed in the wind will be a combination of shock- and photoionized gas. In order to put strong constraints on the [S{\sc iii}]/[S{\sc ii}]\ line ratio in the wind, we explore correcting the line ratios for the shock ionization as follows. As discussed by \citet{b:Jaskot_apj13}, the observed line ratio can be written as: \begin{equation} \frac{A_o}{B_o} = \frac{A_s + A_p}{B_s + B_p}\label{eq:photo} \end{equation} where A and B correspond to the emission lines in question, and the subscripts $o$, $s$, and $p$ indicate the observed, shocked, and photoionized line strengths, respectively. To convert our observed line ratios to the line ratios of photoionized gas, we need to solve for the fraction of shock-ionized relative to photoionized gas, \begin{equation} X = \frac{H_s}{H_p}. \end{equation} Together with Equation~\ref{eq:photo}, this yields: \begin{equation} \frac{A_p}{B_p} = \frac{A_o}{B_o} + X\left(\frac{A_o}{B_o} - \frac{A_s}{B_s}\right) \label{eq:step3} \end{equation}. In Equation \ref{eq:step3}, $\frac{A_s}{B_s}$ can be replaced by the predicted line strengths from MAPPINGS III shock models \citep{b:Allen_apjs08}. Since \citet{b:Veilleux_apj02} observed velocities of up to $v = 250$ km/s in the wind, we explore shocks with $v \leq 950$ km/s (assuming strong shocks). For this calculation we use the MAPPINGS III shock-only models, which provide the most optimistic correction factor for high [S{\sc iii}]/[S{\sc ii}]\ from photoionization. The precursor component to the shock would increase the relative contribution of shocks to the [S{\sc iii}], thus moving the corrected line ratios more towards optically thick than the shock only models. To solve for $X$, we write Equation \ref{eq:step3} for each emission line ratio, [S{\sc iii}]/[S{\sc ii}], [S{\sc iii}]/\Hline{\mbox{$\alpha$}}, and [S{\sc ii}]/\Hline{\mbox{$\alpha$}}. By combining the resultant equations, we finally solve for $X$: \begin{equation} X = \rm \frac{\it{D_1}\left(\frac{[SII]}{\Hline{\mbox{$\alpha$}}}_o\right) + \it{D_2}\left(\frac{[SIII]}{[SII]}_o\right) + \it{D_3}}{\it{D_3D_2}} \end{equation} where \begin{equation} D_1 = \rm \frac{[SIII]}{[SII]}_o - \frac{[SIII]}{[SII]}_s\\ \end{equation} and $D_2$ and $D_3$ are the corresponding values for [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ and [S{\sc iii}]/\Hline{\mbox{$\alpha$}}, respectively. After solving for $X$, we use that value in Equation \ref{eq:step3} to create a new [S{\sc iii}]/[S{\sc ii}]\ map (Figure \ref{f:n1482_shock}). We find that the [S{\sc iii}]/[S{\sc ii}]\ maps remain strongly radiation bounded. \begin{figure} \includegraphics[width=4in]{N1482_S3dS2_shockcor_cut-eps-converted-to.pdf} \caption{Shock-corrected ionization parameter map of [S{\sc iii}]/[S{\sc ii}]. This map is created assuming shock velocities of 950 km/s and the MAPPINGS III shock-only model \citep{b:Allen_apjs08}. Even in the most optimistic correction, the [S{\sc iii}]/[S{\sc ii}]\ ratio transitions to values $<0.70$ close to the disk of the galaxy. N is up and E is to the left. \label{f:n1482_shock}} \end{figure} Since there is evidence that the wind has cleared gas from the ISM of NGC~1482, it is somewhat surprising that this object does not show evidence for an optically thin wind. {\sc \rm HI}\ observations reveal a hole in the {\sc \rm HI}\ distribution at the location of the wind \citep{b:Hota_mnras05,b:Omar_japa05}. Further evidence that the wind has cleared material from the center of the galaxy is found by \citet{b:Sharp_apj10}. They generate a [S{\sc ii}]$\lambda6716/\lambda6732$ ratio map and see a sharp gradient at the base of the wind, which suggests clearing of the inner disk by the wind \citep{b:Sharp_apj10}. Both of these lines of evidence suggest that ionizing radiation can escape along the wind. Perhaps the massive star population in NGC~1482 has aged in the time since the wind launched, and, thus, $Q_0$ has dropped significantly. Another explanation could be that dust entrained in the wind \citep{b:Vagshette_NewA12} provides additional material that absorbs radiation as it travels away from the galaxy. We note here that some of the emission may be shifted out of our very narrow bandpasses. For $v\sim250$\ km/s, the emission will shift by 5\AA\ and 7\AA\ for [S{\sc ii}]\ and [S{\sc iii}], respectively. This is comparable to half the bandpass in each of those filters. Exacerbating this effect, our central bandpass in [S{\sc iii}]\ is shifted $\sim4\AA$\ from the appropriate central bandpass for the redshift of NGC~1482. This means that for NGC~1482 we do not detect some of the redshifted emission from [S{\sc iii}]\ in the wind, although we should detect all of the blueshifted component. That being said, the wind exhibits the transition from high to low ionization parameter gas relatively close to the disk of the galaxy. Furthermore, the [S{\sc iii}]/[S{\sc ii}]\ line ratio is well below 0.5 for most of the wind. Even if we assume that we are missing half of the [S{\sc iii}]\ flux in the wind because of the velocity offset, the difference does not push the [S{\sc iii}]/[S{\sc ii}]\ ratio high enough to match our expectations for optically thin emission. Given the 5 kpc extent of the wind \citep{b:Strickland_apjs04}, the location of the transition zone makes it less likely that the transition is a spurious one. \subsection{Non-Detections}\label{s:nondetect} \subsubsection{NGC 178} NGC 178 is forming stars at a rate of 0.55 \mbox{${\cal M}_\odot$}/yr \citep{b:Oey_apj07} and lies 20.6 Mpc away \citep{b:Meurer_apjs06}. In the emission-line images we see that ionized gas is split between two areas, one around the bulk of the observed continuum, and a separate extra-planar region (Figure \ref{f:img_n178}). Since ionizing radiation emitted at the edge of galaxies is more likely to escape \citep{b:Gnedin_apj08}, it might be reasonable to expect to find some evidence of density bounding in this region of the ionization parameter map. \begin{figure}[h] \centering \includegraphics[width=3.45in]{ha178-eps-converted-to.pdf} \includegraphics[width=3.25in]{178-eps-converted-to.pdf} \caption{Emission-line images of NGC 178. Top: \Hline{\mbox{$\alpha$}}\ Bottom: Three-color composite with rest frame [S{\sc iii}]$\lambda9069$, [S{\sc ii}]$\lambda6716$, and continuum at $\lambda6680$ in red, blue, and green, respectively. At 20.6 Mpc, 10\arcsec = 1 kpc. In this figure, N is up and E is to the left. \label{f:img_n178}} \end{figure} However, the ionization parameter map shows that all nebular regions are optically thick. The ionized gas to the west is comprised of a few separate peaks in the [S{\sc iii}]/[S{\sc ii}]\ ratio map (Figure \ref{f:ratios_178}), all of which exhibit a clear transition to low ionization parameter gas at the edges. \begin{figure}[h] \subfigure[[S{\sc iii}]/[S{\sc ii}]]{ \includegraphics[width=3in]{N178_S3dS2_0_2_b3-eps-converted-to.pdf} \label{f:s3ds2_178}}\\ \subfigure[[S{\sc iii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N178_S3dha_0_05_b3-eps-converted-to.pdf} \label{f:s3dha_178}}\\ \subfigure[[S{\sc ii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N178_S2dha_0_05_b3-eps-converted-to.pdf} \label{f:s2dha_178}} \caption{Emission-line ratio maps of for NGC~178. These images have been binned 3x3, resulting in 0.6\arcsec\ pixel sizes. The maps are 50\arcmin\ on a side with an orientation such that North is up and East is to the left.} \label{f:ratios_178} \end{figure} The [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ map (Figure \ref{f:ratios_178}) is intriguing. There appears to be an envelope of high [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ surrounding the galaxy. This envelope could be caused by the well known [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ gradient in the warm, ionized medium of galaxies \citep{b:Rand_apj98} or could be a signature of shocked gas surrounding the galaxy. We note, however, that there are significant sky gradients and background features around the galaxy that are caused by scattered light within the instrument and by bright time observing conditions (\S \ref{s:data}). These sky features make it challenging to determine the nature of the surrounding low ionization parameter envelope. \subsubsection{NGC 7126} NGC 7126 is an SA(rs)c galaxy \citep{b:deVaucouleurs_rc391}. At 45.5 Mpc away \citep{b:Meurer_apjs06}, it is the most distant galaxy in our sample. The emission-line images reveal an extended spiral structure filled with H{\sc ii}\ regions (Figure \ref{f:img_n7126}). {\sc \rm HI}\ observations show that it is interacting with its nearby companion NGC~7125 \citep{b:Nordgren_aj97}. NGC~7126 does not have a catalogued wind, and, with its high star formation rate, acts as a control against this parameter in the sample \citep{b:Hanish_apj10}. \begin{figure}[h] \centering \includegraphics[width=3.45in]{ha7126_2-eps-converted-to.pdf} \includegraphics[width=3.25in]{7126-eps-converted-to.pdf} \caption{Emission-line images of NGC 7126. Top: \Hline{\mbox{$\alpha$}}\ Bottom: Three-color composite with rest frame [S{\sc iii}]$\lambda9069$, [S{\sc ii}]$\lambda6716$, and continuum at $\lambda6680$ in red, blue, and green, respectively. At 45.5 Mpc, 10\arcsec = 2.3 kpc. In this figure, N is up and E is to the left. \label{f:img_n7126}} \end{figure} \begin{figure}[h] \subfigure[[S{\sc iii}]/[S{\sc ii}]]{ \includegraphics[width=3in]{N7126_S3dS2_0_2_b3-eps-converted-to.pdf} \label{f:s3ds2_7126}}\\ \subfigure[[S{\sc iii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N7126_S3dha_0_05_b3-eps-converted-to.pdf} \label{f:s3dha_7126}}\\ \subfigure[[S{\sc ii}]/\Hline{\mbox{$\alpha$}}]{ \includegraphics[width=3in]{N7126_S2dha_0_05_b3-eps-converted-to.pdf} \label{f:s2dha_7126}} \caption{Emission-line ratio maps of for NGC~7126. These images have been binned 3x3, resulting in 0.6\arcsec\ pixel sizes. The maps are 83.33\arcmin\ on a side with an orientation such that North is up and East is to the left.} \label{f:ratios_7126} \end{figure} The emission-line ratio maps are shown in Figure \ref{f:ratios_7126}. The [S{\sc iii}]/[S{\sc ii}]\ map shows a remarkably low ratio of [S{\sc iii}]/[S{\sc ii}]. While the spiral structure is detected, it does not show up as clearly as it does in the \Hline{\mbox{$\alpha$}}\ ratio maps. In the [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ ratio map, the H{\sc ii}\ regions themselves appear with line ratios consistent with photoionized gas. As one moves to the edges of the individual regions, [S{\sc ii}]/\Hline{\mbox{$\alpha$}}\ increases. This is consistent with expectations of the warm, ionized medium in galaxies \citep{b:Rand_apj98}. We note, however, that these line ratios are a bit high, and they approach values consistent with shock ionization. The [S{\sc iii}]/\Hline{\mbox{$\alpha$}}\ map exhibits relatively low line ratios throughout, even at the centers of the observed H{\sc ii}\ regions. This could be indicative that the ionizing stellar population is composed primarily of B or late O type stars. \section{Discussion}\label{s:discuss} The full sample of seven galaxies provides examples of both optically thin and optically thick extended emission-line gas. By comparing the properties of these galaxies we can gain insight on the processes and properties that influence the escape of ionizing radiation. Table \ref{t:SFR} presents relevant properties for each galaxy in the sample. Columns 1 and 2 are the galaxy name and {\sc \rm HI}\ mass, respectively; Columns 3, 4, and 5 present the \Hline{\mbox{$\alpha$}}\ derived SFR, SFR surface density (\mbox{$\Sigma_{\rm SFR}$}), and the specific SFR (sSFR), respectively. Column 6 provides the references for the star formation rates, {\sc \rm HI}\ masses and \Hline{\mbox{$\alpha$}}\ radii. Finally, Column 7 indicates whether the galaxy shows wind and/or expanding bubble activity, and Column 8 shows the ratio of {\sc \rm HI}\ mass to stellar mass (Table \ref{t:props}). \begin{deluxetable*}{l c c c c c c c} \tablewidth{0pt} \tabletypesize{\footnotesize} \tablecaption{Star Formation Rates \label{t:SFR}} \tablehead{\colhead{Galaxy} & \colhead{log($M_{{\sc \rm HI}}$)} & \colhead{SFR} & \colhead{\mbox{$\Sigma_{\rm SFR}$}$^a$} & \colhead{sSFR} &\colhead{Ref$^b$} & \colhead{Wind or} & \colhead{M$_{{\sc \rm HI}}$/M$_*^c$} \\ \colhead{} & \colhead{[\mbox{${\cal M}_\odot$}]} &\colhead{[\mbox{${\cal M}_\odot$}/yr]} &\colhead{[$\frac{\mbox{${\cal M}_\odot$}}{\rm yr\ kpc^2}$]} & \colhead{Gyr$^{-1}$} & \colhead{} & \colhead{Bubbles?}&\colhead{} } % \startdata \multicolumn{7}{c}{Optically thin extended emission}\\ \hline\\ NGC~3125 &\nodata&0.461 &0.0826& 0.374& (1),n/a, (2) & Y (3) &\nodata\\ NGC~5253 &8.22 & 0.428& 1.04 & 0.382& (1),(4),(5) & Y (3) &0.006\\ \hline\\ \multicolumn{7}{c}{Optically thick extended emission}\\ \hline\\ NGC~1482 & 9.13&3.39 & 0.0271& 0.093& (1),(6),(7) & Y (8) & 0.037 \\ NGC~1705 & 7.96 &0.057& 0.0091& 0.241& (1),(4),(9,7)& Y (10) & 0.389 \\ He~2-10 &8.49&0.872 &0.340 & 0.276 & (1),(11),(12) & Y (13) &0.098\\ \hline\\ \multicolumn{7}{c}{No detected extended emission}\\ \hline\\ NGC~178 & 9.40 &0.549 &0.019 & 0.315& (14),(4),(4) & N & 1.445 \\ NGC~7126 & 10.57$^d$ &5.21 & 0.0467& 0.188& (14),(4),(4) & N &1.349 \enddata \tablenotetext{a}{\mbox{$\Sigma_{\rm SFR}$}\ is calculated from the \Hline{\mbox{$\alpha$}}\ SFR using either the radius of the ionized emission \citep{b:Calzetti_apj10} or the effective \Hline{\mbox{$\alpha$}}\ radius \citep{b:Oey_apj07}.} \tablenotetext{b}{References, the first reference listed in the column is for the \Hline{\mbox{$\alpha$}}\ SFR, the second is for M$_{{\sc \rm HI}}$, and the third is for $R_{\Hline{\mbox{$\alpha$}}}$ -- (1) \citet{b:Calzetti_apj10}; (2) \citet{b:GildePaz_apjs03}; (3) \citet{b:Marlowe_apj95}; (4) \citet{b:Meurer_apjs06}; (5) \citet{b:Calzetti_aj99}; (6) \citet{b:Hota_mnras05}; (7) \citet{b:Kennicutt_apj09}; (8) \citet{b:Veilleux_apj02}; (9) \citet{b:Kennicutt_apjs08}; (10) \citet{b:Meurer_aj92}; (11) \citet{b:Sauvage_aap97}; (12) \citet{b:Johnson_aj00}; (13) \citet{b:Mendez_aap99}; (17) \citet{b:Oey_apj07}} \tablenotetext{c}{$M_*$ from Table \ref{t:props}.} \tablenotetext{d}{Log($M_{{\sc \rm HI}}$) is a combination of NGC 7126 and NGC 7125.} \end{deluxetable*} \subsection{Orientation bias} The interpretation of escaping Lyman continuum observations in starbursts and LBGs is challenging because of the need to disentangle observational effects from intrinsic trends. One issue under debate is whether the relatively few ($\sim10\%$) galaxies that have significant \mbox{$f_{\rm esc}$}\ are representative of the fraction of galaxies that have escaping Lyman continuum. Potential challenges that can modify the intrinsic distribution of \mbox{$f_{\rm esc}$}\ include low redshift interlopers, a preferred direction of escape, and small covering fractions for the escaping Lyman continuum \citep[e.g.,][]{b:Vanzella_apj12,b:Nestor_apj11}. Examining the morphologies of the ionization cones observed in this work can shed light on this issue. Two of the seven starburst galaxies studied here show strong evidence for optically thin ionization cones. In both NGC 5253 \citep[see also][]{b:Zastrow_apj11} and NGC 3125, the cones are aligned along the minor axis. This behavior is consistent with the expectations of massive-star feedback-driven models \citep[e.g.,][]{b:Murray_apj11,b:Kim_arXiv12}. The observed alignment implies a preferred direction; feedback will punch holes in the ISM more easily along the minor axis \citep[e.g.,][]{b:Veilleux_araa05}. In a study of low redshift LBG analogs, \citet{b:Heckman_apj11} measured extreme wind velocities ($v\sim1500\ \rm km/s$) in all galaxies with low optical depth to Lyman continuum. The high velocities indicate the direction of the wind must be nearly aligned with the line of sight to the galaxy. This relationship between low optical depth and high outflow velocities further strengthens the argument that Lyman continuum escape has a preferred direction \citep{b:Zastrow_apj11}. Furthermore, the ionization cones exhibit narrow morphologies and subtend an estimated 4\% of $4\pi$\ steradians. This implies that the covering fraction of escaping radiation will be much less than unity, and is consistent with the results from studies of high redshift star-forming galaxies \citep[e.g.,][]{b:Nestor_apj11,b:Kreimeyer_arXiv13}. We note that the narrow ionization cones do not necessarily imply small \mbox{$f_{\rm esc}$}. If the physical size of the cone's base is significantly larger than the size of the ionizing cluster, then large fractions of the ionizing flux produced by the cluster may still escape through the angularly narrow cone. The combination of small opening angles and preferred directions suggest that if one tried to directly detect escaping Lyman continuum, the axis of escape would need to be close to the observer's line of sight in order to obtain a strong detection. If more galaxies are like NGC 3125 and NGC 5253, then an orientation bias makes it more challenging to detect escaping Lyman continuum in starbursts and LBGs. \subsection{Possible Trends with Galaxy Properties} In light of the massive-star feedback-driven model discussed previously, the first galactic property to examine is the presence of a galactic wind. All of the galaxies that show evidence for extended emission also have evidence for either a galactic wind or superbubble expansion out of the galaxy. Thus confirming the link between winds and the fate of ionizing radiation \citep[e.g.,][]{b:Heckman_apj11}. However, NGC~1705 and NGC~1482, the two galaxies that have the most clearly established winds, both appear optically thick. We note that both galaxies without extended emission, NGC 178 and NGC 7126, do not have galactic winds. Thus, as discussed in \citet{b:Heckman_aj97}, we see that while wind activity is important for the escape of Lyman continuum, other factors clearly play a determining role. Another critical property to examine is star formation rate. Models predict that escaping ionizing radiation will be associated with extreme feedback from strongly star forming systems \citep[e.g.,][]{b:Clarke_mnras02,b:Wise_apj09,b:Paardekooper_aap11}. From Table \ref{t:SFR}, it is immediately apparent that SFR(\Hline{\mbox{$\alpha$}}) is not directly correlated with the optical depth of extended emission. The galaxy with the highest SFR, NGC 7126, is one that most clearly does not have extended optically thin emission. Furthermore, the galaxy with the second highest SFR, NGC 1482, is optically thick. A more meaningful comparison is to look at the \mbox{$\Sigma_{\rm SFR}$}\ and the specific SFR. These measures of the SFR take into account the sizes and masses of the galaxies. A more massive galaxy would need a higher SFR to produce enough feedback for ionizing radiation to escape. Table \ref{t:SFR} shows that NGC~178 and NGC~7126, which show no evidence for extended emission, do have low \mbox{$\Sigma_{\rm SFR}$}, which is consistent with our expectations. However, distinguishing between the galaxies that have optically thin and thick emission is more challenging. NGC~5253 has the highest \mbox{$\Sigma_{\rm SFR}$}, but NGG~3125 has \mbox{$\Sigma_{\rm SFR}$}\ lower than He~2-10, which is likely optically thick. One explanation is that the higher SFR in He~2-10 contains contamination from its AGN. However, the AGN in He~2-10 is not particularly luminous in \Hline{\mbox{$\alpha$}}\ \citep{b:Reines_nature11} making the significance of its contribution unclear. NGC~1705 has comparatively low \mbox{$\Sigma_{\rm SFR}$}. Paired with the lower \mbox{$\Sigma_{\rm SFR}$}\ of NGC~1482, the rates are consistent with the hypothesis that optical depth is correlated with \mbox{$\Sigma_{\rm SFR}$}. However, we cannot extrapolate any conclusions on trends with \mbox{$\Sigma_{\rm SFR}$}\, based only on our findings for these individual galaxies. The comparison between sSFR is similarly fraught. NGC~5253 and NGC~3125 do have among the highest sSFR, but their sSFRs are not significantly higher than those of the galaxies with optically thick emission. Furthermore, NGC~178, which doesn't have clear evidence for extended emission, has a rather high sSFR. Thus, our small sample supports previous work in its conclusion that strong SFR is necessary but not sufficient for significant \mbox{$f_{\rm esc}$}\ \citep{b:Heckman_apj01}. We caution the reader that our sample is small, and the range of star-formation rates is also small. Therefore, our sample does not provide the leverage to firmly establish any apparent correlation between optical depth and \mbox{$\Sigma_{\rm SFR}$}\ or sSFR. As an additional consequence of the small sample size, the galaxies that seem anomalous, such as NGC 178 with its high sSFR but lack of extended emission and He~2-10, which is optically thick despite its high \mbox{$\Sigma_{\rm SFR}$}, may not be significant outliers. In addition to SFR, the recent star formation history will have an influential role on the escape of ionizing radiation. As a stellar population ages, the rate of ionizing photons produced (\mbox{$Q_0$}) decreases and the hardness of the radiation field softens. The ages of the clusters responsible for the ionization cones in both NGC~5253 and NGC~3125 are between 2-5 Myr old \citep{b:Calzetti_aj97,b:Westera_aap04,b:Chandar_apj05,b:Raimann_mnras00}. In contrast with this, NGC~1705 is in a post-burst state (cluster ages $\sim 10-15$ Myr) in which most of the O stars in the main super star clusters have died off \citep{b:Heckman_aj97}. Thus, the softer and fainter B stars dominate the radiation field and do not provide enough ionizing radiation to allow for escape. This may similarly offer a partial explanation for the line ratios observed in He~2-10. While the main star forming region, embedded in the center of the galaxy, contains many young clusters with ages ranging from 2-20 Myr \citep{b:Johnson_aj00}, the second star-forming region, further towards the eastern edge of the galaxy is likely much older with fewer of its massive stars remaining. Furthermore, visual examination of the images of He~2-10 show closed bubbles, rather than fractured filaments (Figure \ref{f:img_he210}). We note that for radiation fields dominated by late type stars, the relationship between the ionization parameter map and the optical depth is complicated. In these situations, a transition to [S{\sc ii}]\ dominated gas does not necessarily indicate an optically thick nebula \citep{b:Pellegrini_apj12}. If the radiation field is soft but sufficiently strong, it is possible that not enough gas is present to absorb the lower energy photons, yet, the ionization parameter map still exhibits a transition zone to [S{\sc ii}]. Furthermore, whether the population is a single burst or an extended burst also influences the fate of ionizing radiation. While feedback generates low density passageways in the ISM, simulations of superbubble evolution show that there is a delay between superbubble formation and the actual escape of ionizing radiation \citep{b:Dove_apj00,b:Fujita_apj03}. During the phase prior to superbubble break-out, ionizing radiation gets trapped by the bubble walls \citep{b:Fujita_apj03}. Once the bubble breaks out of the disk, ionizing radiation may escape relatively unhindered. However, by this point in time, the rate of ionizing photons produced by the massive star population is considerably weaker than in the first few Myr \citep{b:Dove_apj00,b:Fujita_apj03}. If the starburst occurs over a short duration, essentially a single burst, then it is less likely to leak significant ionizing radiation. If instead the burst is more extended, the later episodes of star formation can capitalize on the passageways carved by prior star forming episodes. Observationally, \citet{b:McQuinn_apj10b} showed that extended bursts with typical timescales of hundreds of Myr are common in dwarf starbursts. Our data further support this scenario. In NGC~5253, the current burst of star formation occurred within the last 3--5 Myr \citep{b:Calzetti_aj97}. While it is unlikely that any superbubbles formed from this epoch will have broken from the disk, there is evidence for prior star formation episodes with ages on the order of 10--100 Myr \citep{b:Caldwell_apj89}. Thus, the ionization cone in NGC~5253 probably formed through low density regions cleared out by feedback from earlier episodes of star formation. In NGC~3125, the super star cluster at the base of the ionization cone has an age between 1--3 Myr, while the region surrounding it has an overall mean age of between 8--10 Myr \citep{b:Westera_aap04,b:Chandar_apj05}. While the initial episode of star formation in this burst is younger than that shown for NGC~5253, the dominant star formation is on the periphery of the galaxy, which makes it easier for ionizing radiation to escape \citep{b:Gnedin_apj08}. Furthermore, \citet{b:Raimann_mnras00} stacked a sample of similar HII galaxies in which NGC~3125 was the dominant source and found that 65\% of the stellar population is younger than 5 Myr, 32\% is between 5 and 100 Myr, and the remaining 3\% of the population us is older than 100 Myr. While the SFH analysis is not based solely on NGC~3125 it does point in the direction of NGC~3125 having multiple recent episodes of star formation. The young ages of the ionizing populations in NGC~3125 and NGC~5253 may represent a special age range needed to find escaping ionizing radiation. In a recent paper, \citet{b:Jaskot_apj13} study a sample of $z\sim0.1-0.3$ galaxies whose extremely high [O{\sc iii}]/[O{\sc ii}]\ suggest that they are likely optically thin. These galaxies all have ionizing populations that are between 3-5 Myr \citep{b:Jaskot_apj13}, similar to the ages observed in galaxies with ionization cones studied in this work. Furthermore, the two local galaxies that have direct \mbox{$f_{\rm esc}$}\ measurements, both have significant populations of young stars \citep{b:Leitet_aap13, b:Leitet_aap11}. Finally, it has been suggested that the mass of the galaxy is correlated with the escape fraction of ionizing radiation. There are some simulations that predict higher \mbox{$f_{\rm esc}$}\ with larger galaxy mass \citep[e.g.,][]{b:Gnedin_apj08}. Others claim the opposite because the smaller galaxy's weaker potential well would make escape easier \citep{b:Yajima_mnras11,b:Razoumov_apj10}. Based on our small sample here, we do not find evidence for or against either claim. The two galaxies for which the extended ionized gas is optically thin have masses that are in the middle to low range of masses in the sample (Tables \ref{t:props} and \ref{t:SFR}). Interesting to note, these two galaxies have significantly higher fractions of {\sc \rm HI}\ relative to the stellar mass of the galaxy (Table \ref{t:SFR}) than the sample galaxies with extended ionized gas. Meanwhile, NGC~5253 has an exceedingly low {\sc \rm HI}\ mass fraction, even when compared to the other galaxies with winds. This is consistent with the idea that the clearing, or consumption, of the neutral gas in the galaxy plays a role in regulating the escape of ionizing radiation. \section{Conclusions} \label{s:conc} The passage of ionizing radiation through, and possibly out of, a galaxy has ramifications on our understanding of cosmic reionization. In this paper, we study the extended, low surface brightness, emission-line gas of seven dwarf starburst galaxies. Using narrowband [S{\sc iii}], [S{\sc ii}], and \Hline{\mbox{$\alpha$}}\ images and the technique of ionization parameter mapping, we evaluate the optical depth of the extended emission. In two of the galaxies, we discover optically thin ionization cones extending along the minor axis. These cones suggest that ionizing radiation is escaping the main body of the galaxy. The narrow morphology suggests that ISM morphology is a critical determinant of whether ionizing photons will escape a galaxy. In three of the galaxies, we find that the established galactic winds are most likely optically thick. These galaxies, despite the presence of strong feedback do not show clear evidence for escaping ionizing radiation. The two remaining galaxies, despite the strong star formation show little evidence of extended emission. Our most convincing examples of optically thin emission are found in narrow ionization cones. These narrow cones illustrate a scenario in which the covering fraction of Lyman continuum is significantly less than unity, as has been suggested by other work \citep[e.g.,][]{b:Nestor_apj11}. Furthermore, both cones are aligned with the minor axis of the galaxy, indicating a preferred direction. The small opening angle and preferred direction suggest that unless the galaxy orientation is such that the axis of escape aligns with the line of sight, it will be challenging to directly detect escaping radiation. If other starbursts are similar to NGC~3125 and NGC~5253, an orientation bias at least partially explains the low detection rate of ionizing radiation in starburst galaxies and LBGs. In addition to observational biases, we explore the galactic properties that may regulate the escape of ionizing radiation. While galactic winds and bubble activity encourage escaping radiation, not all starbursts with galactic wind have escaping Lyman continuum. One explanation is that by the time the bubble or wind breaks out of the ISM, the ionizing population has aged and is no longer producing ionizing radiation as prodigiously \citep[e.g.,][]{b:Dove_apj00,b:Fujita_apj03}. Alternatively, the ionizing population is too young to have carved sufficient low-density paths out of the ISM for escape. This suggests an optimal age between 3-5 Myr for the escape of ionizing radiation. Based on the recent star-formation histories of NGC~3125 and NGC~5253, a starburst with a few recent star forming episodes strikes the balance between the time it takes for bubbles to burst and maximizing the number of ionizing photons produced that have clear passage out of the galaxy. Another potential contributing factor is the concentration of star formation in the burst. In the examples of optically thin ionization cones, we see that they emanate from specific, concentrated star forming regions, rather than extended star-forming areas. In contrast, NGC~7126 shows no sign for feedback-driven extended emission, and has its copious star formation spread throughout the disk of the galaxy. This suggests that concentrated star formation will be more effective at clearing out the ISM. However, we must be careful in drawing strong conclusions from so few objects. He~2-10 and NGG~1705, which also have concentrated star formation, are most likely optically thick to ionizing radiation. Thus, while the concentration of star formation is important, the interplay between the star formation and the ISM morphology most likely regulates the escape of ionizing radiation from starbursts and Lyman break galaxies. \acknowledgments Support for this work was provided by the National Science Foundation, grants AST-0806476, AST-0907758, and AST-1210285 to MSO; and AST-0606932 and AST-1009583 to SV and MM. We also acknowledge a Margaret and Herman Sokol Faculty Award to MSO. MM acknowledges support by NASA through a Hubble Fellowship grant HST-HF51308.01-A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. We are grateful to Colin Slater for help collecting observations and the UM FANG research group, Eric Pellegrini, Amy Reines, Lee Hartmann, and Timothy McKay for helpful discussions.
\section{Introduction} If cosmological dark matter (DM) is similar in character to a WIMP (weakly-interacting massive particle), it may possess a weak-scale cross-section for interaction with ordinary nucleons. In this case, collisions between DM particles in the halo of the Milky Way and nuclei in the Sun will slow some of the DM enough to gravitationally bind it to the Sun \cite{Press85,Griest87,Gould87b}. If enough is captured and remains in the Sun, DM may impact the structure and evolution of the Sun itself \cite{Steigman78,Spergel85,Faulkner85}. The effects of captured DM particles on the Sun and other stars have been a subject of investigation for many decades (see e.g.~\cite{Scott09,Turck12,Zurek13} for reviews). Much work has concentrated on the detectable signals of annihilating DM in stellar cores, such as $\sim$GeV-scale neutrinos escaping from the Sun \cite{Krauss86, Gaisser86, Griest87, Gandhi:1993ce, Bottino:1994xp, Bergstrom98b, Barger02, Desai04, Desai08, IceCube09, IceCube09_KK, IC40DM, SuperK11, IC22Methods, Silverwood12, IC79} and the effects of core heating by annihilation products \cite{SalatiSilk89, BouquetSalati89a, Moskalenko07, Spolyar08, Bertone07, Fairbairn08, Scott08a, Iocco08a, Iocco08b, Scott09, Casanellas09, Ripamonti10, Zackrisson10a, Zackrisson10b, Scott11}, or the MeV-scale neutrinos from these decay products \cite{Rott13, Bernal:2012qh}. The weakly-interacting nature of WIMP-like DM\footnote{Strictly, the WIMP paradigm implies certain weak-scale annihilation and nuclear scattering cross-sections, a GeV--TeV scale mass and a corresponding natural reproduction of the observed relic density via thermal production. Here all that is important is the interaction with nucleons, so throughout this paper we use `WIMP' loosely to refer to any DM with the requisite nuclear interaction, regardless of its mass or annihilation cross-section.} can also make it a medium for energy transport in the Sun or other stars \cite{Steigman78,Spergel85,Faulkner85}, the implications of which have been studied extensively \cite{Gilliland86, Renzini87, Spergel88, Faulkner88, Bouquet89, BouquetSalati89b, Salati90, Dearborn90b, Dearborn90, GH90, CDalsgaard92, Faulkner93, Iocco12}. Even when that energy transport is highly localised (\ie conductive), WIMPs may cool the core by absorbing heat, which can then be deposited at larger radii. These changes can be potentially constrained with the temperature-sensitive $^8$B neutrino flux, or complementarily, with precision helio/asteroseismological measurements of the sound speed, small frequency separations and gravity modes \cite{Lopes02a,Lopes02b,Bottino02,Frandsen10,Taoso10,Cumberbatch10,Lopes10,Lopes:2012,Lopes:2013,Casanellas:2013,Lopes:2014}. So far all studies of energy transport by DM particles have assumed that the cross-section for scattering between WIMPs and nuclei is independent of their relative velocity and the momentum exchanged in the collision. This is also a common assumption in analyses of direct searches for dark matter. Among these, DAMA \cite{Bernabei08}, CoGeNT \cite{CoGeNTAnnMod11}, CRESST-II \cite{CRESST11} and CDMS II \cite{Agnese:2013rvf} have reported excess events above their expected backgrounds, consistent with recoils caused by WIMP-like dark matter. On the other hand, XENON10 \cite{Angle:2011th}, XENON100 \cite{XENON2013}, COUPP \cite{COUPP12}, SIMPLE \cite{SIMPLE11}, Super-Kamiokande \cite{SuperK11} and IceCube \cite{IC79}, under the assumption of constant scattering cross-sections and a standard halo model, would together exclude such ``detections'' from being observations of WIMPs. However, there are assumptions built into these exclusions. Some recent work (e.g.\ \cite{Bozorgnia:2013hsa, Mao:2013nda}) has focused on the astrophysical uncertainties. A greater deal of interest has developed in scattering cross-sections with a non-trivial dependence upon the relative velocity or momentum transfer of the collision, as it appears that such velocity-- or momentum-dependent scattering could reconcile the various direct detection results \cite{Masso09, Feldstein:2009tr, Chang:2009yt, Feldstein10, Chang10, An10, Chang:2010en, Barger11, Fitzpatrick10, Frandsen:2013cna}.\footnote{This may be more difficult now that LUX \cite{LUX13} has released first results \cite{Gresham13}, however.} The focus on constant cross-sections in stellar dark matter studies is therefore not because these are the only interesting DM models, but simply because the required theoretical background for including the effects of velocity or momentum-dependent couplings to nucleons in stellar structure simulations does not exist in the literature. This is in contrast to direct detection, where such effects can be easily computed (e.g.\ \cite{Cirelli13}). Here we provide such calculations for a range of non-constant cross-sections. This work will allow the full transport theory to be incorporated into a solar evolution code that also treats the capture and annihilation of DM, in order to derive self-consistent limits on DM models with non-standard cross-sections. Our other motivation is the well-known Solar Abundance Problem: recent photospheric analyses \cite{APForbidO,CtoO,AspIV,AspVI,AGS05,ScottVII,Melendez08,Scott09Ni,AGSS} indicate that the solar metallicity is 10--20\% lower than previously thought, putting predictions of solar evolution models computed with the improved photospheric abundances at stark odds with sound-speed profiles inferred from helioseismology \cite{Bahcall:2004yr, Basu:2004zg, Bahcall06, Yang07, Basu08, Serenelli:2009yc}. A deal of effort has been devoted to finding solutions to this quandary \cite{Bahcall05, Badnell05, Arnett05, Charbonnel05, Guzik05, Castro07, Christensen09, Guzik10, Serenelli11, Frandsen10, Taoso10, Cumberbatch10, Vincent12}, but so far no proposal has really proven viable. Upcoming solar neutrino experiments such as SNO+, Hyper-Kamiokande and LENA should allow the solar core metallicity and temperature to be measured even more accurately than can be done with helioseismology alone \cite{Haxton13,Lopes13b,Smirnov13}, hopefully shedding some light on the discrepancy. One might postulate that the disagreement is in fact caused by enhanced energy transport in the solar core mediated by non-standard WIMP-nucleon couplings; the calculations we carry out here allow this proposition to be tested. This paper is structured as follows: first, in Section \ref{sec:backbackground} we provide some background on WIMP-nucleon couplings and how one goes about calculating energy transport by WIMPs, introducing the thermal conduction coefficients $\alpha$ and $\kappa$. In Section \ref{sec:background} we briefly summarise the relevant equations from Ref.\ \cite{GouldRaffelt90a} that allow calculation of $\alpha$ and $\kappa$. In Section \ref{sec:non-standard} we modify this treatment to account for non-trivial velocity and momentum structure in the scattering cross-section, and show the effect on $\alpha$ and $\kappa$. In Section \ref{sec:luminosity} we provide some examples of the effect on energy transport in the Sun, then finish with some discussion and concluding remarks in Section \ref{sec:conclusion}. \section{Preliminary theory} \label{sec:backbackground} The quantity that describes microscopic interactions between WIMPs and Standard Model nuclei is the total cross-section $\sigma$ for WIMP-nucleon scattering. Determining the impacts of energy conduction by a population of weakly-interacting particles in a star means working out their influence on bulk stellar properties like the radial temperature, density and pressure profiles. Initial solutions were obtained by way of simple analytic approximations \cite{Faulkner85,Spergel85,Gilliland86}, but this quickly gave way to the established approach of solving the Boltzmann collision equation (BCE) inside the spherically symmetric potential well of a star \cite{Nauenberg87,GouldRaffelt90a,GouldRaffelt90b}. The formalism and a general solution for standard WIMP-quark couplings were described by Gould and Raffelt \cite{GouldRaffelt90a}. There are three quantities that, together, are sufficient to describe the transport of energy by a diffuse, weakly-interacting massive particle $\chi$ in a dense medium such as a star:\begin{itemize} \item $l_\chi$, defined as the inverse of the mean number of WIMP-nucleon interactions per unit length; this gives the typical distance travelled by a WIMP between scatterings; \item $\alpha$, related to the thermal diffusion coefficient; this parameterises the efficiency of a species' diffusion inside the potential well; \item $\kappa$, the dimensionless thermal conduction coefficient; this describes the efficiency of energy transfer from layer to layer in the star. \end{itemize} Given a form of the WIMP-nucleon cross-section $\sigma$ as a function of the microscopic kinetic variables ($v_{\rm rel}, q$, etc.), the coefficients $\alpha$ and $\kappa$ are quantities that are averaged over the kinetic gas distributions, thus depending only on $\mu$, the ratio between the WIMP mass and the mass of the nucleus with which it scatters. They can therefore be computed once and for all for each type of WIMP-nucleon coupling. In the following sections, we compute the thermal conduction coefficients $\alpha$ and $\kappa$ of WIMPs with non-standard couplings to the standard model, going beyond the $\sigma = const.$ case computed explicitly by Ref.\ \cite{GouldRaffelt90a}. Because these computations must be done on a case by case basis, here we focus on the specific velocity- and momentum-dependent cross-sections $\sigma \varpropto v_{\rm rel}^{2n}$ and $\sigma \varpropto q^{2n}$, with $n = \{-1,1,2\}$.\footnote{We do not consider cases where $n \le -2$, as this causes the total momentum transfer to diverge; conduction properties would then intrinsically depend on the chosen cutoff scale.} Here $v_{\rm rel}$ is the WIMP-nucleus relative velocity, and $q$ is the momentum transferred during a collision. Our choice of these types of couplings is motivated in part by a number of recent direct-detection experiments \cite{Bernabei08,CoGeNTAnnMod11,CRESST11,Agnese:2013rvf}, but they can easily occur theoretically \cite{Fitzpatrick13,Kumar13}. Mixed couplings (whether linear combinations of these couplings or cross-terms between them) also each require explicit calculation, and cannot be treated using combinations of the results we present here; we leave these for future work. The dominant operators that arise in the most basic models of interactions between WIMPs $\chi$ and standard model particles (namely quarks, denoted $Q$) are $O_{SI} = \bar \chi \chi \bar Q Q$ and $ O_{SD} = \bar \chi \gamma_\mu \gamma_5 \chi \bar Q \gamma^\mu \gamma_5 Q$. These respectively lead to the regular spin-independent $\ssi$ and spin-dependent $\ssd$ cross-sections, which have no velocity or momentum structure beyond standard kinematic factors, and have been the focus of exclusion efforts in direct experimental searches to date. There is, however, no guarantee that the dominant coupling between the dark sector and the SM is as simple as $O_{SI}$ or $O_{SD}$. Possibilities include a finite particle radius (the dark sector analogue of the nuclear form factor), a vector coupling to quarks, parity-violating interactions such as $\bar \chi \gamma_5 \chi \bar Q Q$ or $\bar \chi \gamma_\mu \gamma_5 \chi \bar Q \gamma^\mu Q$ or even a small dipole or anapole interaction with the standard model \cite{Pospelov00,Sigurdson04,Chang:2009yt,Feldstein:2009tr,Feldstein10,Chang:2010en,Fitzpatrick10,Barger11,Frandsen:2013cna,DelNobile:2013cva}. If these terms dominate, then the interaction cross-section will be proportional to some power of the momentum transfer $q$: \begin{equation} \sigma = \sigma_0 \left(\frac{q}{q_0}\right)^{2n}, \label{pdep} \end{equation} where $\sigma_0$ is the cross-section at some normalisation momentum $q_0$. Models charged under multiple gauge fields can display interactions that go as a sum of different powers of $q^2$ \cite{Feldstein:2009tr}. Even for scattering via standard $O_{SI}$ or $O_{SD}$ couplings, scattering between WIMPs and nuclei can be strongly momentum-dependent due to the nuclear form factor. This is especially important at high momentum transfer, so has a large impact on rates at which WIMPs can be scattered to below the local escape velocity and captured by a star. In this paper we neglect the influence of nuclear form factors, mainly because the majority of WIMP scattering in stellar interiors is with hydrogen and helium, at much lower momentum transfer than successful capture events. In the cases where momentum is exchanged with heavier nuclei (mainly C, N and O), the suppression due to a Helm form factor is only a few percent at relevant values of $q$ -- much smaller, for example, than uncertainties on the the local interstellar WIMP density or velocity distribution. The form factors should in principle be taken into account if the impact of a given WIMP on a particular star is to be calculated in its fullest, goriest detail. Operators also exist which produce interactions that depend on the relative velocity $\vrel$ between the WIMP and the nucleus: \begin{equation} \sigma = \sigma_0 \left(\frac{\vrel}{v_0}\right)^{2n}. \label{vdep} \end{equation} The collisional cross-section $\sigma$ is proportional to the squared matrix element amplitude $|\mathcal M|^2$, which is typically a function of the centre-of-mass energy. This dependence can be expanded as Taylor series of $(\vrel/c)^2$, where the only sizeable contribution in the non-relativistic limit is the constant term. However, in some models the form of $|\mathcal M|^2$ causes a partial or exact cancellation of the constant term, leaving $\sigma \propto \vrel^2$. This is called \textit{p-wave} suppression. The case in which the second term is also suppressed -- leading to $\sigma \propto \vrel^4$, is referred to as \textit{d-wave}. On the other hand, DM models which scatter through the exchange of a massive force carrier exhibit a Sommerfeld-like enhancement. These can display ``resonant'' behaviour, which leads to a scattering cross-section proportional to $\vrel^{-2}$ \cite{Sommerfeld,Hisano05,AHDM}. Given a certain WIMP mass, a good choice for the reference velocity $v_0$ in a star is the temperature-dependent velocity \begin{equation} \label{vt} v_T(m_\chi,r) = [2 k_{\rm B} T(r)/\mx]^{1/2}. \end{equation} The typical thermal WIMP velocity can be obtained by multiplying $v_T$ by a further factor of $(3/2)^{1/2}$ \cite{GouldRaffelt90a}. Here $T(r)$ refers to the stellar temperature profile as function of radius $r$. In this paper we adopt $v_0=v_T(m_\chi=20\,\mathrm{GeV},r=0)=110 $\,km\,s$^{-1}$ in the Sun, and provide a simple means for rescaling our results to other values of $v_0$. Our chosen value is comparable to the typical relative velocities in direct detection experiments, where $v_0 \sim 250$\,km\,s$^{-1}$ is often employed. Similarly for $q_0$, where our default is $q_0=40$\,MeV: this is about the typical momentum transferred both in thermal collisions in the Sun and direct detection. This value corresponds to nuclear recoil energies of $E_{\rm R} = q^2/2m_{\rm nuc} \sim 10$\,keV in direct detection experiments. As for $v_0$, our results can be easily rescaled to any other $q_0$; the process is described in the passage following (\ref{sigmaqm2}). In this paper we will not focus on specific particle physics models or the WIMP-nucleon couplings they give rise to. Rather, we will take the general forms (\ref{pdep}) and (\ref{vdep}), and compute the effect on thermal conduction by an additional heavy species inside a star. \section{Thermal Conduction in a Star} \label{sec:background} \subsection{Theory} The theoretical framework of energy transport by WIMPs was constructed in detail by Gould and Raffelt \cite{GouldRaffelt90a}. Transport by a weakly-coupled species can occur in two distinct regimes. If the typical inter-scattering distance $l_\chi$ of a particle $\chi$ is much larger than the typical geometric length scale $r_\chi$ (which in this case is the length scale of the WIMP distribution in a star, not the stellar radius), one reaches the Knudsen limit of non-local transport. In the opposite regime, the Knudsen number $K \equiv l_\chi/r_\chi$ is $\ll 1$, and WIMPs scatter many times per scale height as they travel through the medium. Here we are interested in the latter regime, where local thermal equilibrium (LTE) holds. In \cite{GouldRaffelt90a}, Gould and Raffelt analysed the conduction of energy by a collection of massive particles $\chi$ in the LTE regime. They modelled energy transport using the Boltzmann collision equation: \begin{equation} DF = l_\chi^{-1} C F. \label{BCE} \end{equation} $F(\vect u, \vect r, t)$ here is the phase space density for particle velocity $\vect u$ and position $\vect r$ at time $t$. The typical inter-scattering distance for a WIMP is the inverse of the mean number of interactions it undergoes per unit length \begin{equation} \label{lchi} l_\chi(\vect u,\vect r) = \left[\sum_i n_i(r) \langle\sigma_i(\vect{u}) \rangle \right]^{-1}, \end{equation} were the index $i$ denotes the individual nuclear species present in the stellar gas, and the angular brackets represent thermal averaging. $n_i(\vect r)$ is simply the number density of species $i$ at location $\vect r$. The relative weighting of different WIMP velocities in the thermal average depends on the local WIMP velocity distribution, which is itself a function of the temperature with height in the star $T(\vect r)$; the thermal averaging is therefore itself understood to be dependent on $\vect r$. The differential operator is $D(\vect u, \vect r, t) = \del_t + \vect u \cdot \nabla_{\vect r} +\vect g(\vect r) \cdot \nabla_{\vect u} $, where $\vect g(\vect r)$ is the gravitational acceleration. To proceed, we make three key approximations: \begin{enumerate} \item The dilute gas approximation: WIMP-WIMP interactions can be neglected because they are much less frequent than WIMP-nucleon interactions.\footnote{This amounts to the restriction on the WIMP self-interaction term $\sigma_{\chi \chi} \ll \sigma_{\chi n} N_{nuc}/N_\chi$. This is not a strong restriction \textit{per se}, but one must be careful when considering certain DM models with strong self-interactions. For example, DM with large enough capture rates and allowed self interaction cross-sections $\sigma_{\chi \chi} \sim$ 1 cm$^2$/g can lead to large WIMP populations in the sun, violating this condition. This was not taken into account, for example, in the analysis of \cite{Taoso10}.} \item Local isotropy: this allows us to drop any $\theta, \phi$ dependence in the collision operator $C$. \item The conduction approximation: energy transport is conductive if $l_\chi$ is smaller than the other two length scales in the problem. That is, \begin{itemize} \item $l_\chi \ll r_\chi \implies K \ll 1$, \ie local thermal equilibrium. \item $l_\chi \ll |\nabla \ln T(\vect r)|^{-1}$, \ie the typical inter-scattering distance is much smaller than the length scale over which the temperature changes. This means that $\varepsilon \equiv l(r) |\nabla \ln T(r)| \ll 1$, allowing a perturbative expansion in powers of $\varepsilon$. \end{itemize} \end{enumerate} Actually, approximation 3 is only strictly necessary for the computation of $\alpha$ and $\kappa$ rather than their application, as correction factors exist for making the conduction treatment approximately applicable even in the non-local regime when $K > 1$ (see e.g.\ \cite{Bottino02,Scott09}). In practice both conditions in approximation 3. break down for cross-sections $\sigma \simeq 10^{-39}$\,cm$^2$ and smaller -- we will discuss the correction factors explicitly in Section \ref{sec:luminosity}. The collision operator $C$ has a straightforward intuitive definition: $CF$ is a function of the WIMP velocity $\vect u$, position $\vect r$ and time $t$, and represents the local rate of change of the WIMP phase space distribution. It can be formally written in terms of the components $\Cin$ and $\Cout$, \begin{eqnarray} CF &=& \int \ud^3 v\, C'(\vect u,\vect v, \vect r, t)F(\vect v, \vect r, t), \, \mathrm{with} \label{Cdef} \\ C' &\equiv& \Cin (\vect u,\vect v, \vect r, t) - \Cout (\vect v, \vect r, t)\delta^3(\vect v - \vect u). \end{eqnarray} $\Cin(\vect u,\vect v, \vect r, t)$ is the rate at which particles with velocity $\vect v$ are scattered to velocity $\vect u$. $\Cout(\vect v,\vect r,t)$ is the rate at which particles with velocity $\vect v$ are scattered to any other velocity. Performing the integral of $\Cin$ in (\ref{Cdef}) over incoming velocities $\vect v$ gives the total rate of scattering \textit{to} velocity $\vect u$ from any velocity, whereas the integral over $\Cout$ gives the total rate of scattering \textit{from} velocity $\vect u$ to any other velocity. The difference in these two rates is therefore the net rate of change in the WIMP phase space distribution. The BCE was solved by \cite{GouldRaffelt90a} perturbatively, with the expansion \begin{equation} F = F_0 + \varepsilon F_1 + ... \label{fperturb} \end{equation} where $\varepsilon\ll1$ is an expansion parameter (as explained above),\footnote{Note that in the case where the temperature gradient in a stellar core is so steep that $\varepsilon$ is not small, the entire treatment of energy transport by WIMPs presented here and in \cite{GouldRaffelt90a} breaks down because a perturbative expansion is not possible, even though $K$ may still be much less than 1. In such interesting cases the phase-space distribution would be strongly non-Boltzmannian, and an entirely different solution to the BCE would be required.} and $F_0$ is a regular Boltzmann distribution. Such a distribution is in thermal equilibrium by definition, so $CF_0 = 0$. Using this expansion to first order, dropping the term $\varepsilon D F_1$ because it is much smaller than $DF_0$, the first order BCE is \begin{equation} DF_0 = \frac{\varepsilon}{l_\chi}CF_1. \label{firstorderBCE} \end{equation} Assuming that the thermal timescale in a star is sufficiently long compared to the timescale for conductive energy transport by multiple WIMP scatterings, the solution of the BCE will be time-independent (stationary), such that the time-derivative in $D$ can be ignored. In this case, the left-hand side of (\ref{firstorderBCE}) is \begin{equation} DF_0(\vect u,r) = \frac{\varepsilon}{l_\chi} \left[ \vect \alpha(r) + \frac{\mx u^2}{2 T(r)}\frac{\nabla \ln T(r)}{|\nabla \ln T(r) | } \right] \vect u F_0(\vect u, r), \end{equation} where $\vect \alpha$ is a separation constant. Assuming the stellar temperature gradient to be spherically symmetric, both $DF_0$ and $F_1$ are pure dipoles. We use the standard notation for spherical harmonics: $Y_j^m(\theta,\varphi)$, where $j$ corresponds to the degree (related to total angular momentum) and $m$ is the order (angular momentum solely in the $\hat r$ direction). We can therefore express the BCE entirely in terms of monopole ($j=m=0$) and dipole ($j = 1, m = 0,\pm 1)$ components of the various quantities; this is the Rosseland approximation. It is convenient to transform our working variables to dimensionless quantities. We define the WIMP-to-nucleus mass ratio \begin{equation} \mu \equiv \mx/ m_{\rm nuc}, \label{mudef} \end{equation} and divide velocities by the local temperature-dependent velocity $v_T$ (\ref{vt}) \begin{eqnarray} \vect x &\equiv& \vect v/v_T, \label{xdef} \\ \vect y &\equiv& \vect u/v_T, \label{ydef} \\ \vect z &\equiv& \vect v_{\rm nuc} / v_T, \label{zdef} \end{eqnarray} with $v_{\rm nuc}$ denoting the velocity of a nucleus. We also define a dimensionless (angle-dependent, differential) cross-section $\shat$, and a total dimensionless (angle-independent, non-differential) cross-section $\shat_{\rm tot}$. The total dimensionless cross-section is simply $\shat$ integrated over the centre-of-mass scattering angle $\theta_{\rm CM}$: \begin{equation} \label{stotdef} \shat_{\rm tot}(v_{\rm rel}) \equiv \int_{-1}^1 \ud\cos\theta_{\rm CM}\, \shat(v_{\rm rel},q), \end{equation} remembering that $q$ is itself a function of both $v_\mathrm{rel}$ and $\cos\theta_{\rm CM}$. The original dimensionless differential cross-section ($\shat$) is defined by the requirement that the \textit{total} dimensionless cross-section be unity at $v_{\rm rel} = v_T$, \ie \begin{equation} \label{stotdef2} \shat_{\rm tot}(v_T) = 1. \end{equation} Similarly, we define the normalised phase space distribution functions \begin{equation} f_\nu^{j,m}(x,\vect r)\,\ud x \equiv \frac{1}{n_{\chi, \mathrm{LTE}}(r)} F_\nu^{j,m}(v,\vect r)\,\ud v, \label{fnormeq} \end{equation} where $n_{\chi, \mathrm{LTE}}(r)$ is the WIMP number density in the LTE (conductive) approximation, $\nu$ is the expansion order in $\varepsilon$, $j$ is the degree and $m$ is the order of the spherical harmonic expansion. The first order equation (\ref{firstorderBCE}) can then be expressed as the combination of \begin{equation} \left(\alpha_m y - \delta_{m0} y^3 \right) f_0^{0,0}(y,r) = \int \ud x\, C(y,x,r)f_1^{1,m}(x,r) \label{firstorderBCEv2} \end{equation} and the stationarity condition, expressed as the absence of a net WIMP flux across any given surface inside the star: \begin{equation} \int \ud x\, x f_1^{1,m} (x,r) = 0. \label{stationarity} \end{equation} Here $\alpha_m$ are thermal diffusivity coefficients corresponding to the dipole coefficients in the spherical harmonic expansion of $\hat{\vect u}\cdot\vect \alpha (r)$. In the Dirac notation of \cite{GouldRaffelt90a}, where \begin{eqnarray} \ket{f} &\equiv& f(y),\\ Q\ket{f} &\equiv& \int \ud x\, Q(y,x)f(x),\\ \braket{g}{f} &\equiv& \int \ud x\, g(x)f(x),\\ \bra{g}Q\ket{f} &\equiv& \int \ud y \ud x\, g(y)Q(y,x)f(x), \end{eqnarray} we see that (\ref{firstorderBCEv2}) and (\ref{stationarity}) become quite compact: \begin{eqnarray} \alpha_m (r) \ket{y f_0^{0,0}} -\delta_{m0} \ket{y^3 f_0^{0,0}} &=& C\ket{f_1^{1,m}}, \label{keteq}\\ \braket{y}{f_1^{1,m}} &=& 0. \end{eqnarray} In terms of the inverse of the collisional operator $C$, we can write the solution to the first-order BCE as \begin{equation} \ket{f_1^{1,m}} = \alpha_m C^{-1} \ket{y f_0^{0,0}} - \delta_{m0}C^{-1}\ket{y^3 f_0^{0,0}}. \label{keteq2} \end{equation} The diffusivity coefficients are then fully specified by multiplying (\ref{keteq2}) by $\bra{y}$: \begin{eqnarray} \alpha_{\pm 1} &=& 0, \\ \alpha_0 &=& \frac{\bra{y} C^{-1} \ket{y^3 f_0^{0,0}}}{\bra{y} C^{-1} \ket{y f_0^{0,0}}}. \label{eqn:alpha} \end{eqnarray} This can be used to obtain the stationary WIMP density profile \cite{GouldRaffelt90a} \begin{eqnarray} \label{LTEdens} n_{\chi,{\rm LTE}}(r) &=& n_{\chi,\mathrm{LTE}}(0)\left[\frac{T(r)}{T(0)}\right]^{3/2} \\ &\times&\exp\left[-\int^r_0 \ud r'\,\frac{k_{\rm B}\alpha(r')\frac{\ud T(r')}{\ud r'} + m_\chi\frac{\ud \phi(r')}{\ud r'}}{k_{\rm B}T(r')}\right],\nonumber \end{eqnarray} where $\phi(r)$ refers to the gravitational potential at height $r$ in the star. Together with the thermal conductivity $\kappa$ \begin{equation} \kappa = \frac{\sqrt{2}}{3} \braket{y^3}{f_1^{1,0}}, \end{equation} one can then use (\ref{LTEdens}) to finally obtain the luminosity carried by WIMP scattering, \begin{eqnarray} \label{LTEtransport} L_{\chi,{\rm LTE}}(r) &=& 4\pi r^2 \zeta^{2n}(r) \kappa(r)n_{\chi,{\rm LTE}}(r)l_\chi(r) \nonumber \\ &\times& \left[\frac{k_\mathrm{B}T(r)}{m_\chi}\right]^{1/2}k_\mathrm{B}\frac{\ud T(r)}{\ud r}, \end{eqnarray} and the corresponding energy injection rate per unit mass of stellar material: \begin{equation} \label{epsLTE} \epsilon_{\chi,{\rm LTE}}(r) = \frac{1}{4\pi r^2 \rho(r)}\frac{\ud L_{\chi,{\rm LTE}}(r)}{\ud r}. \end{equation} Here $\rho(r)$ is the stellar density, $\zeta(r)= v_0/v_T(r)$ for velocity-dependent scattering and $\zeta(r)= q_0/[m_\chi v_T(r)]$ for momentum-dependent scattering. The factors of $\zeta$ in (\ref{LTEtransport}) connect the velocity (momentum) scale at which the reference cross-section $\sigma_0$ is defined, to the thermal scale $v_T$ ($m_\chi v_T$) at which the dimensionless conductivity $\kappa$ is computed. Note that as per Ref.\ \cite{Scott09}, our sign convention is such that positive $L$ and $\epsilon$ refer to energy injection rather than evacuation, which differs from Ref.\ \cite{GouldRaffelt90a}. The problem of calculating the rate of energy transport by WIMPs is now one of explicitly calculating $\alpha$ and $\kappa$ for the WIMP-nucleus mass ratio and cross-section of interest. The only position dependence in the dimensionless BCE arises through the temperature; the transformation to dimensionless variables $x, y$, etc., means that the collision operator $C$, as well as $\alpha$ and $\kappa$, contain no $r$-dependence at all. This allows $\alpha$ and $\kappa$ to be expressed solely in terms of the mass ratio $\mu$. In a real star, more than one species $i$ of gas is present however, in different ratios at different $r$. Thus, to compute (\ref{LTEdens}) and (\ref{LTEtransport}) $\alpha$ and $\kappa$ must be computed for the specific mixture of gases at each height \cite{Scott09}: \begin{equation} \alpha(r,t) = \sum_i \frac{\sigma_i n_i(r,t)}{\sum_j \sigma_j n_j(r,t)} \alpha_i(\mu_i) \end{equation} and \begin{equation} \kappa(r,t) = \left[l_\chi(r,t)\sum_i\left\{\kappa_i(\mu_i)l_i(r,t)\right\}^{-1} \right]^{-1}. \label{kappaTotal} \end{equation} Here $n_i(r,t)$ is the number density of species $i$, while $\sigma_i$ and $l_i(r,t) \equiv (n_i(r)\langle\sigma_i(\vect u) \rangle )^{-1}$ are respectively the scattering cross-section and the typical interscattering distance of WIMPs with species $i$; this is in contrast with the complete sum in (\ref{lchi}). Henceforth we denote $\alpha \equiv \alpha(\mu)$ and $\kappa \equiv \kappa(\mu)$. Gould and Raffelt \cite{GouldRaffelt90a} calculated these quantities for a range of mass ratios, but only the case of velocity- and momentum-independent cross-sections; here we extend the treatment to more general cross-sections. \subsection{Calculations of $\alpha$ and $\kappa$} Because we plan on evaluating the collision operator $C$ numerically, it makes sense to express the functions $\ket{f(y)}$ as one-dimensional ``vectors'' and the operators as two-dimensional matrices. This allows direct calculation of a well-defined inverse operator $C^{-1}$, simply by inverting the corresponding matrix for $C$. Although the integrals of the form $Q \ket{f(y)}$ are from zero to infinity, the exponential behaviour of $f(y)$ means that the $y > 3$ contribution to each integral is less than a fraction of a percent. We therefore truncate the vectors and matrices at $y = 5$. $\Cout$ is relatively straightforward to construct: \begin{equation} \Cout(x, r) = \int \ud^3 z\, |\vect x - \vect z| \hat \sigma_{\rm tot}(v_T | \vect x - \vect z |) F_{\rm {nuc}}(\vect z), \label{CoutDefinition} \end{equation} where $F_{\rm {nuc}}(\vect z)$ is the velocity distribution of the nuclei, \begin{equation} F_{\rm {nuc}}(\vect z) = (\pi \mu)^{-3/2} e^{-|\vect z|^2/\mu}. \label{fnucdef} \end{equation} An expression for $\Cin$ is not as obvious. We transform to centre of mass (CM) coordinates, where $\vect a$ is the velocity of the CM frame, and $\vect b$ and $\vect b'$ are respectively the incoming and outgoing WIMP velocities in that frame.\footnote{In the language of Ref.~\cite{GouldRaffelt90a} $\vect a$ and $\vect b$ correspond to $\vect s$ and $\vect t$, respectively.} These quantities are illustrated in Figure \ref{kinfig}. \begin{figure}[tbp] \includegraphics[width=.5\textwidth]{figures/Cinkinematics} \caption{\textit{Scattering kinematics in the centre of mass frame, for the computation of $\Cin$. The velocity of the CM with respect to the lab frame is $a$. The incoming and outgoing WIMP velocities are $b$ and $b'$ respectively. Nucleon velocities are not shown, as they are simply related to the WIMP quantities in this frame.}} \label{kinfig} \end{figure} \begin{table*}[t!] \caption{Analytic forms of the integral (\ref{CoutDefinition}), given as a function of $w \equiv y/\sqrt{\mu}$. Expressions are identical for velocity-dependent and momentum-dependent cross-sections of the same order $2n$.} \label{tab:coutv} \vspace{2mm} \begin{tabular}{l@{\hspace{1cm}}l} \hline\hline\vspace{1mm} $\sigma=const.$& $\Cout(y) = \mu^{1/2} \left[ \left(w + \frac{1}{2w}\right)\mathrm{erf} (w) + \frac{1}{\sqrt{\pi}}\exp (-w^2)\right]$\\\vspace{1mm} $\sigma\varpropto v_\mathrm{rel}^2, q^2$ & $\Cout(y) = \mu^{3/2} \left[\frac{3 + 12w^2 + 4 w^4}{4w} \mathrm{erf}(w) +\frac{5 + 2 w^2}{2 \sqrt{\pi}}\exp(-w^2) \right]$ \\\vspace{1mm} $\sigma\varpropto v_\mathrm{rel}^4, q^4$& $\Cout(y) = \mu^{5/2} \left[\frac{ \left(15+8 w^6+60 w^4+90 w^2\right)}{8w}\mathrm{erf}(w)+\frac{ \left(33 +4 w^4+28 w^2\right)}{4\sqrt{\pi}}\exp(-w^2)\right]$\\\vspace{1mm} $\sigma\varpropto v_\mathrm{rel}^{-2}, q^{-2}$ & $\Cout(y) = \mu^{-1/2} w^{-1} \mathrm{erf}(w) $\\ \hline\hline \end{tabular} \end{table*} Ref.\ \cite{GouldRaffelt90a} then finds \begin{eqnarray} \label{cinint} \Cin^j(y,x,r) &=& (1+\mu)^4 \frac{y}{x}\int_0^\infty \ud a\, \int_0^\infty \ud b\, F_{\rm nuc}(\vect z) 2 \pi b \langle P_j \hat \sigma \rangle \nonumber \\ &\times& \Theta(y-|a-b|)\Theta(a+b - y)\nonumber\\ &\times& \Theta(x-|a-b|)\Theta(a+b-x). \end{eqnarray} $\langle P_j (\cos \theta_{\rm lab}) \hat \sigma \rangle$ is the angle-averaged product of the differential cross-section with the $j$-th Legendre polynomial, as expanded around transverse scattering angles in the lab frame. The angle averaging is performed in the azimuthal direction around the $\vect a$ axis in the CM frame. The integral over the zenith angle is performed implicitly when integrating over $a$ and $b$. Note that in this angle-average, $P_j$ is a function of $\theta_{\rm lab}$, whereas $ \shat $ is a function of $\vrel$ and $\theta_{\rm CM}$. The two angles are related as \begin{eqnarray} \cos \theta_{\rm CM} &=& A + B\cos \phi_{a b'}, \\ \cos \theta_{\rm lab} &=& G + B\frac{b^2}{xy} \cos \phi_{a b'}, \\ \end{eqnarray} with \begin{eqnarray} A &\equiv& \cos \theta_{a b'} \cos \theta_{a b} \label{Adef} \\ &=& \frac{(x^2-a^2-b^2)(y^2-a^2-b^2)}{4a^2 b^2}, \nonumber\\ B &\equiv& \sin \theta_{a b'} \sin \theta_{a b}, \label{Bdef} \\ G &\equiv& \frac{(x^2+a^2-b^2)(y^2+a^2-b^2)}{4a^2 xy}.\label{Gdef} \end{eqnarray} As we are only solving the BCE to first order, the only relevant term in (\ref{cinint}) is $\langle P_1 \hat \sigma \rangle$. Noting that \mbox{$\vrel=(1+\mu)bv_T$,} we then have \begin{eqnarray} \langle P_j \shat \rangle = \langle P_1 \shat \rangle &=& \frac{1}{2\pi}\int_0^{2\pi} d \phi_{a b'} \left(G + B\frac{b^2}{xy} \cos \phi_{a b'}\right) \nonumber\\ &\times& \shat \left[(1+\mu)bv_T, A + B\cos \phi_{a b'}\right]. \end{eqnarray} \section{conduction coefficients for non-standard WIMPs} \label{sec:non-standard} We write the cross-sections of interest to us explicitly in terms of the dimensionless variables. We will find $\Cout$ will usually have a tractable analytic form, but that the evaluation of the kinematic integrals for $\Cin$ (\ref{cinint}) must be done numerically, which we do using the CUBPACK multi-dimensional adaptive cubature integration package \cite{CUBPACK}. In every case, we produce an $N \times N$ linearly-spaced matrix for $C(y,x)$, which we explicitly invert to find $\alpha$, $\ket{f_1^{1,0}(y)}$ and $\kappa$. We find $N=500$ to be sufficient: both $\alpha$ and $\kappa$ vary by less than one part in $10^{5}$ when going from $N = 400$ to $N = 500$, across the full range of $\mu$ values we consider. For the standard $\sigma = const.$ case, our method of explicit matrix inversion yields identical results to the values of $\alpha(\mu)$ and $\kappa(\mu)$ presented in Table I of Ref.\ \cite{GouldRaffelt90a}, where a slightly different iterative method was used. These values are shown in the figures in this section and are given explicitly in Table \ref{aktable} at the end of the text. Our values of $\alpha$ agree exactly within given precision with Gould and Raffet for small $\mu$, and to within a part in $\sim 10^4$ as $\mu \rightarrow 100$. Our $\kappa$ results agree within $10^{-3}$ at small $\mu$, up to $\sim 1\%$ as $\mu$ approaches 100. These differences are small enough not to have any appreciable impact on modelled energy transport. We have furthermore cross-checked our results for the collision operator $\Cin$ and $\Cout$ as well as the first-order solution $f_1^{1,0}(y)$ presented in the figures of Ref.\ \cite{GouldRaffelt90a}. \subsection{Velocity-dependent scattering} \label{subsec:vdepscat} The relative velocity is the difference between the incoming WIMP speed $x$ and the nucleus speed $z$; expressed in terms of these velocities, the dimensionless differential cross-section is just \begin{equation} \hat \sigma_{v^{2n}} = \frac12|\vect x - \vect z | ^{2n}. \end{equation} $\Cout$ is then modified by extra powers of the relative velocity. The angular and radial integrals can be performed analytically, and the results are given in Table \ref{tab:coutv}. $\Cin$ is also modified in a simple way. The angular dependence drops out in the centre of mass frame where we compute $\Cin$ and $\vect x - \vect z = (1 + \mu) \vect b$, so that \begin{equation} \hat \sigma_{v^{2n}} = \frac12(1+\mu)^{2n} b^{2n}. \label{cinv2} \end{equation} These are related to (\ref{vdep}) via: \begin{equation} \sigma = 2 \sigma_0 \zeta^{-2n} \shat_{v^{2n}}, \end{equation} where $\zeta$ is defined below (\ref{epsLTE}). There is one final complication in the computation of $\Cin$ (\ref{cinint}) for the $\vrel^{-2}$ case: the expression (\ref{cinv2}) diverges for $b \rightarrow 0$, complicating the numerical integration of $\Cin$, even though the integral itself is finite. We address this by imposing a cutoff velocity $\om$ such that: \begin{equation} \shat_{v^{-2}} \rightarrow \frac{1 + \om^2}{2\left(|\vect x - \vect z |^2 + \om^2\right) } = \frac{1 + \om^2}{2(1+\mu)^2 b^2 + 2\om^2}, \label{v2cutoff} \end{equation} where the factor in the numerator ensures that $\hat \sigma_{\rm tot} (v_T) = 1$. Both $\alpha$ and $\kappa$ converge to finite values as $\om^2 \rightarrow 0$. This convergence is illustrated in Figure \ref{fig:vm2cutoff}. As $\omega^2$ is varied from $10^{-3}$ to $10^{-2}$, $\alpha(\mu)$ varies by less than one part in $10^4$, whereas $\kappa(\mu)$ changes by less than $1\%$ at low $\mu$, up to $\sim 4\%$ for $\mu = 100$. The thermally-averaged cross-section, necessary for computing a particle's typical inter-scattering distance $l_\chi$, can potentially depend on non-zero $\om$. A large value is reasonable, for example, in the case of resonant Sommerfeld-enhanced scattering, where the cross-section can reach a saturation value for $\vrel/c \approx 10^{-5}-10^{-3}$, below which $\sigma$ becomes constant. Values of $\Cin(y,x)$ for $\mu = 1$ are illustrated on the left-hand side of Figure \ref{Cinfig}. For positive powers of $\vrel$, this results in enhanced scattering of particles with high incoming velocity $v$; conversely this enhancement is present a low incoming velocities for $\sigma \propto \vrel^{-2}$. \begin{figure*}[p] \begin{tabular}{c@{\hspace{0.04\textwidth}}c} \multicolumn{2}{c}{\includegraphics[height = 0.32\textwidth]{figures/f1}} \\ \includegraphics[height=0.32\textwidth]{figures/f2} & \includegraphics[height=0.32\textwidth]{figures/f6} \\ \includegraphics[height=0.32\textwidth]{figures/f4} & \includegraphics[height=0.32\textwidth]{figures/f7} \\ \includegraphics[height = 0.32\textwidth]{figures/f3} & \includegraphics[height = 0.32\textwidth]{figures/f8} \\ \end{tabular} \caption{\it Partial dimensionless collision operator $\Cin(y,x)$ with $\mu = \mx/m_{\rm nuc} = 1$ for the various types of cross-section, as a function of the WIMP's dimensionless incoming velocity $x$ and outgoing velocity $y$. The angular integral in the $q^{-2}$ case (bottom-right panel) is regulated with $\xi = 10^{-8}$ (\ref{P1qm2}) to remove the divergence at $x = y$. Note also the log scale in this panel. } \label{Cinfig} \end{figure*} \begin{figure*}[tbp] \begin{tabular}{c@{\hspace{0.04\textwidth}}c} \includegraphics[width=0.48\textwidth]{figures/alphav2p2} & \includegraphics[width=0.48\textwidth]{figures/kappav2p2} \\ \includegraphics[width=0.48\textwidth]{figures/alphavm2pm2} & \includegraphics[width=0.48\textwidth]{figures/kappavm2pm2} \\ \end{tabular} \caption{\textit{ Dimensionless thermal diffusivity $\alpha$ (left) and conductivity $\kappa$ (right) as a function of $\mu = \mx/\mn$. Curves are shown for velocity-dependent scattering cross-sections $\sigma \propto \vrel^{2n}$ and for momentum-dependent cross-sections $\sigma \propto q^{2n}$. The upper panels show curves for positive values of $n$, and bottom panels show results for $n = -1$. In every case the thick red line is the fiducial case $\sigma = const.$, which was calculated explicitly by Gould and Raffelt \cite{GouldRaffelt90a}, and is identical to the result presented in their Figure 5 (panel b); this is in contrast to the results obtained by \cite{Nauenberg87}, who did not include the $\Cin$ contribution . Note that, because we regulate the cross-section for $q^{-2}$-dependent scattering using the momentum transfer cross-section (\ref{sTdef}), the $\kappa$ curves for $\vrel^{-2}$ and $q^{-2}$ are identical.}} \label{v2p2fig} \end{figure*} \begin{figure*}[tbp] \begin{tabular}{c@{\hspace{0.04\textwidth}}c} \includegraphics[width=0.48\textwidth]{figures/vm2cutoff_alpha} & \includegraphics[width=0.48\textwidth]{figures/vm2cutoff_kappa} \\ \end{tabular} \caption{\textit{The effect of a cutoff $\omega$ to the enhancement on $\alpha$ and $\kappa$, as defined in (\ref{v2cutoff}). Increasing the cutoff from $\omega = 0$ to $\omega \gg 1$ smoothly transforms $\alpha$ and $\kappa$ from the $\sigma \propto \vrel^{-2}$ case to the $\sigma = const.$ case. This type of behaviour could occur, for example, in models of Sommerfeld-enhanced DM scattering, which corresponds to an attractive force with a massive force carrier. }} \label{fig:vm2cutoff} \end{figure*} In Figure \ref{v2p2fig}, the thermal conduction coefficients $\alpha$ and $\kappa$ are shown for $\vrel^{2}$ scattering (blue dashed lines), and for $\vrel^4$ (cyan dotted lines). For the $\vrel^{-2}$ case, $\kappa$ and $\alpha$ converge to well-defined curve as $\om^2 \rightarrow 0$; these are the values we show in the bottom panels of Figure \ref{v2p2fig} (dark red dashed lines), and later use to compute $L(r)$. Numerical values are provided in Table \ref{aktable}. The effect of positive powers of $\vrel$ is to suppress $\alpha$ at low values of $\mu$, and to suppress $\kappa$ for large values of $\mu$. $\alpha$ is a measure of how well WIMPs can diffuse outward in the potential well of a star. This suppression ultimately means that the distribution $n_\chi(r)$ will be more compact than in the standard case. The effect of a $\kappa$ suppression, meanwhile, implies that the WIMP population will be less efficient at heat transport via scattering with light nuclei such as hydrogen. In the lower panel of Figure \ref{v2p2fig} we show that the opposite behaviour is true for $\sigma \propto \vrel^{-2}$: a ``fluffier'' WIMP distribution around the stellar core, with enhanced heat transport via collisions. In the case of a large cutoff velocity $\omega$, the thermal conduction coefficients lie in between the $\sigma = const.$ and $\sigma \propto \vrel^{-2}$ cases. We illustrate this in Figure \ref{fig:vm2cutoff}, for several values of the dimensionless cutoff velocity $\omega$. Sommerfeld-enhanced DM models, therefore, would benefit from some of the enhanced conduction of the $\vrel^{-2}$ case. \subsection{Momentum-dependent scattering} The momentum transferred during a collision, $q = \sqrt{2\mn E_{\rm R}}$ is equal to the total momentum gained or lost by the WIMP: \begin{eqnarray} q^2 &=& \mx^2 |\vect u - \vect v|^2 \nonumber \\ &=& q_0^2 \zeta^{-2} |\vect x - \vect y|^2 \nonumber \\ &=& 2 b^2 \zeta^{-2} q_0^2 (1-\cos \theta_{\rm CM}). \label{pdepeq} \end{eqnarray} The latter expression allows us to write the cross-section in terms of the centre-of-mass velocity of the incoming WIMP, $b$, and the CM scattering angle $\theta_{\rm CM}$, as the scattering event does not change the WIMP's speed in the CM frame, only its direction. Then \begin{equation} \shat_{q^{2n}} = 2^{-n}H_n|\vect x - \vect y|^{2n} = H_nb^{2n} (1-\cos \theta_{\rm CM})^{n}, \label{sigmaofq} \end{equation} where the normalisation factor required to ensure that $\hat\sigma_{\rm tot}(v_T) = 1$ is \begin{equation} \label{H_n} H_n = \frac{(1+\mu)^{2n}}{\int_{-1}^1 (1-\cos \theta')^{n}\, \ud \cos \theta' }. \end{equation} The structure of (\ref{sigmaofq}) shows us that \begin{equation} \shat_{q^{2n}} = 2\shat_{v^{2n}}\frac{(1-\cos \theta_{\rm CM})^{n}}{\int_{-1}^1 (1-\cos \theta')^{n}\, \ud \cos \theta' }, \end{equation} so \begin{equation} \shat_{\mathrm{tot},q^{2n}} = \shat_{\mathrm{tot},v^{2n}} = (1+\mu)^{2n} b^{2n}, \end{equation} as we would expect, given that $\shat_\mathrm{tot}$ is independent of $\theta_{\rm CM}$. This also means that $C_{\mathrm{out},q^{2n}} = C_{\mathrm{out},v^{2n}}$. In terms of the dimensionful cross-section (\ref{pdep}), \begin{equation} \sigma = \frac{2^n}{H_n} \sigma_0 \zeta^{-2n} \shat_{q^{2n}}. \end{equation} The component $\Cin$ requires an explicit evaluation of $\langle P_1 \shat (\cos \theta_{\rm CM}) \rangle$: \begin{eqnarray} \langle P_1 \shat_0 \rangle &=& \frac12 G, \label{P1q0} \\ \langle P_1 \shat_{q^{2}} \rangle &=& \frac12b^2(1+\mu)^2 \left[G(1-A) - \frac{b^2B^2}{2xy}\right], \label{P1q2} \\ \langle P_1 \shat_{q^{4}} \rangle &=& \frac38b^4(1+\mu)^4 \Bigg[ \left\{G(1-A) - \frac{b^2B^2}{xy}\right\}(1-A) \nonumber\\ &+& \frac{GB^2}{2} \Bigg], \label{P1q4} \\ \langle P_1 \shat_{q^{-2}} \rangle &=& \frac{1}{\ln 2 - \ln \xi}b^{-2}(1+\mu)^{-2} \nonumber \\ &\times&\left[\frac{(1-A+\xi)b^2 + Gxy}{xy\sqrt{(1-A+\xi)^2-B^2}} - \frac{b^2}{xy}\right], \label{P1qm2} \end{eqnarray} where $A$, $B$ and $G$ are defined in (\ref{Adef}--\ref{Gdef}). The parameter $\xi$ comes from the replacement $\cos \theta_{\rm CM} \rightarrow \cos \theta_{\rm CM} - \xi$. We make this substitution because for $n = -1$, the factor $(1 - \cos \theta)^{-1}$ in (\ref{sigmaofq}) causes the cross-section to diverge for forward scattering at small angles; likewise for the integral (\ref{stotdef}). For the computation of $\alpha$ in (\ref{eqn:alpha}), the powers of $C^{-1}$ allow the divergence to cancel. However, $\kappa$, which governs momentum transfer, formally diverges. Fortunately, the divergence at $\theta_{\rm CM}=0$ corresponds to forward scattering, in which no momentum is actually transferred. We regulate this divergence with the ``momentum transfer cross-section,'' which comes from plasma physics \cite{Krstic} but has also been applied to parameterise transport by WIMPs \cite{Tulin:2013teo}: \begin{align} \label{sTdef} \shat_{\rm tot,T}(v_{\rm rel}) &\equiv \int_{-1}^1 \ud\cos\theta_{\rm CM}\, \shat_{\rm T}(v_{\rm rel},q), \\ \shat_{\rm T}(v_{\rm rel},q) &\equiv (1-\cos\theta_{\rm CM})\shat(v_{\rm rel},q), \end{align} where $\shat$ is as per (\ref{sigmaofq}), but with the replacement $n\to n+1$ in the denominator of (\ref{H_n}). This gives \begin{equation} \label{P1qm2reg} \langle P_1 \shat_{\mathrm{T},q^{-2}} \rangle = \frac12 \frac{G}{b^2 (1+ \mu)^2}, \end{equation} leading to a finite, well-defined value of $\kappa$. The behaviour of $\Cin$ as a function of $x$ and $y$ for momentum-dependent cross-sections is shown in the right-hand panels of figure $\ref{Cinfig}$. The two upper right-hand panels show an enhancement in the scattering rate when the momentum transfer is large, over both the constant and velocity-dependent cases, which favour collisions closer to the $x = y$ line. The lower-right panel shows the $q^{-2}$ case computed with (\ref{P1qm2}) and $\xi = 10^{-8}$, illustrating the divergent behaviour in the forward-scattering rate, and a general suppression of collisions which lead to appreciable momentum exchange. The values of $\alpha$ and $\kappa$ are shown in dot-dashed orange for the $q^{2}$ case and in dot-dashed green for $q^{-2}$ in Figure \ref{v2p2fig}. The $q^4$ case is also shown (dashed purple lines). The behaviour of $\alpha$ and $\kappa$ for a $q^{2n}$ cross-section is qualitatively similar to the $\vrel^{2n}$ case. Indeed, $\kappa$ is identical for the $q^{-2}$ and $\vrel^{-2}$ cases, thanks to the momentum-transfer cross-section, which essentially deletes the angular dependence in the $q^{-2}$ case and makes $C$ identical for momentum and velocity-dependent cross-sections when $n=-1$. In the next section we will show that, once this behaviour is combined with the different typical inter-scattering distances in models with momentum-dependent and velocity-dependent scattering, the conduction of energy can be greatly modified with respect to the fiducial constant cross-section case. \section{Effect on energy transport in the Sun} \label{sec:luminosity} We now illustrate the effect of a non-standard WIMP-nucleon cross-section on the conduction of energy within the Sun. Our discussion applies equally well to any star, but modelling and measurements of the Sun and its properties are far more precise than for other stars. We use the temperature, density and elemental composition profiles computed with the AGSS09ph solar model\footnote{publicly available at \url{http://www.mpa-garching.mpg.de/~aldos/}} \cite{AGSS,Serenelli:2009yc}. Scattering cross-sections $\sigma_0 \lesssim 10^{-38}$\,cm$^2$ allowed by bounds from direct detection experiments can easily lead to energy transport outside the LTE regime, \ie non-local transport. To account for this, based on the results of Gould \& Raffelt \cite{GouldRaffelt90a,GouldRaffelt90b}, Refs.\ \cite{Bottino02,Scott09} introduce the quantities: \begin{eqnarray} \mathfrak{h}(r) &=& \left(\frac{r - r_\chi}{r_\chi}\right)^3 + 1, \label{hr}\\ \mathfrak f(K) &=& \frac{1}{1+ \left(\frac{K}{K_0}\right)^{1/\tau}}, \label{fK} \end{eqnarray} where $K$ is the Knudsen number, $K_0 \simeq 0.4$ and $\tau \simeq 0.5$. The WIMP scale height $r_\chi$ as a function of the central temperature $T_c$ and density $\rho_c$ is \begin{equation} r_\chi = \left(\frac{3 k_B T_c}{2\pi G \rho_c \mx}\right)^{1/2}. \label{scaleheight} \end{equation} The WIMP distribution becomes a combination of the isothermal and LTE distributions: \begin{equation} n_\chi(r) = \mathfrak f(K) n_{\chi,{\rm LTE}} + \left[1 - \mathfrak f(K) \right] n_{\chi,{\rm iso}}, \end{equation} where \begin{equation} \label{isodens} n_{\chi,{\rm iso}}(r,t) = N(t)\frac{e^{-\frac{r^2}{r_\chi^2}}}{\pi^{3/2}r_\chi^3}. \end{equation} The total luminosity can be finally written: \begin{equation} L_{\chi,\rm total}(r,t) = \mathfrak f(K) \mathfrak h(r,t)L_{\chi,{\rm LTE}}(r,t). \end{equation} The expressions (\ref{hr}, \ref{fK}) are fits to the results of direct Monte Carlo solutions to the Boltzmann Equation for $n=0$ in (\ref{pdep}, \ref{vdep}). In principle similar solutions should be obtained and fitted for $n\ne0$, but carrying out such Monte Carlo simulations is well beyond the scope of this paper. We expect that the form of the suppression functions will be broadly similar to (\ref{hr}, \ref{fK}) for all $n$, however. Conductive transport is generically expected to be most efficient for some $K$ close to but less than 1, because the larger the inter-scattering distance $l_\chi$ of a WIMP relative to the scale height $r_\chi$, the further it will have travelled (in an effective sense) in the star, and therefore the greater impact it will have on the local luminosity function, when it deposits its energy -- but beyond $K\sim1$, the conduction approximation breaks down, and transport must be less efficient. Transport is also enhanced in the surface regions of a star due to the enhancement of (\ref{LTEdens}) at large $r$ compared to (\ref{isodens}), and suppressed in the stellar core generically, as the radial height is actually less than $l_\chi$, causing the impacts of conductive transport to be predominantly felt `downstream' (i.e.\ at radii larger than $l\kappa(\mu)/\kappa(0)$; see \cite{GouldRaffelt90a} for details). For the calculation of WIMP conductive luminosities, we require the typical inter-scattering distance (\ref{lchi}), which depends on the local velocity-averaged scattering cross-section. Taking a Maxwell-Boltzmann distribution for the velocities of the WIMPs and nuclei once more, the distribution of their relative velocities, in normalised units, is \begin{equation} F_{\chi-n}(\vect x - \vect z) = \left[\pi(1+\mu)\right]^{-3/2}e^{-\frac{|\vect x - \vect z|^2}{1+\mu}}. \end{equation} This then gives: \begin{eqnarray} \left\langle\sigma_{v^2}\right\rangle &= \sigma_0 \zeta^{-2} \left\langle \frac{\vrel^2}{v_T^2} \right\rangle &= \frac32\sigma_0\zeta^{-2}(1+\mu), \label{sigmav2}\\ \left\langle\sigma_{v^4}\right\rangle &= \sigma_0 \zeta^{-4} \left\langle \frac{\vrel^4}{v_T^4} \right\rangle &= \frac{15}{4}\sigma_0\zeta^{-4}(1+\mu)^2, \\ \left\langle\sigma_{v^{-2}}\right\rangle &= \sigma_0 \zeta^2 \left\langle \frac{v_T^2}{\vrel^2} \right\rangle &= 2\sigma_0\zeta^{2}(1+\mu)^{-1}, \end{eqnarray} where $\zeta \equiv v_0/v_T$. The averaged $q$-dependent cross-sections follow from these, as $\langle b^{2n} \rangle = \langle (\vrel/v_0)^{2n} \rangle/(1+\mu)^{2n}$: \begin{eqnarray} \left\langle\sigma_{q^2}\right\rangle &=& 6 \sigma_0 \zeta^{-2} (1+\mu)^{-1}, \\ \left\langle\sigma_{q^4}\right\rangle &=& 40 \sigma_0 \zeta^{-4} (1+\mu)^{-2}, \\ \left\langle\sigma_{q^{-2}}\right\rangle &=& \sigma_0 \zeta^2 (1+\mu), \label{sigmaqm2} \end{eqnarray} where for $q$-dependent cross-sections $\zeta \equiv q_0/ (\mx v_T)$. We have again regulated the divergence in the $q^{-2}$ case using the momentum transfer cross-section. Here we take the opportunity to point out that our results for $\alpha$ and $\kappa$ are fully independent of $\zeta$, and therefore $v_0$ and $q_0$. To apply them in cases where $v_0$ or $q_0$ differ from the canonical values we have assumed in this paper, one need only use the desired $v_0$ or $q_0$ when implementing (\ref{LTEtransport}) and the appropriate one of (\ref{sigmav2})--(\ref{sigmaqm2}). To properly compute the effect of WIMPs in the Sun, we will need to include the conduction coefficients in a full solar evolution code that also includes a detailed computation of the velocity/momentum-dependent capture, evaporation and annihilation rates (if any); that work will appear soon \cite{Vincent14}. For the purpose of this paper, we simply assume a steady state with a set number of WIMPs per baryon in the present Sun ($n_\chi/n_\mathrm{b}$), in order to directly compare the conductive efficiencies of cross-sections with different velocity and momentum scalings. Likewise, we simply choose a few benchmark values of $\mx$, $\sigma_0$, $v_0$ and $q_0$ at which to make these comparisons, and assume for the sake of example that $\sigma_0$ scales with nuclear mass as it does for isospin-conserving spin-independent couplings (\ie proportional to the square of the atomic number). \begin{figure*}[htbp] \begin{tabular}{c@{\hspace{0.04\textwidth}}c} \includegraphics[width=0.48\textwidth]{figures/epsilon_v_m_1} & \includegraphics[width=0.48\textwidth]{figures/epsilon_v_m_20} \\ \end{tabular} \caption{\textit{ Absolute energy transported $|\epsilon|$ by WIMPs of mass $\mx = 1$ GeV (left) and $\mx = 20$ GeV (right) with a velocity dependent cross-section. Transported energy is negative at small $R$ (below the dip) and becomes positive at larger $R$, representing transport from the solar core into outer layers. We have taken the AGSS09ph solar model \cite{Serenelli:2009yc} with $n_\chi/n_\mathrm{b}=10^{-16}$ WIMPs per baryon, $v_0 = 110$ km/s, and $\sigma_0 = 10^{-39}$ cm$^{2}$.}} \label{fig:etransv} \end{figure*} In Figure \ref{fig:etransv} we illustrate the effect on energy transport by 1\,GeV\footnote { Although the tendency in solar dark matter literature at present is to assume that evaporation is efficient below 4\,GeV and inefficient above it, this is a simplification of a result obtained by Gould \cite{Gould87a} in the context of now-defunct models designed to explain the solar neutrino problem. The reality is far more nuanced than this naive approximation would suggest: the efficiency of evaporation depends sensitively upon mass-matching between different nuclei in the Sun and the WIMP, as well as the thermal velocity widths of the nuclei and their radial stratification, and the actual Knudsen number of the system, which depends directly on the cross-section itself. This sensitivity will be further modified by any velocity or momentum-dependence to the cross-section. These effects are ignored in even the best modern analyses, so the true impact of evaporation for modern dark matter models, both at low and high WIMP masses, still remains to be reliably calculated. } and 20\,GeV WIMPs in the Sun, using $n_\chi/n_\mathrm{b} = 10^{-16}$ WIMPs per baryon, $v_0 = 110$ km/s, and $\sigma_0 = 10^{-39}$ cm$^{2}$. These roughly correspond to results of \cite{Taoso10}, which provide some small effects on solar structure. The combined effect of $\alpha$, $\kappa$ and $l_\chi$ gives a suppression of energy transport in the case of a $\vrel^2$ or $\vrel^4$ cross-section, but an enhancement for the $\vrel^{-2}$ case, due to the increased scattering at low velocity. \begin{figure*}[tbp] \begin{tabular}{c@{\hspace{0.04\textwidth}}c} \includegraphics[width=0.48\textwidth]{figures/epsilon_q_m_1} & \includegraphics[width=0.48\textwidth]{figures/epsilon_q_m_20} \\ \end{tabular} \caption{\textit{ Absolute energy transported $|\epsilon|$ by WIMPs of mass $\mx = 1$ GeV (left) and $\mx = 20$ GeV (right) with a momentum-dependent cross-section and $q_0= 40$ MeV. We have taken the AGSS09ph solar model \cite{Serenelli:2009yc} with $n_\chi/n_\mathrm{b}=10^{-16}$ WIMPs per baryon, and $\sigma_0 = 10^{-39}$ cm$^{2}$.}} \label{fig:etransq} \end{figure*} Figure \ref{fig:etransq} shows the absolute energy transported by WIMPs with a momentum-dependent cross-section for $\mx = 1$\,GeV and $\mx = 20$\,GeV, again with $n_\chi/n_\mathrm{b} = 10^{-16}$ WIMPs per baryon and $\sigma_0 = 10^{-39}$ cm$^{2}$. Here we use $q_0 = 40$\,MeV, which corresponds to a nuclear recoil energy of $E_R = 11$\,keV in a germanium direct detection experiment, and 29\,keV in a silicon detector. It is apparent that at higher masses, a $q^{2n}$ cross-section with negative $n$ can result in heat transport that is enhanced by many orders of magnitude relative to standard WIMPs with a constant cross-section. At lower masses, the opposite is true: negative $n$ suppresses heat transport, whereas positive $n$ can greatly enhance it. \begin{figure*}[tbp] \begin{tabular}{c@{\hspace{0.04\textwidth}}c} \includegraphics[width=0.48\textwidth]{figures/etrans1MeV} & \includegraphics[width=0.48\textwidth]{figures/etrans40MeV}\\ \end{tabular} \caption{\textit{Total transferred energy $ r_\odot^{-3}\int_0^{r_\odot} |\epsilon(r)|r^2 \ud r$ as a function of the WIMP mass for momentum-dependent scattering. Here $\sigma_0 = 10^{-39}$ cm$^2$, $n_\chi/n_b = 10^{-16}$ WIMPs per baryon and $q_0 = 1$\,MeV (left) or $q_0 = 40$\,MeV (right). These reference momenta can be compared with nuclear recoil energies in direct detection experiments using the expression $q_0 = \sqrt{2 m_{\rm nuc} E_R}$. For recoils on Germanium, for instance, they correspond to nuclear recoil energies of 7 eV and 11 keV, respectively. These show that, depending on the $q$ region probed by direct detection experiments, solar observations may be far more, or far less sensitive than underground detectors.}} \label{fig:eps_mdep} \end{figure*} Figure \ref{fig:eps_mdep} illustrates this point in more detail. Here we show the full dependence of the total transported energy on the WIMP mass $\mx$, for each type of $q$-dependent scaling. We also compare this behaviour for two different example values of $q_0$. For $q_0 = 1$\,MeV -- corresponding to recoil energies in detectors of only a few eV -- the behaviour of $\epsilon$ as a function of the sign of $n$ is reversed below $\mx \sim 2$\,GeV. The value of $q_0 = 40$\,MeV that we choose for the right-hand figure is closer to the region probed by underground direct detection experiments, and shows a reversal at slightly higher masses, around $\mx \sim 4$\,GeV. This shows that for the mass region probed by such experiments, the effects of solar heat transport can be much larger than previously computed in the $\sigma = const.$ case and that for such models, the comparison between experiments and solar effects should be made on a case-by-case basis. In contrast with the behaviour shown in Figure \ref{fig:eps_mdep} for $q$-dependent scattering, the total energy transported by $\vrel$-dependent scattering does not show a reversal with respect to $n$ when the WIMP mass is changed. This can also be seen by comparing Figures \ref{fig:etransv} and \ref{fig:etransq}: for $q$, the impact depends on WIMP mass, whereas for $\vrel$-dependent scattering, negative $n$ enhances energy transport and positive $n$ decreases it, independent of the mass. The qualitative difference between the impacts of the WIMP mass for $\vrel$ and $q$-dependent cross-sections is due to the additional factor of $(1+\mu)^2$ in the thermally-averaged cross-section for $q$-dependent scattering. This adds an additional dependence on WIMP mass to the typical inter-scattering distance, Knudsen number, and therefore the total $K$-dependent luminosity suppression. Again, we stress that to compute an accurate profile of energy transport, one should include the feedback on solar structure itself during stellar evolution. \section{Discussion and Conclusion} \label{sec:conclusion} Figures \ref{fig:etransv}--\ref{fig:eps_mdep} contrast the potential impacts on energy transport in the Sun of nuclear scattering cross-sections with different momentum and velocity scalings. The differences we see suggest that solar physics will be a very useful complement to direct detection experiments in ruling out or confirming specific forms of $\sigma(\vrel,q)$. It is worth noting that the cross-sections which give rise to enhanced energy transport will also enhance or suppress the capture of DM by the Sun in the first place. In order to become gravitationally bound to the Sun, a free WIMP must lose enough kinetic energy to fall below the local escape velocity. A scattering cross-section that goes as $\vrel^{-2}$ would boost the chance of a free WIMP at the low end of the Galactic DM velocity distribution scattering off a nucleus, whereas a $q^{2n}$ cross-section with $n > 0$ would enhance the rate of collisions leading to large enough energy losses for the WIMP's speed to fall below $v_{\rm esc}$. The combination of enhanced capture and heat transport is therefore expected to lead to observable effects on the solar structure for some models, even for very small scattering cross-sections. Finally we point out the possible application of this enhanced heat transport to the so-called Solar Abundance Problem. It has been suggested \cite{Frandsen10,Taoso10} that at intermediate radii, $\mx = 5$\,GeV DM can partly reconcile the discrepancy between the results of solar models computed with the latest abundances, and the sound speed profile inferred from helioseismology. In this scenario, DM accumulation is usually maximised by taking it to be asymmetric (non-annihilating). Simulations \cite{Cumberbatch10} indicate that the effects of energy transported from the core can propagate to higher radii and affect the sound speed up to the base of the convection zone. However, to do so without the effects of momentum- or velocity-dependent couplings, the scattering cross-section must be extremely high ($\sim$$10^{-35}$\,cm$^2$), in order to accumulate sufficient quantities of DM to have an observable effect. The enhancement provided to capture and conduction by $\vrel^{2n}$ or $q^{2n}$ scattering of DM on nuclei has the potential to finally make DM a viable solution to the abundance problem. There is thus a strong possibility that models such as Sommerfeld-enhanced or dipole DM, whose motivation stem from different areas of DM phenomenology, are particularly suited to resolving this outstanding problem in solar physics. The framework developed above -- an extension of the process established by Gould and Raffelt \cite{GouldRaffelt90a} -- allows the computation of the thermal conduction parameters $\alpha$ and $\kappa$ for DM particles whose interactions with nuclei are velocity or momentum-dependent. We have computed these parameters explicitly for $\sigma \propto \vrel^{2n}$ and $q^{2n}$, for $n = \{-1,1,2\}$. Finally, we have shown the impact on heat transfer in the Sun with respect to the fiducial constant $\sigma$ case, for parameters currently allowed by direct detection experiments. Our results indicate that this impact could be rather substantial. In a follow-up paper \cite{Vincent14}, we plan to incorporate these results into state-of-the-art solar simulation software to include realistic DM capture and the feedback on solar structure itself. \textit{Note added: } As we were finalising this paper for resubmission, ref.\ \cite{Lopes14} appeared on the arXiv, investigating the impacts on the solar structure of a model with a massive force carrier, similar to the $\vrel^{-2}$ case with a cuttoff $\omega$ that we studied. That paper employed the original formalism of \cite{GouldRaffelt90a}, which we have shown here to be invalid for such dark matter models, and considered only mean WIMP velocities rather than full velocity distributions as we do here. Nonetheless, the authors did employ a full solar evolutionary model, calculate the capture and annihilation rates for their model, and find promising results with regards to the impacts on the Sun and the potential for solving the Solar Abundance Problem. This makes an analysis of solar models with velocity- and momentum-dependent energy transport correctly implemented on the basis of the results in this paper \cite{Vincent14}, all the more interesting a prospect. \section*{Acknowledgements} We thank Aldo Serenelli for useful discussions. AV acknowledges support from FQRNT and European contracts FP7--PEOPLE--2011--ITN and PITN--GA--2011--289442--INVISIBLES. PS is supported by a Banting Fellowship, administered by the Natural Science and Engineering Research Council of Canada.
\section{Diffraction - a historical perspective} Diffractive scattering of electrons on various nuclei provided important insights to the charge density distributions of spherical nuclei. The detailed analysis resulted in simple observations by Hofstadter and colleagues, that were summarized in the Nobel lecture of Hofstadter as follows: \begin{itemize} \item The volume of spherical nuclei is proportional to the mass number $A$. \item The surface thickness is constant, independent of $A$. \end{itemize} These observations revealed structures in atomic nuclei on the femtometer scale. They imply also that the central charge density of large nuclei is approximately constant. For more details, we recommend the Nobel Lecture (1961) by R. Hofstadter ~\cite{HofstadterNobel}. The results summarized there were one of the first observations of images on the femtometer scale, corresponding to nuclear charge density distributions. The more recent historical overview of ref.~\cite{Glauber:2006gd} discusses applications of multiple diffraction theory to high energy particle and nuclear physics. Recently, with 7 and 8 TeV colliding energies of proton-proton reactions at CERN LHC, the resolution of diffractive images in elastic proton-proton scattering reached the sub-femtometer scales, as we demonstrate below. Our talk at the Low-X 2013 conference discussed two model classes: the Bialas-Bzdak and the Glauber-Velasco models. Our conference contribution follows the lines of that presentation, except for the details of on results from the Bialas-Bzdak model, for which we direct the interested readers to suitable references. \section{Diffraction in quark-diquark models} In a series of papers, Bialas and Bzdak discussed a quark-diquark model of elastic proton proton scattering~\cite{Bialas:2006qf,Bzdak:2007qq,Bialas:2006kw,Bialas:2007eg}. Recently, this Bialas-Bzdak or BB model was tested in details on elastic proton-proton scattering data both at the ISR and LHC energies~\cite{Nemes:2012cp} and developed further to obtain a more realistic description of the dip region of elastic pp scattering~\cite{CsorgO:2013kua}. In the BB model, the differential cross-section of elastic proton-proton scattering is given by the following formula \begin{equation} \frac{d\sigma}{dt}=\frac{1}{4\pi}\left|T\left(\Delta\right)\right|^2, \label{e:1} \end{equation} where $\Delta=|{\vec\Delta}|$ is the modulus of the transverse component of the momentum transfer. In the high energy limit, $s\rightarrow\infty$, $\Delta^{2} \simeq -t$, where $t$ is the squared four-momentum transfer. The amplitude of the elastic scattering in momentum space, $T(\vec{\Delta})$ is given by the Fourier-transform of the amplitude in impact parameter space, \begin{equation} T(\vec{\Delta})=\int\limits^{+\infty}_{-\infty}\int\limits^{+\infty}_{-\infty}{t_{el}(\vec{b})e^{i\vec{\Delta} \cdot \vec{b}}\text{d}^2b}= 2\pi\int\limits_0^{+\infty}{t_{el}\left(b\right)J_0\left(\Delta b\right)b {\text d}b}, \label{e:2} \end{equation} where the impact parameter is denoted by ${\vec b}$ and $b=|\vec{b}|$. From unitarity conditions one obtains \begin{equation} t_{el}(\vec{b})=1-\sqrt{1-\sigma(\vec{b})}.\label{e:3} \end{equation} The inelastic proton-proton cross-section in the impact parameter space for a fixed impact parameter $\vec{b}$ is given by the following integral \begin{equation} \sigma(\vec{b})=\int\limits^{+\infty}_{-\infty}...\int\limits^{+\infty}_{-\infty} {\text{d}^2s_q \text{d}^2s'_q \text{d}^2s_d \text{d}^2s'_d D(\vec{s_q},\vec{s_d}) D(\vec{s_q}',\vec{s_d}') \sigma(\vec{s_q},\vec{s_d};\vec{s_q}',\vec{s_d}';\vec{b})}, \label{e:4} \end{equation} where the integral is taken over the two-dimensional transverse position vectors of the quarks $\vec{s_{q}}$, $\vec{s_{q}}'$ and diquarks $\vec{s_{d}}$, $\vec{s_{d}}'$. \section{Bialas - Bzdak model of elastic pp scattering} The BB model approximates the quark-diquark distribution inside the proton with a Gaussian form~\cite{Bialas:2006qf,Bzdak:2007qq,Bialas:2006kw,Bialas:2007eg} \begin{equation} D\left(\vec{s_q},\vec{s_d}\right)=\frac{1+\lambda^2}{\pi R_{qd}^2}e^{-(s_q^2+s_d^2)/R_{qd}^2}\delta^2(\vec{s_d}+\lambda \vec{s_q}),\;\lambda=m_q / m_d,\label{e:5} \end{equation} and, in order to define a model that can be analytically integrated and compared to data in a straight-forward way, the model is formulated in simple and if possible Gaussian terms. The BB model also supposes that protons are scattered elastically if and only if all of their constituents are scattered elastically \begin{equation} \sigma(\vec{s_q},\vec{s_d};\vec{s_q}',\vec{s_d}';\vec{b}) = 1-\prod_{a,b \in \{q,d\}}\left[1-\sigma_{ab}(\vec{b}+\vec{s_a}'-\vec{s_b} )\right], \label{e:6} \end{equation} where the inelastic differential cross-sections of the constituents are parametrized with Gaussian distributions as well \begin{equation} \sigma_{ab}\left(\vec{s}\right) = A_{ab}e^{-s^2/R_{ab}^2},\;R_{ab}^2=R_a^2+R_b^2. \label{e:7} \end{equation} Here $A_{ab}$ are suitably chosen normalization factors, $a,b$ stand for $q,d$, denoting quarks and/or diquarks, while $R_q$ and $R_d$ stand for the Gaussian radii, that characterize in the BB model the quark and diquark inelastic scattering cross-sections, respectively. \begin{figure}[H] \includegraphics[width=0.40\textwidth]{Figs/single.eps} \includegraphics[width=0.40\textwidth]{Figs/qq.eps} \centering \caption{Scheme of the scattering of two protons, when the proton is assumed to have a quark-diquark structure. The diquark is assumed to be scattered as a single entity (left) or as composition of two quarks (right). This figure is a snapshot and all the model parameters follow a Gaussian distribution. Note, that a center of mass energy dependent Lorentz-contraction determines the longitudinal scale parameters.} \label{scattering sit1} \end{figure} This BB model comes in two different realizations, corresponding to two different pictures of the proton: in one of the cases, the diquark is assumed to have a structureless Gaussian distribution, while in the other case, the diquark is assumed to scatter as a loosely bound state of two correlated quarks. These scenarios are indicated by the $ p = (q, d) $ and the $p = (q, (q,q))$ labels. As noted by Bialas and Bzdak, models with three uncorrelated quarks in the proton, labelled as $ p= (q,q,q)$ were tested before at ISR energies and they are known to fail, with other words, we know that the quarks are correlated inside the protons~\cite{Czyz:1969jg}. In its original form, the BB model has been integrated analytically, both for the $p = (q,d)$ and the $p=(q,(q,q)) $ scenarios, {\it assuming} that the real part of forward scattering is negligible. \begin{figure}[H] \includegraphics[width=0.49\textwidth]{Figs/c223.eps} \includegraphics[width=0.49\textwidth]{Figs/c123.eps} \centering \caption{(Color online.) Results of Minuit fits of both versions of the Bialas-Bzak model at ISR energies. Left panel indicates the scenario $p = (q,d)$, where the diquark is assumed to scatter as a single entity while the right panel indicates the scenario $p = (q,(q,q))$, where the diquark inside the proton is considered to be a scattering object consisting of two quarks.} \label{Figure:2} \end{figure} The two panels of Figure ~\ref{Figure:2} indicate CERN Minuit fit results of the BB model to differential cross-section data on elastic proton-proton scattering at the ISR energy of $\sqrt{s}= 23.5$ GeV. Left plots correspond to the scenario $p = (q, d)$ while the right panel stands for the scenario $p = (q, (q, q))$. The top parts show the data points and the result of the best fit, while the lower parts indicate the relative deviation of the model from data in units of measured error bars. As the original BB model is singular around the dip, 3 data points were left out from the optimalization, that are located closest to the diffractive minimum, and indicated with filled (red) circles in Figure~\ref{Figure:2}. The fit range was restricted to $0.36 \le - t \le 2.5$ GeV$^2$, so that a fair comparison could subsequently be made with the first TOTEM results on proton-proton elastic scattering at LHC energy of 7 TeV of ref. ~\cite{Antchev:2011zz}. The best fits are shown with a solid (black) line in the fitted range, while their extrapolation to lower $t$ values are also shown, with dashed (green) lines. The confidence levels, after fixing the values of $\lambda$ and $A_{qq}$ to 0.5 and 1, respectively, come very close to 0.1\%, indicating that the fit quality is similar, statistically acceptable in both scenarios. Similar fit qualities were reported at each ISR energies of 30.7 GeV, 52.8 GeV and 62.5 GeV, see ref. ~\cite{Nemes:2012cp} for details. Figure ~\ref{Figure:3} shows the comparison of the BB model to TOTEM data on elastic pp scattering at 7 TeV LHC energies, indicating a qualitative change, as compared to the fit results at ISR energies: the quality of this fit is statistically not acceptable, CL is significantly below 0.1\%, and the fit deviates from the data in particular in the dip region. The bottom parts indicate, that the shape of the differential cross-section in the dip region, around the first diffractive minimum is not reproduced correctly by the original BB model at 7 TeV LHC energies, and, as also can be seen on this Figure ~\ref{Figure:3}, this shortcoming cannot be fixed by leaving out a few data points around the diffractive minimum from the optimalization procedure. The details of these BB fits are described in ref. ~\cite{Nemes:2012cp}. \begin{figure}[H] \includegraphics[width=0.49\textwidth]{Figs/c27000.eps} \includegraphics[width=0.49\textwidth]{Figs/c17000.eps} \centering \caption{(Color online.) The result of the fit of the original version of the Bialas-Bzdak model at $7$ TeV LHC energies in two different scenarions: left panel stands for the $p=(q,d)$ scenario, when the diquark is assumed to scatter as a single entity, while the right panel stands for the $p=(q,(q,q))$ case, when the internal structure of the diquark is resolved as a correlated system of two quarks. } \label{Figure:3} \end{figure} Recently, two of us generalized the Bialas-Bzdak model by adding a small real part to the forward scattering amplitude, to investigate, if the description of the dip region can be improved can be made statistically acceptable in this way. The results of this scenario are described in detail in ref.~\cite{CsorgO:2013kua}. A {\it small} real part was added to the forward scattering amplitude by using an analogy of with the Glauber-Velasco model, and assuming that even if all the parton level scatterings are elastic, the proton-proton scattering can, with a small probability, become inelastic. In this manner, a parton level $\rho$ parameter was introduced. The results, detailed in ref.~\cite{CsorgO:2013kua}, indicate that a small real part indeed improves the agreement of the BB model with data in the dip region, and the fits become statistically acceptable in the whole $t$ region, including all the data points from dip region, {\it if} the energy of the collisions is limited to the ISR energy range of $\sqrt{s} = 23.5 $ - $ 62.5 $ GeV. At the LHC energy of $\sqrt{s} = 7$ TeV, the generalized Bialas-Bzdak or the $\alpha$BB model resulted in an improvement, that reduced the disagreement between the BB model and the data substantially and filled the dip region rather dramatically. However, even this improvement did not result in a statistically acceptable fit quality to the differential cross-section of elastic proton-proton collisions at this LHC energy, although good quality fits were obtained if the fit range was limited to the dip region. As a consequence, we kept on searching for a model that is able to describe elastic pp scattering data at LHC energies, and investigated the performance of the Glauber-Velasco model~\cite{Glauber:1984su}. Before reporting the results, let us summarize what we have learned till now from the detailed fits using the original Bialas-Bzdak or BB model, and its generalized version when a small real part is added to its forward scattering amplitude, see refs.~\cite{Nemes:2012cp,CsorgO:2013kua} for further details. \section{What have we learnt so far ?} The original version of the Bialas-Bzdak model gave a statistically acceptable description of elastic pp scattering data at ISR energies, if the data points close to the diffractive minimum were left out from the fit. If these data points were included and also a small real part was added to the model, as detailed in ref.~\cite{CsorgO:2013kua}, the fits at the ISR energies from $\sqrt{s} = $ 23.5 GeV to 62.5 GeV become statistically acceptable, good quality fits, in the fit range of 0.36 $\le -t \le $ 2.5 GeV$^2$. Two model parameters could be fixed at all energies ($A_{qq} = 1$ and $\lambda = 1$) while maintaining the statistically acceptable fit quality. The parameter $\alpha$, that was introduced as a parton level ratio of the real to imaginary part of the forward scattering amplitude, remained indeed in the region of very small values, $\alpha = 0.01 \pm 0.01$ except at 52.8 GeV, where $\alpha = 0.02 \pm 0.01$ value was found. Although these $\alpha$ parameters are within errors consistent with zero, a small but non-vanishing value provided qualitatively better fits in the dip region, as detailed in ref.~\cite{CsorgO:2013kua}. The best fit parameters, that described the quark structure of the protons geometrically, took also rather interesting values. For example, the quark radius $R_q$ within 2 standard deviations was consistent with an energy independent value of $R_q = 0.27 \pm 0.01$ fm. The diquark size indicated a nearly constant value, varying between $R_d = 0.71 \pm 0.01$ to $0.77 \pm 0.01$ fm, slighlty increasing with increasing $\sqrt{s}$. Although the fit to the TOTEM data 7 TeV were not statistically acceptable, the best parameter values for the quark and diquark radii were in the same range, except a slight decrease of the diquark size in the $p = (q,(q,q))$ model at 7 TeV. We observed that the biggest variation, when the energy is increased to 7 TeV, is observable in the scale that measures the typical quark-diquark distance, $R_{qd}$. This value was in the range of $R_{qd} = 0.23\pm 0.01$ fm at ISR energies in the $p = (q, (q,q)) $ model, while it increased to the value of $0.73 \pm 0.01 $ fm at 7 TeV. Similar trend of increasing quark-diquark separation is seen in the $p = (q,d)$ scenario. Graphically, the evolution of the proton elastic scattering structure is illustrated on Figure~\ref{Figure:3}, where the best fit parameters are also indicated on the sub-plots, generated for the case of the $p=(q,(q,q))$ scenario. The same qualitative behaviour of increasing quark-diquark distance is observed also in the $p = (q, d)$ picture, see ref.~\cite{CsorgO:2013kua} for further details. \begin{figure}[H] \centering \includegraphics[trim = 6mm 12mm 2mm 4mm, clip, width=0.4\linewidth]{Figs/c23.5.eps} \includegraphics[trim = 6mm 12mm 2mm 4mm, clip, width=0.4\linewidth]{Figs/c30.7.eps}\\ \includegraphics[trim = 6mm 12mm 2mm 4mm, clip, width=0.4\linewidth]{Figs/c52.8.eps} \includegraphics[trim = 6mm 12mm 2mm 4mm, clip, width=0.4\linewidth]{Figs/c62.5.eps}\\ \includegraphics[trim = 6mm 12mm 2mm 4mm, clip, width=0.4\linewidth]{Figs/c7000.eps} \caption{Visualization of the fit results of Bialas-Bzdak models, extended to a small real part, for the case of $p = (q, (q, q)) $, when the diquark is assumed to be resolvable as a weekly bound state of two quarks. The main effect of increasing $\sqrt{s}$ is apparently the increasing value of $R_{qd}$, the typical quark-diquark distance.} \label{Figure:4} \end{figure} As discussed both in refs.~\cite{Nemes:2012cp} and ~\cite{CsorgO:2013kua}, the $p = (q,d)$ and the $p = (q, (q,q))$ models provide similar quality of data description both at ISR energies (where they are both statistically acceptable) and at 7 TeV LHC energy (where both fail to describe TOTEM data in a statistically acceptable manner). Nevertheless we compare the best fit values of the different energies, to try to get a qualitative insight assuming, that the missing element of the model will not modify drastically the best fit parameters at LHC energies. Given that the $p = (q,d)$ and the $p = (q,(q,q))$ Bialas-Bzdak models correspond to two different assumptions about the internal structure of the protons, it was a kind of surprize for us, that the measured total pp cross-section $\sigma_{tot}$ was phenomenologically related to the parameters of the BB model in a model-independent way, i.e. the following relation is approximately valid for both scenarios: \begin{equation} \sigma_{tot} \approx 2 \pi R_{eff}^2 \,\, = \,\, 2 \pi (R_q^2 + R_d^2 + R_{qd}^2). \end{equation} This approximation was found to be valid within a relative error of about 9 \% at ISR energies, while at the LHC energies it yields only an ball-park value, order of magnitude estimation ($\frac{\sigma_{tot}}{2 \pi R_{eff}^2} = 1.42$). We also have observed an interesting scaling property of the differential and the total proton-proton elastic scattering cross-section, namely the product of the total cross-section times the $t$ of the dip is within errors a constant: \begin{equation} t_{dip}\sigma_{tot} \approx C \end{equation} where $C = 54.8 \pm 0.7$ mb GeV$^2$ from a fit. We find that this relation is valid within 5 \% relative error at each ISR and also at 7 TeV LHC energies. A similar relation holds for a light scattering from a black disc, however, with a significantly different constant value, $C_{blackdisc} \approx 35.9 $ mb GeV$^2$. \begin{figure}[H] \centering \includegraphics[width=0.48\textwidth]{Figs/tdip_sigmatot.eps} \caption{ The $\frac{|t_{dip}| \cdot \sigma_{tot,exp}}{C}$ ratio, directly obtained from experimental data. The dashed line indicates $1$, which value within errors is consistent with all the data from $\sqrt{s} = 23.5 $ GeV to 7 TeV. } \label{Figure:5} \end{figure} Given that there are theoretically well established formulas for the description of the rise of the total pp scattering cross-section with increasing energies, the above formula can be well used to predict the position of the (first) minimum or the dip in the differential cross-section of pp collisions and also can be extrapolated, or, predicted for pA and AB collisions ~\cite{inprep}. Given that we could not find a statistically acceptable quality fit with the Bialas-Bzdak model to 7 TeV TOTEM data on elastic pp scattering at LHC, neither in the original form, nor when a small real part is added to the forward scattering amplitude of this model, we started to look for alternative interpretations and derivations of $d\sigma/dt$. One possibility is to allow for not only small values of the real part of the forward scattering amplitude, but still keep the basic structure of the Bialas-Bzdak model. The studies in this direction will be reported elsewhere. In the next section we report about the other natural direction, that we investigated in detail. In particular, when we added a small real part to the forward scattering amplitude to the Bialas-Bzdak model in ref.~\cite{CsorgO:2013kua}, we were introducing a parton level $\rho$ parameter inspired by the Glauber-Velasco model of refs.~\cite{Glauber:1984su,Glauber:1987sf}. In the next section, we summarize this model and report about its first comparisions to TOTEM data. \section{Overview of the model of Glauber and Velasco} In this section, we follow the lines of the presentation of the Glauber-Velasco model, as described in refs.~\cite{Glauber:1984su,Glauber:1987sf}. The Glauber diffractive multiple scattering theory is utilized to describe elastic collisions of two nucleons, which are pictured as clusters of partons. The parton distributions are assumed to have form factors given by the experimentally measured electric charge form factors. Differential cross sections calculated in this way showed good agreement with the experimentally measured ones over a broad range of $pp$ and $p\bar{p}$ energies, when the parton-parton scattering amplitude is given a suitable parametrization~\cite{Glauber:1984su,Glauber:1987sf}. The range of the parton-parton interaction derived from these data is found to increase steadily with energy. The absorption processes that take place are localized in the overall nucleon-nucleon interaction by calculating the shadow profile function. The emerging picture corresponded to an opaque region of interaction that grows in radius with increasing energy. The surface region of the interaction seems however to maintain a remarkably fixed shape as the radius grows. In the multiple diffraction theory of Glauber and Velasco, the elastic scattering amplitude for diffractive collisions can be written as an impact parameter integral \begin{equation} F(t)=i\int_0^{\infty} J_0\left(b\sqrt{-t}\right)\left\{1-\exp\left[ -\Omega\left(b\right)\right]\right\}bdb. \end{equation} Any particular model is characterized by the opacity function $\Omega(b)$, which in general may be a complex valued function. If we picture the two colliding nucleons as clusters of partons that scatter one another with the averaged scattering amplitude $f(t)$, then the opacity function can be written in the form of an integral over momentum transfers $q$, \begin{equation} \Omega\left(b\right)=\frac{\kappa}{4\pi} \left(1-i\alpha\right)\int_0^{\infty} {J_0\left(q\,b\right)G_{p,E}^2\left(-t\right)}\frac{f(t)}{f(0)}q dq, \end{equation} where $q = \sqrt{-t}$. The constants $\kappa$ and $\alpha$ in this expression are real-valued and need to be determined empirically. The function $G_{p,E}(t)$ is the form factor for the parton density in the proton. Following ref.~\cite{Glauber:1987sf}, we shall assume it to be the same as the observed electric form factor for the proton. One choice of parametrization we have investigated is \begin{equation} \frac{f(t)}{f(0)} = \frac{e^{i\left(b_1\left|t\right|+b_2\,t^2\right)}} {\sqrt{1+a\left|t\right|}}.\\ \end{equation} The BSWW form factor, corresponding to the distribution of electric charge in the proton, is described with a four-pole parametrization~\cite{Borkowski_1975} \begin{equation} G_{p,E}\left(q^2\right)=\sum_{i=1}^{n}\frac{a_i^{E} \left(m_{i}^{E}\right)^2} {\left(m_{i}^{E}\right)^2 + q^2}\,,\, \sum_{i=1}^{n}a_i^E=1\,,\, G_{p.E}(0)=1.\\ \label{BSWW_four_pol_parametrization} \end{equation} The differential cross-section for elastic pp collisions is evaluated as \begin{equation} \frac{d\sigma_{el}}{d\left|t\right|}=\pi\left|F(t)\right|^2. \end{equation} The parameters of the BSWW form factor are given by the following table: \begin{table}[!h] \centering \begin{tabular}{|r|r|} \hline $a_i^E$ & $(m_i^E)^2$ (fm$^{-2}$) \\ \hline 0.219 & 3.53 \\ \hline 1.371 & 15.02 \\ \hline -0.634 & 44.08 \\ \hline 0.044 & 154.20 \\ \hline \end{tabular} \caption{Best fit parameters~\cite{Borkowski_1975} of the four-pole fit of Eq.~(\ref{BSWW_four_pol_parametrization}).} \end{table} \begin{figure}[H] \centering \includegraphics[width=0.49\linewidth]{Figs/Glauber_pp_7000_from_0_36_to_2_5.eps} \caption{Glauber-Velasco model fit to the differential cross-sectio of proton-proton elastic scattering data at $7$ TeV, in the range of $0.36 \le -t \le 2.5 $ GeV$^2$. } \label{Figure:6} \end{figure} Figure~\ref{Figure:6} indicates, that the Glauber-Velasco model is able to describe successfully the differential scattering cross-section of elastic pp collisions at the 7 TeV LHC energies: the fit quality is statistically acceptable, with CL $> 0.1$ \%. We have tested the model at the ISR energy range of 23.5 GeV - 62.5 GeV too, where the similarly good quality fits were found. The detailed results will be reported in a manuscript that is currently under preparation. \section{Summary} In summary, we have analized elastic proton-proton scattering data from the 23.5 GeV ISR energies to 7 TeV LHC energies, using various forms of the Bialas- Bzdak model. We found that the scenario when the proton is considered to be a quark-diquark state provides a fit quality that is similar to the case when the diquark is resolved as a correlated quark-quark system within the framework of the same model. Adding a small real part to the forward scattering amplitude of the original Bialas-Bzdak model provides a statistically acceptable description of elastic pp scattering data at the ISR energies, however, even this generalized Bialas-Bzdak model fails to describe TOTEM data on elastic pp scattering at 7 TeV. Given that the generalization of the Bialas-Bzdak model followed the lines of the Glauber-Velasco model, we tested also the performance of the Glauber-Velasco model in its original form, and found that it was describing elastic proton-proton scattering both at ISR and at LHC energies when the fit range was restricted to $ 0.36 \le -t \le 2.5$ GeV$^2$. \section{Acknowledgments} T. Cs. would like to thank for professor Glauber for his kind hospitality at Harvard University as well as for his inspiring and fruitful visits to Hungary and CERN, that made the completion of these results possible. He also would like to express his gratitude to the organizers of the Low-X 2013 conference for an invitation and for creating an inspiring and useful meeting. This research was supported by the Hungarian OTKA grant NK 101428 and by a Ch. Simonyi fund, as well as by the Hungarian Academy of Sciences, by a HAESF Senior Leaders and Scholars fellowship and by the US DOE.
\section{Introduction} The recent precise measurements for the neutrino sector indicate that the mixing angle $\theta_{13}$ has finite non-zero value \cite{Abe:2011sj,Adamson:2011qu,Abe:2011fz,An:2012eh,Ahn:2012nd}. This fact indicates the modification of models such as Tri-Bi Maximal mixing \cite{tribimaximal} which derives $\theta_{13}=0$. Several ideas to overcome this problem have been proposed so far, based on flavor symmetries \cite{Altarelli:2010gt,Ishimori:2010au}, perturbation of Pontecorvo-Maki-Nakagawa-Sakata (PMNS) matrix $U_{PMNS}$ from symmetric textures \cite{Wang:2013wya}, texture zeros of neutrino mass matrix $M_{\nu}$ \cite{Frampton:2002yf,Meloni:2012sx, Grimus:2012zm, Fritzsch:2011qv,Lashin:2011dn}, anarchy \cite{Hall:1999sn, Gluza:2011nm} and vanishing minors of $M_{\nu}$ \cite{vanishingminor, Araki:2012ip} etc. For models of two texture zeros in $M_{\nu}$, with four independent parameters, it has been shown \cite{Meloni:2012sx} that the following two patterns \begin{equation} {\rm A}_1~:~M_{\nu}({\rm A}_1)=\left(\begin{array}{ccc} 0&0&\times\\ 0&\times &\times\\ \times&\times&\times\\ \end{array}\right),~~ {\rm A}_2~:~M_{\nu}({\rm A}_2)=\left(\begin{array}{ccc} 0&\times&0\\ \times&\times &\times\\ 0&\times&\times\\ \end{array}\right), \label{a1a2} \end{equation} where $\times$ denotes non-zero matrix elements, require less fine-tuning of parameters to satisfy the current experimental bounds. Moreover it is discussed in Ref. \cite{Grimus:2012zm} that if all parameters are real and if ratio of non-zero matrix elements are small integer (1 or 2), textures given in Eq. (\ref{a1a2}) with two independent parameters are in good agreement with the experimental data. In this paper, we consider a mass matrix $M_{\nu}$ with {\it one} real parameter in the case of textures Eq. (\ref{a1a2}), such as \begin{equation} M_{\nu}=m\left( \begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&3\\ \end{array}\right), \label{mnu0} \end{equation} where $m$ is a mass parameter determined by experimental constraints of two mass-squared differences $\Delta m_{21}^2$ and $\Delta m_{31}^2$ of neutrinos. First we perform numerical calculation in the case of two texture zeros A$_1$ and A$_2$ with four real parameters, and find several candidates of $M_{\nu}$ with one parameter by assuming that ratio of matrix elements are small real number (including integer). After that, we discuss symmetry realization of a candidate Eq.(\ref{mnu0}) by non-Abelian discrete symmetry $A_5'$\cite{Hashimoto:2011tn}. The $A_5'$ symmetry contains three- and five-dimensional irreducible representation ${\bf 3}$ and ${\bf 5}$, and the singlet ${\bf 1}$ from their tensor product ${\bf 3 \cdot 5 \cdot 3}={\bf 1}+\cdots$ enters all the elements of $M_{\nu}$ with desired weights if one assigns ${\bf 3}$ and ${\bf 5}$ for neutrinos and Higgs bosons, respectively { \footnote{ In Ref.\cite{Everett:2008et}, the $A_5$ symmetry is discussed which has similar feature.} }. This paper is organized as follows. In the next section, we perform numerical calculation of neutrino mass matrix $M_{\nu}$ given in Eq.(\ref{a1a2}), and show allowed regions of non-zero matrix elements. Several explicit forms of $M_{\nu}$ with one real free parameter are also given. In Section 3, we discuss symmetry realization of a concrete example of $M_{\nu}$ given in Eq.(\ref{mnu0}) by the binary icosahedral symmetry $A_5'$. Section 4 is devoted to the conclusions. In Appendix, the Higgs potential in our model based on the $A_5'$ symmetry is given. \section{One-parameter Texture of $M_{\nu}$} In this section, we first perform numerical calculation for Majorana neutrino mass matrix $M_{\nu}$ with two zero entries in order to find textures with one real free parameter. Next, we give some examples of $M_{\nu}$ with one parameter based on the numerical calculation and related quantities such as the PMNS matrix and mass squared differences. From a standpoint of flavor symmetry, it is preferable that ratios of non-zero elements of $M_{\nu}$ are simple small real number. \subsection{Numerical Calculation in Two Texture Zeros} We assume two zero entries in $M_{\nu}$ \cite{Frampton:2002yf} in our numerical calculation. As discussed in Refs.\cite{Meloni:2012sx,Grimus:2012zm,Fritzsch:2011qv}, if one includes the CP-violating phases, the textures given in Eq.(\ref{a1a2}) require less fine-tuning of parameters to satisfy the current experimental constraints given below, although seven patterns of $M_{\nu}$, such as A$_{1,2}$, B$_{1,2,3,4}$ and C,{\footnote{Since we focus on the pattern A$_{1}$ and A$_{2}$ due to the conclusions of Refs. \cite{Meloni:2012sx,Grimus:2012zm}, we do not show the explicit form of patterns B$_{1,2,3,4}$ and C. See Refs. \cite{Frampton:2002yf,Fritzsch:2011qv} for the definition of those.}} are consistent with the experiments \cite{Fritzsch:2011qv} . Moreover if all parameters are real, the patterns A$_1$ and A$_2$ with two free parameters show good agreement with the experiments \cite{Grimus:2012zm}. Two matrices in Eq.(\ref{a1a2}) are related to each other by the relation \begin{eqnarray} PM_{\nu}({\rm A}_1)P=M_{\nu}({\rm A}_2),~~{\rm with}~~ P=\left(\begin{array}{ccc} 1&0&0\\ 0&0 &1\\ 0&1&0\\ \end{array}\right), \end{eqnarray} and the both cases lead to the Normal Hierarchy (NH) of the neutrino masses $(m_3>m_2>m_1)$. In order to find texture of $M_{\nu}$ with one real free parameter, we perform numerical calculation in the following procedure; \begin{enumerate} \item We assume all parameters in $M_{\nu}$ to be real, and choose the basis in which the left-handed charged leptons are mass eigenstates. The PMNS matrix $U_{PMNS}$ with vanishing CP-violating phases and neutrino mass eigenvalues $(m_1,m_2,m_3)$ are defined as \begin{eqnarray} M_{\nu}=U_{PMNS}\left( \begin{array}{ccc} m_1&&\\&m_2&\\&&m_3\\ \end{array}\right)U_{PMNS}^T, \label{diag} \end{eqnarray} where $m_2=\pm \sqrt{\Delta m_{21}^2+m_1^2}$ and $m_3=\pm \sqrt{\Delta m_{31}^2+m_1^2}$ with $\Delta m_{ij}^2=m_i^2-m_j^2$. \item We focus on the pattern A$_1$ and A$_2$ given in Eq.(\ref{a1a2}). \item The global fit data \cite{GonzalezGarcia:2012sz} for the case of NH at the 3$\sigma$ level \footnote{See also Refs. \cite{Tortola:2012te,Fogli:2012ua} for the other global fit results.} , \begin{eqnarray} &&\sin^2 \theta_{12}=[0.267,0.344],~ \sin^2 \theta_{23}=[0.342,0.667],~ \sin^2 \theta_{13}=[0.0156,0.0299],\nonumber\\&& \Delta m_{21}^2=[7.00,8.09] \times 10^{-5}~\mbox{eV}^2,~ \Delta m_{31}^2=[2.276,2.695] \times 10^{-3}~\mbox{eV}^2, \label{exp} \end{eqnarray} are used. We randomly input the above constraints except $\Delta m_{31}^2$ into the right-hand side of Eq.(\ref{diag}), and find $m_1$ and $\Delta m_{31}^2$ by solving equations $(M_{\nu})_{11}=0$ and $(M_{\nu})_{12(13)}=0$ for the case A$_{1(2)}$. \end{enumerate} \begin{figure}[h] \begin{center} \includegraphics[width=87mm]{A1112.eps} \includegraphics[width=87mm]{A1113.eps} \vspace{-1cm} \caption{Neutrino mass $m_1$ and $\Delta m_{31}^2$ in the case of $(M_{\nu})_{11}=(M_{\nu})_{12}=0$ (left panel) and $(M_{\nu})_{11}=(M_{\nu})_{13}=0$ (right panel). In both panels, allowed region of $\Delta m_{31}^2$ given in Eq.(\ref{exp}) is represented by light-red (dark) regions.} \label{fig1} \end{center} \end{figure} Figure \ref{fig1} shows the results for the case of A$_1~:~(M_{\nu})_{11}=(M_{\nu})_{12}=0$ (left panel) and A$_2~:~(M_{\nu})_{11}=(M_{\nu})_{13}=0$ (right panel) in the $m_1-\Delta m_{31}^2$ plane. In both panels, allowed region of $\Delta m_{31}^2$ given in Eq.(\ref{exp}) is represented by the light-red (dark) regions. All the dots in both panels satisfy the global fit constraints Eq.(\ref{exp}) except $\Delta m_{31}^2$, and one can see that there exist dots in the allowed region of $\Delta m_{31}^2$. Therefore we confirm that both the cases A$_{1}$ and A$_{2}$ are consistent with the current experimental bounds\cite{Meloni:2012sx,Grimus:2012zm}. \begin{figure}[h] \begin{center} \includegraphics[width=87mm]{1112.eps} \hspace{3mm} \includegraphics[width=87mm]{1113.eps} \vspace{-1cm} \caption{Correlations between $(M_{\nu})_{23}/(M_{\nu})_{13}$ and $(M_{\nu})_{33}/(M_{\nu})_{22}$ (left panel) for the case of A$_1$, and $(M_{\nu})_{23}/(M_{\nu})_{12}$ and $(M_{\nu})_{33}/(M_{\nu})_{22}$ (right panel) for the case of A$_2$. The symbols $+$, grey points and $\ast$ represent $1 \leq |(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|<2$, $2 \leq |(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|<3$ and $3 \leq |(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|\leq 4$ for the left (right) panel, respectively.} \label{fig2} \end{center} \end{figure} Next we discuss correlations between non-zero elements of $M_{\nu}$. Figure \ref{fig2} shows the allowed region in $(M_{\nu})_{23}/(M_{\nu})_{13}- (M_{\nu})_{33}/(M_{\nu})_{22}$ plane (left panel) for the case of A$_1$, and in $(M_{\nu})_{23}/(M_{\nu})_{12}-(M_{\nu})_{33}/(M_{\nu})_{22}$ plane (right panel) for the case of A$_2$. In both panels, we have chosen the non-zero elements such that all experimental constraints Eq. (\ref{exp}) are satisfied. The symbols $+$, grey points and $\ast$ represent $1 \leq |(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|<2$, $2 \leq |(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|<3$ and $3 \leq |(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|\leq4$, respectively for the case A$_1$(A$_2$). There are no solutions for $|(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|<1$ and $4<|(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|$. From these figures, one finds solutions at $|(M_{\nu})_{23}/(M_{\nu})_{13(12)}|\sim 2~{\rm or}~3$, and $0.5 \raise0.3ex\hbox{$\;<$\kern-0.75em\raise-1.1ex\hbox{$\sim\;$}}(M_{\nu})_{33}/(M_{\nu})_{22}\lsim2$ with $1<|(M_{\nu})_{22}/ (M_{\nu})_{13(12)}|<4$. \begin{table}[h] \centering {\fontsize{9}{12} \begin{tabular}{|c||c|c|c|c|} \hline $M_{\nu}$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&3\\ \end{array}\right)$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&\sqrt{3}&2\sqrt{2}\\ 1&2\sqrt{2}&2\sqrt{2}\\ \end{array}\right)$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&\sqrt{3}&2\sqrt{2}\\ 1&2\sqrt{2}&3\\ \end{array}\right)$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&2\sqrt{2}\\ \end{array}\right)$ \\\hline $U_{PMNS}$ & \!\!$\left(\begin{array}{ccc} \!\!\!\!\frac{3+\sqrt{7}}{2\sqrt{7+2\sqrt{7}}}&\!\!\!\!\frac{1}{\sqrt{3}}&\!\!\!\!\frac{3-\sqrt{7}}{2\sqrt{7-2\sqrt{7}}}\\ \!\!\!\!-\frac{1+\sqrt{7}}{2\sqrt{7+2\sqrt{7}}}&\!\!\!\!\frac{1}{\sqrt{3}}&\!\!\!\!-\frac{1-\sqrt{7}}{2\sqrt{7-2\sqrt{7}}}\\ \!\!\!\!\frac{1}{\sqrt{7+2\sqrt{7}}}&\!\!\!\!-\frac{1}{\sqrt{3}}&\!\!\!\!\frac{1}{\sqrt{7-2\sqrt{7}}} \!\!\!\!\\ \end{array}\right)\!\!$ & \!\!$\left(\begin{array}{ccc} \!\!\!\!0.815&\!\!\!\!0.560&\!\!\!\!0.147\\ \!\!\!\!-0.518&\!\!\!\!0.592&\!\!\!\!0.617\\ \!\!\!\!0.259&\!\!\!\!-0.579&\!\!\!\!0.773\!\!\!\\ \end{array}\right)$\!\!& \!\!$\left(\begin{array}{ccc} \!\!\!\!0.803&\!\!\!\!0.578&\!\!\!\!0.145\\ \!\!\!\!-0.534&\!\!\!\!0.589&\!\!\!\!0.606\\ \!\!\!\!0.265&\!\!\!\!-0.565&\!\!\!\!0.782\!\!\!\\ \end{array}\right)$\!\! & \!\!$\left(\begin{array}{ccc} \!\!\!\!0.819&\!\!\!\!0.559&\!\!\!\!0.136\\ \!\!\!\!-0.504&\!\!\!\!0.581&\!\!\!\!0.640\\ \!\!\!\!0.279&\!\!\!\!0.592&\!\!\!\!0.756\\ \end{array}\right)$\!\! \\\hline $\sin^2 \theta_{12}$ &$\frac{2}{87}(28-5\sqrt{7})=0.3396$ & $ 0.321$&$0.341$ & $0.319$ \\\hline $\sin^2 \theta_{23}$ & $\frac{1}{29}(17-2\sqrt{7})=0.4037$&$0.389$& $0.376$ & $0.417$ \\\hline $\sin^2 \theta_{13}$ &$\frac{1}{3}-\frac{5}{6\sqrt{7}}=0.01836$&$0.0215$ & $0.0211$ &$0.0186$ \\\hline $m_1$ & $ (3-\sqrt{7})m$&$0.318m$& $0.330m$ & $0.341m$ \\\hline $m_2$ &$-m$ &$-1.03m$& $-0.977m$ & $-1.06m$ \\\hline $m_3$ & $(3+\sqrt{7})m$& $5.28m$&$5.38m$ & $5.55m$ \\\hline $\Delta m_{21}^2$ & $(-15+6\sqrt{7})m^2$& $0.966m^2$&$0.846m^2$ & $1.00 m^2$ \\\hline $\Delta m_{31}^2$ & $12\sqrt{7}m^2$&$27.7m^2$& $28.8m^2$ & $30.6m^2$ \\\hline \!\!\!\!\!\!$\begin{array}{c}|m|\vspace{-2mm}\\ ( 10^{-3}{\footnotesize \mbox{eV}})\\ \end{array}$\! \!\!\!\!\!\!\!&$ [8.95,~9.21]$ &$[9.06,~9.15]$&$[9.10,~9.67]$& $[8.62,~8.97]$\\\hline \end{tabular}% } \caption{Examples of one-parameter neutrino mass matrix $M_{\nu}$ and related quantities for $|(M_{\nu})_{23}/(M_{\nu})_{13}|\simeq 3$ for the case of A$_1$. } \label{tab1} \end{table} \begin{table}[h] \centering {\fontsize{9}{12} \begin{tabular}{|c||c|c|c|c|} \hline $M_{\nu}$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&\sqrt{6}&2\\ 1&2&5/2\\ \end{array}\right)$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&5/2&2\\ 1&2&5/2\\ \end{array}\right)$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&2\sqrt{2}&2\\ 1&2&\sqrt{5}\\ \end{array}\right)$ & $m\left(\begin{array}{ccc} 0&0&1\\ 0&2\sqrt{2}&\sqrt{5}\\ 1&\sqrt{5}&2\sqrt{2}\\ \end{array}\right)$ \\\hline $U_{PMNS}$ & \!\!$\left(\begin{array}{ccc} \!\!\!\!0.825&\!\!\!\!0.543&\!\!\!\!0.157\\ \!\!\!\!0.312&\!\!\!\!-0.669&\!\!\!\!0.675\\ \!\!\!\!-0.472&\!\!\!\!0.507&\!\!\!\!0.721\\ \end{array}\right)$\!\! & \!\!$\left(\begin{array}{ccc} \!\!\!\!0.828&\!\!\!\!0.539&\!\!\!\!0.156\\ \!\!\!\!0.306&\!\!\!\!-0.667&\!\!\!\!0.679\\ \!\!\!\!-0.470&\!\!\!\!0.515&\!\!\!\!0.717\\ \end{array}\right)$\!\!& \!\!$\left(\begin{array}{ccc} \!\!\!\!0.822&\!\!\!\!0.552&\!\!\!\!0.143\\ \!\!\!\!0.287&\!\!\!\!-0.618&\!\!\!\!0.732\\ \!\!\!\!-0.492&\!\!\!\!0.560&\!\!\!\!0.666\\ \end{array}\right)$ \!\! & \!\!$\left(\begin{array}{ccc} \!\!\!\!0.842&\!\!\!\!0.521&\!\!\!\!0.139\\ \!\!\!\!0.299&\!\!\!\!-0.664&\!\!\!\!0.685\\ \!\!\!\!-0.449&\!\!\!\!0.535&\!\!\!\!0.715\\ \end{array}\right)$\!\! \\\hline $\sin^2 \theta_{12}$ &$0.303$ & $ 0.297$&$0.311$ & $0.277$ \\\hline $\sin^2 \theta_{23}$ & $0.467$&$0.473$& $0.547$ & $0.478$ \\\hline $\sin^2 \theta_{13}$ &$0.0247$&$0.0242$ & $0.0205$ &$0.0192$ \\\hline $m_1$ & $ -0.572m$&$-0.567m$& $-0.599m$ & $-0.533m$ \\\hline $m_2$ &$0.933m$ &$0.956m$& $1.02m$ & $1.03m$ \\\hline $m_3$ & $4.59m$& $4.61m$&$4.65m$ & $5.16m$ \\\hline $\Delta m_{21}^2$ & $0.544m^2$& $0.592m^2$&$0.671m^2$ & $0.70 m^2$ \\\hline $\Delta m_{31}^2$ & $20.7m^2$&$20.9m^2$& $21.3m^2$ & $26.4m^2$ \\\hline \!\!\!\!\!\!$\begin{array}{c}|m|\vspace{-2mm}\\ ( 10^{-3}{\footnotesize \mbox{eV}})\\ \end{array}$\! \!\!\!\!\!\!\! &$ [11.3,~11.4]$ &$[10.9,~11.3]$&$[10.3,~11.0]$& $[9.54,~10.1]$\\\hline \end{tabular}% } \caption{Examples of one-parameter neutrino mass matrix $M_{\nu}$ and related quantities for $|(M_{\nu})_{23}/(M_{\nu})_{13}|\simeq 2$ for the case of A$_1$. } \label{tab1-1} \end{table} From the numerical calculation above, we find some examples of $M_{\nu}$ with one real free parameter $m$ as listed in Table \ref{tab1} for $|(M_{\nu})_{23}/(M_{\nu})_{13}|\simeq 3$ and in Table \ref{tab1-1} for $|(M_{\nu})_{23}/(M_{\nu})_{13}|\simeq 2$ for the case A$_1$ \footnote{See Ref.\cite{Araki:2012ip} for another example of the one-parameter texture $M_{\nu}=m\left( \begin{array}{ccc} 0&0&1\\0&3&2\\1&2&2\\ \end{array}\right)$.}. The bound of $|m|$ is obtained by the overlap of two constraints of $\Delta m_{21}^2$ and $\Delta m_{31}^2$ given in Eq.(\ref{exp}). One can see that all textures are in agreement with the current experimental bounds at the 3 $\sigma$ level given in Eq.(\ref{exp}). In addition to the mass matrices shown in Tables \ref{tab1} and \ref{tab1-1}, matrices with different sign are also candidates of one-parameter $M_{\nu}$, such as \begin{eqnarray} \left( \begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&3\\ \end{array}\right)\to \left( \begin{array}{ccc} 0&0&1\\ 0&2&-3\\ 1&-3&3\\ \end{array}\right),~ \left( \begin{array}{ccc} 0&0&1\\ 0&-2&3\\ 1&3&-3\\ \end{array}\right),~ \left( \begin{array}{ccc} 0&0&1\\ 0&-2&-3\\ 1&-3&-3\\ \end{array}\right), \label{sign} \end{eqnarray} with the same observables $\sin^2 \theta_{12,23,13}$, $\Delta m^2_{21,31}$ and consequently the same bounds of $|m|$, while the sign of mass eigenvalues $m_{1,2,3}$ and that of the PMNS matrix $U_{PMNS}$ are different. For the case A$_2$, all mass matrices $M_{\nu}({\rm A}_2)$ obtained from $M_{\nu}({\rm A}_1)$ listed in Tables \ref{tab1}, \ref{tab1-1} and Eq.(\ref{mnu0}), are also consistent with Eq.(\ref{exp}). The PMNS matrix for the case A$_2$ is given from that for the case A$_1$ by $U_{PMNS}({\rm A}_2)=PU_{PMNS}({\rm A}_1)$, {\it i.e.}, $\sin^2 \theta_{23}({\rm A}_2)=1-\sin^2 \theta_{23}({\rm A}_1) $ and the other quantities remain unchanged. In the next section, we discuss symmetry realization of the mass matrix \begin{equation} M_{\nu}=m\left( \begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&3\\ \end{array}\right), \label{mnu} \end{equation} by the binary icosahedral symmetry $A_5'$, as an example. \section{Symmetry Realization} In this section, we discuss symmetry realization of the one-parameter texture Eq.(\ref{mnu}) by the binary icosahedral symmetry $A_5'$ and the Higgs potential of our model. \subsection{Mass Matrices} \begin{table}[thbp] \centering {\fontsize{10}{12} \begin{tabular}{||c|c|c||c|c|c|c|c||} \hline\hline ~~~~ & ~~ $L_a$~~ & ~~$e^c_a$~~ & ~~ $\Phi_0$~~ & $\Phi_a$ & $\Phi'_A$ &$\Delta_A$ \\\hline $(SU(2)_L,U(1)_Y)$ & $({\bf{2}},-1/2)$ & $({\bf{1}},1)$ & $({\bf{2}},1/2)$ & $({\bf{2}}, 1/2)$ & $({\bf{2}},1/2)$ & $({\bf{3}},1)$ \\\hline $A_5'$ & ${\bf 3}_a$ & ${\bf 3}_a$ & ${\bf 1}$ & ${\bf 3}_a$ & ${\bf 5}_A$& ${\bf 5} _A$\\\hline \end{tabular}% } \caption{The particle contents ($a=1-3$ and $A=1-5$). } \label{tab2} \end{table} Now we discuss symmetry realization of the texture given in Eq.(\ref{mnu}) for neutrino mass matrix. We consider the binary icosahedral symmetry $A_5'$ \cite {Hashimoto:2011tn} as a flavor symmetry. The $A_5'$ symmetry contains three- and five-dimensional irreducible representations ${\bf 3}$ and ${\bf 5}$, and their tensor product gives $A_5'$ invariance ${\bf 1}$. Moreover the resultant singlet ${\bf 1}$ from ${\bf 3 \cdot 5 \cdot 3}$ enters all the elements of $M_{\nu}$ with desired weights if neutrinos and Higgs bosons are embedded into ${\bf 3}$ and ${\bf 5}$, respectively. Therefore the $A_5'$ symmetry is preferable for the one-parameter texture. For the lepton sector, the particle contents are shown in Table \ref{tab2}, where $L_a,~e^c_a,~(\Phi_{0,a},~\Phi'_A)$ and $\Delta_A$ ($a=1-3,~A=1-5$) are the $SU(2)_L$ doublet leptons, $SU(2)$ singlet leptons, $SU(2)$ doublet Higgs fields and $SU(2)$ triplet Higgs fields, respectively. Since right-handed neutrinos are absent in our model, the $SU(2)_L$ doublet Higgs fields are responsible only for masses of charged leptons, while neutrino masses are generated by the vacuum expectation values (VEVs) of $SU(2)_L$ triplet Higgs fields through the type-II seesaw mechanism. If one assigns ${\bf 1}$ for quarks, only $\Phi_0$ couples to quarks and gives their masses. Although there are no predictions in the quark sector, it is ensured to be the same as the standard model. Now we give the multiplication rules of the $A_5'$ group relevant for the Yukawa interactions in our model. For ${\bf 3}$ and ${\bf 5}$ irreducible representations \begin{eqnarray} &&{\bf 3}=\left( \begin{array}{c} x_1\\x_2\\x_3\\ \end{array}\right),~ \left( \begin{array}{c} y_1\\y_2\\y_3\\ \end{array}\right),~ {\bf 5}=\left( \begin{array}{c} X_1\\X_2\\X_3\\X_4\\X_5\\ \end{array}\right),~ \left( \begin{array}{c} Y_1\\Y_2\\Y_3\\Y_4\\Y_5\\ \end{array}\right), \end{eqnarray} their tensor products are given by \footnote{See Ref.\cite{Hashimoto:2011tn} for the complete multiplication rules.} \begin{eqnarray} && {\bf 3}\otimes {\bf 3} =\left( x_1 y_3-x_2 y_2+x_3 y_1\right)_{{\bf 1}} +\frac{1}{\sqrt{2}}\left( \begin{array}{c} x_1 y_2-x_2y_1\\ x_1y_3-x_3y_1\\ x_2y_3-x_3y_2 \end{array}\right)_{{\bf 3}} +\left(\begin{array}{c} x_1 y_1\\ \frac{1}{\sqrt{2}}\left(x_1y_2+x_2y_1 \right)\\ \frac{1}{\sqrt{6}}\left( x_1y_3+2x_2y_2+x_3y_1\right)\\ \frac{1}{\sqrt{2}}\left( x_2y_3+x_3y_2\right)\\ x_3y_3\\ \end{array}\right)_{{\bf 5}},\nonumber\\&& {\bf 5}\otimes {\bf 5} =\left( X_1Y_5-X_2Y_4+X_3Y_3-X_4Y_2+X_5Y_1\right)_{\bf 1}+\cdots. \label{rule} \end{eqnarray} From the particle contents shown in Table \ref{tab2} and the multiplication rules Eq. (\ref{rule}), the $A_5'$ invariant Yukawa interactions are given by \begin{eqnarray} {\cal L}_Y&=&y_{\Delta}L_a^T i \sigma_2 \Delta_A L_b +y_1 \Phi_0^{\dag}L_a e_b^c +y_2 \Phi_a^{\dag}L_b e_c^c +y_3 {\Phi'}_A^{\dag}L_a e_b^c+c.c., \label{yukawa} \end{eqnarray} where all indices are summed up in $A_5'$ invariant way in accordance with Eq(\ref{rule}). After the electroweak symmetry breaking by the VEVs of the Higgs fields defined by \begin{equation} \langle \Delta_A \rangle=\frac{1}{\sqrt{2}}v_{\Delta A},~~ \langle \Phi_{0,a} \rangle=\frac{1}{\sqrt{2}}v_{0,a},~~ \langle \Phi'_{A} \rangle=\frac{1}{\sqrt{2}}V_{A}, \end{equation} with $v_{0}^2+\sum_a v_a^2+\sum_A(V_A^2+2v_{\Delta A}^2)=(246\mbox{GeV})^2$ , we obtain the following mass matrix \begin{equation} M_{\nu}=\frac{y_{\Delta}}{\sqrt{2}}\left( \begin{array}{ccc} v_{\Delta 5}&-\frac{1}{\sqrt{2}}v_{\Delta 4}&\frac{1}{\sqrt{6}}v_{\Delta 3}\\ -\frac{1}{\sqrt{2}}v_{\Delta 4}&\frac{2}{\sqrt{6}}v_{\Delta 3}& -\frac{1}{\sqrt{2}}v_{\Delta 2}\\ \frac{1}{\sqrt{6}}v_{\Delta 3}&-\frac{1}{\sqrt{2}}v_{\Delta 2}& v_{\Delta 1}\\ \end{array}\right), \label{mnu2} \end{equation} for neutrino sector, and \begin{equation} M_e=\frac{1}{\sqrt{2}}\left( \begin{array}{ccc} y_3 V_5&\frac{1}{\sqrt{2}}y_2 v_3-\frac{1}{\sqrt{2}}y_3V_4&y_1 v_0 -\frac{1}{\sqrt{2}}y_2v_2+\frac{1}{\sqrt{6}}y_3 V_3\\ -\frac{1}{\sqrt{2}}y_2 v_3-\frac{1}{\sqrt{2}}y_3V_4& -y_1 v_0+\frac{2}{{\sqrt{6}}}y_3V_3&\frac{1}{\sqrt{2}}y_2 v_1-\frac{1}{\sqrt{2}}y_3 V_2\\ y_1v_0+\frac{1}{\sqrt{2}}y_2 v_2+\frac{1}{\sqrt{6}}y_3V_3& -\frac{1}{\sqrt{2}}y_2 v_1-\frac{1}{\sqrt{2}}y_3 V_2&y_3 V_1\\ \end{array}\right), \label{me2} \end{equation} for charged lepton sector. If the Higgs fields obtain the following VEVs, \begin{eqnarray} &&v_{\Delta 1}=3 v_{\Delta},~v_{\Delta 2}=-3\sqrt{2} v_{\Delta},~ v_{\Delta 3}=\sqrt{6} v_{\Delta},~v_{\Delta 4}=v_{\Delta 5}=0,\nonumber\\&& v_0\neq 0,~v_2\neq 0,~V_3\neq 0,~v_1=v_3=V_1=V_2=V_4=V_5=0,~ \label{vev} \end{eqnarray} one finds that the mass matrices $M_{\nu}$ and $M_e$ have the desired form \begin{equation} M_{\nu}=\frac{y_{\Delta}v_{\Delta}}{\sqrt{2}} \left( \begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&3\\ \end{array}\right),~~~ M_e=\left(\begin{array}{ccc} 0&0&m_e\\ 0&-m_{\mu}&0\\ m_{\tau}&0&0\\ \end{array} \right), \label{mnume} \end{equation} where \begin{eqnarray} && m_e=\frac{1}{\sqrt{2}}\left( y_1 v_0-\frac{1}{\sqrt{2}}y_2v_2+\frac{1}{\sqrt{6}}y_3V_3\right),\\&& m_{\mu}=\frac{1}{\sqrt{2}}\left(y_1v_0-\frac{2}{\sqrt{6}}y_3V_3 \right),\\&& m_{\tau}=\frac{1}{\sqrt{2}}\left( y_1 v_0+\frac{1}{\sqrt{2}}y_2v_2+\frac{1}{\sqrt{6}}y_3V_3\right), \end{eqnarray} and $M_eM_e^{\dag}={\rm diag}(m_e^2,~m_{\mu}^2,~m_{\tau}^2)$. The VEVs of the triplet Higgs $\Delta_A$ gives additional contributions to the $\rho$ parameter \begin{equation} \rho=\frac{v^2} {v^2+2\sum_{A} v_{\Delta A}^2} = \frac{v^2}{v^2+66 v_{\Delta}^2}, \end{equation} where $v^2=v_0^2+\sum_a v_a^2+\sum_A V_A^2=v_0^2+v_2^2+V_3^2$ because of Eq. (\ref{vev}). The experimental value $\rho_{\rm exp}=1.0004^{+0.0003}_{-0.0004}$ constrains $v_{\Delta}$ to be smaller than about $0.61\mbox{GeV}$ at the 95$\%$ confidence level. Therefore we assume $v_{\Delta}\ll v_0,v_2,V_3$. As already discussed in the last section, the mass matrices Eq.(\ref{mnume}) with $m=y_{\Delta}v_{\Delta}/\sqrt{2}$ corresponding to Eq.(\ref{mnu}) are compatible with the current experimental bounds. \subsection{Higgs Sector} As for the Higgs sector, the total Higgs potential is given in Appendix in symbolic form. Here we mention the tadpole conditions. The $A'_5$ invariant scalar mass terms are given by \begin{eqnarray} &&V_{\rm mass}=m_0^2 \Phi_0^{\dag}\Phi_0 +m_3^2\left( \Phi_1^{\dag}\Phi_3-\Phi_2^{\dag}\Phi_2+\Phi_3^{\dag}\Phi_1 \right) \nonumber\\&& +m_5^2 \left( {\Phi'}_1^{\dag}\Phi'_5- {\Phi'}_2^{\dag}\Phi'_4+ {\Phi'}_3^{\dag}\Phi'_3- {\Phi'}_4^{\dag}\Phi'_2+ {\Phi'}_5^{\dag}\Phi'_1\right)\nonumber\\&& +m_{\Delta}^2{\rm Tr}\left[ \Delta_1^{\dag}\Delta_5- \Delta_2^{\dag}\Delta_4 + \Delta_3^{\dag}\Delta_3- \Delta_4^{\dag}\Delta_2+ \Delta_5^{\dag}\Delta_1\right]. \label{scalarmass} \end{eqnarray} Since the total $A'_5$ invariant potential does not have any accidental symmetry, our model does not suffer from the problem of massless Goldstone bosons. By imposing the conditions for the VEVs Eq.(\ref{vev}), the tadpole conditions give the scalar masses \begin{eqnarray} m_0^2&\simeq&-\lambda_1 v_0^2+\left( \frac 12 \tilde \epsilon_1 -\frac{V_3}{\sqrt{6}v_0}\tilde \kappa_8\right)v_2^2 -\frac 12\left( \tilde \epsilon_2-\frac{V_3}{v_0}\kappa_2^{(1)}\right)V_3^2, \label{higgsm1}\\ m_{3}^2 &\simeq& -\frac 12 \tilde \epsilon_1 v_0^2 +\left( \lambda_2^{(1)}+\frac 23 \lambda_2^{(3)}\right)v_2^2 -\left( \tilde \epsilon_4+2 \tilde \epsilon_5 +\tilde \epsilon_6\right)V_3^2 +\frac{\sqrt{6}}{3}\tilde \kappa_8 v_0V_3 -2\sqrt{2}v_{\Delta}\mu_1,\label{higgsm2}\\ m_{5}^2&\simeq& -\frac 12 \tilde \epsilon_2 v_0^2 +\left( \tilde \epsilon_4+2 \tilde \epsilon_5 +\tilde \epsilon_6 -\frac{v_0}{\sqrt{6}V_3}\tilde \kappa_8\right)v_2^2 -\tilde \lambda_3 V_3^2 +\frac 32 \kappa_2^{(1)}v_0 V_3 -2\sqrt{3}v_{\Delta}\left( \mu_2-6 \mu_3\right),\label{higgsm3}\\ m_{\Delta}^2&\simeq& -\frac 12 \tilde \epsilon_3 v_0^2 +\left( \tilde \epsilon_{7}+ \tilde \epsilon_{8}+\frac{\mu_1}{3\sqrt{2}v_{\Delta}}\right)v_2^2 -\frac 12\left( \tilde \lambda_4+\tilde \lambda_5 +\frac{\mu_2-6\mu_3}{\sqrt{3}v_{\Delta}}\right)V_3^2 +\tilde \kappa_2 v_0 V_3, \label{higgsm4} \end{eqnarray} where \begin{eqnarray} \tilde \lambda_i&=&\lambda_i^{(1)}-\lambda_i^{(7)}-6\lambda_i^{(8)}+36\lambda_i^{(9)}~ (i=3,4,5),\nonumber\\ \tilde \epsilon_{i}&=&\epsilon_{i}^{(1)}+\epsilon_{i}^{(2)} +2 \epsilon_{i}^{(3)}~(i=1,2),~ \tilde \epsilon_3=\epsilon_3^{(1)}+\epsilon_3^{(2)},~ \tilde \epsilon_{i}=\frac 12 \epsilon_i^{(1)}+\frac{1}{\sqrt{6}}\epsilon_i^{(3)} +\frac{3}{10}\epsilon_i^{(4)}~(i=4,5,6,7,8),\nonumber\\ \tilde \kappa_2&=&\kappa_2^{(2)}+\kappa_2^{(3)},~ \tilde \kappa_8=\kappa_8^{(1)}+\kappa_8^{(2)}+\kappa_8^{(3)}. \label{lambda} \end{eqnarray} The coupling constants $\lambda$s, $\epsilon$s and $\kappa$s are given in Appendix. In Eqs. (\ref{higgsm1})-(\ref{higgsm4}), the terms proportional to $v_{\Delta}^2$ have been neglected. The pattern \begin{equation} M_{\nu}=\left( \begin{array}{ccc} 0&0&1\\ 0&2&3\\ 1&3&2\sqrt{2}\\ \end{array}\right), \end{equation} listed in Table \ref{tab1} can be realized by the $A_5'$ symmetry in similar fashion, with corresponding tadpole conditions. \section{Conclusions} We have studied neutrino mass matrix $M_{\nu}$ with two texture zeros and its symmetry realization. After confirming that the cases A$_{1}$ and A$_2$ satisfy the current experimental constraints by numerical calculation, we have found some examples of $M_{\nu}$ with {\it one} real parameter. Since the magnitude of non-zero elements of $M_{\nu}$ is restricted in small region, one can extract examples of $M_{\nu}$ with simple forms. While there exist infinite number of candidates of one-parameter $M_{\nu}$, such simple forms are preferable in the standpoint of flavor symmetry. Next we have discussed symmetry realization of one-parameter $M_{\nu}$ and the Higgs potential based on the binary icosahedral symmetry $A_5'$. The $A_5'$ symmetry contains three- and five-dimensional irreducible representations, and their tensor product ${\bf 3 \cdot 5 \cdot 3}$ enters all the elements of $M_{\nu}$ with definite weights. If one assigns ${\bf 5}$ to $SU(2)_L$ triplet Higgs $\Delta$, desired neutrino mass matrix is obtained by the type-II seesaw mechanism and by choosing the vacua of the Higgs potential. While we have shown one example of symmetry realization in this paper, the $A_5'$ symmetry can work for the other one-parameter $M_{\nu}$ because of its multiplication rules. \section*{Acknowledgments} We would like to thank H. Okada and A. Shibuya for useful discussions. \begin{appendix} \section{Higgs Potential} Here we show the Higgs potential of our model with the A$_5'$ symmetry. In order to avoid redundancy, we give a symbolic form of the potential. The Higgs fields are denoted by \begin{equation} \Phi_0\equiv {\bf 1},~\Phi_a\equiv {\bf 3},~\Phi'_A\equiv {\bf 5},~ \Delta_A \equiv \Delta \end{equation} where $a=1-3$ and $A=1-5$ for three- and five- dimensional representation, respectively. Their Hermitian conjugate fields are denoted by $\Phi_a^{\dag} \equiv {\bf 3}^{\dag}$ etc., while ${\bf 3}^{\dag}$ and ${\bf 3}$ obey the same multiplication rules. We represent a product \begin{eqnarray} {\bf 3}\otimes {\bf 3} ={\bf 1}\oplus {\bf 3} \oplus {\bf 5}~&\rightarrow&~ {\bf 33}=({\bf 33})_{\bf 1}+({\bf 33})_{\bf 3}+({\bf 33})_{\bf 5} =({\bf 33})_{\alpha},\nonumber\\ &{\rm or}&~{\bf 3^{\dag}3}=({\bf 3^{\dag}3})_{\bf 1}+({\bf 3^{\dag}3})_{\bf 3} +({\bf 3^{\dag}3})_{\bf 5}=({\bf 3^{\dag}3})_{\alpha}, \end{eqnarray} depending on the gauge quantum number of the Higgs fields. The index $\alpha$ must be summed up for all possible combinations of the tensor product. For example in the case of $({\bf 3^{\dag} 3})^2$, since only the products ${\bf 11},~{\bf 33}$ and ${\bf 55}$ can be invariant under A$_5'$, we denote $\lambda_2^{(\alpha)}({\bf 3}^{\dag}{\bf 3})_{\alpha}({\bf 3}^{\dag}{\bf 3})_{ \alpha}$ as \begin{eqnarray} &&({\bf 3^{\dag} 3})^2~:~ \lambda_2^{(\alpha)}({\bf 3}^{\dag}{\bf 3})_{\alpha}({\bf 3}^{\dag}{\bf 3})_{ \alpha} =\lambda_2^{(1)}({\bf 3}^{\dag}{\bf 3})_{\bf 1}({\bf 3}^{\dag}{\bf 3})_{ \bf 1} +\lambda_2^{(2)}({\bf 3}^{\dag}{\bf 3})_{\bf 3}({\bf 3}^{\dag}{\bf 3})_{ \bf 3} +\lambda_2^{(3)}({\bf 3}^{\dag}{\bf 3})_{\bf 5}({\bf 3}^{\dag}{\bf 3})_{ \bf 5}\nonumber\\&& =\lambda_2^{(1)}\left( \Phi_1^{\dag}\Phi_3-\Phi_2^{\dag}\Phi_2+\Phi_3^{\dag}\Phi_1\right)^2\nonumber\\&& +\lambda_2^{(2)}\left[2\left(\Phi_1^{\dag}\Phi_2-\Phi_2^{\dag}\Phi_1 \right) \left(\Phi_2^{\dag}\Phi_3-\Phi_3^{\dag}\Phi_2 \right) -\left(\Phi_1^{\dag}\Phi_3-\Phi_3^{\dag}\Phi_1 \right)^2 \right]\nonumber\\&& +\lambda_2^{(3)}\left[ 2\left(\Phi_1^{\dag}\Phi_1\right)\left(\Phi_3^{\dag}\Phi_3\right) -\left(\Phi_1^{\dag}\Phi_2+\Phi_2^{\dag}\Phi_1 \right) \left(\Phi_2^{\dag}\Phi_3+\Phi_3^{\dag}\Phi_2 \right) +\frac{1}{6}\left( \Phi_1^{\dag}\Phi_3+\Phi_2^{\dag}\Phi_2+\Phi_3^{\dag}\Phi_1\right)^2 \right].\nonumber\\ && \end{eqnarray} Moreover in what follows, trace for $\Delta$ is omitted, and ``~$\cdot$~'' denotes the Pauli matrices $\sigma^{1,2,3}$. In the notation described above, the Higgs potential except the bilinear terms given in the main text can be written down as \begin{eqnarray} {\bf 3}^2{\bf 5},{\bf 5}^3,{\bf 3}{\bf 5}^2&:& \mu_1{\bf 3} \Delta^{\dag} {\bf 3} +\mu_2({\bf 5} \Delta^{\dag} )_{\bf 5_1}{\bf 5} +\mu_3({\bf 5} \Delta^{\dag})_{\bf 5_2} {\bf 5} +\mu_4{\bf 3} \Delta^{\dag} {\bf 5}+h.c., \end{eqnarray} for the trilinear terms, and \begin{eqnarray} {\bf 1}^4&:&\lambda_1 ({\bf 1}^{\dag} {\bf 1})^2,\nonumber \\ {\bf 3}^4&:&\lambda_2^{(\alpha)}({\bf 3}^{\dag}{\bf 3})_{\alpha} ({\bf 3}^{\dag}{\bf 3})_{ \alpha},\nonumber\\ {\bf 5}^4&:& \lambda_3^{(\alpha)}({\bf 5}^{\dag}{\bf 5})_{\alpha}({\bf 5}^{\dag}{\bf 5})_{\alpha} +\lambda_4^{(\alpha)}({\bf 5}^{\dag}{\bf 5})_{\alpha}({\Delta}^{\dag}{\Delta})_{\alpha} +\lambda_5^{(\alpha)}({\bf 5}^{\dag}\cdot{\bf 5})_{\alpha}({\Delta}^{\dag} \cdot{\Delta})_{\alpha}\nonumber\\ &&+\lambda_6^{(\alpha)}((\Delta^{\dag}\Delta)_{\alpha})^2 +\lambda_7\det\left[ \Delta^{\dag}\Delta\right] ,\nonumber\\ {\bf 1}^2{\bf 3}^2&:& \epsilon_1^{(1)}({\bf 1}^{\dag}{\bf 1})_{{\bf 1}}({\bf 3}^{\dag}{\bf 3})_{{\bf 1}} +\epsilon_1^{(2)}({\bf 1}^{\dag}{\bf 3})_{{\bf 3}}({\bf 3}^{\dag}{\bf 1})_{{\bf 3}} +\epsilon_1^{(3)}\left[( ({\bf 1}^{\dag}{\bf 3})_{{\bf 3}})^2+h.c.\right],\nonumber\\ {\bf 1}^2{\bf 5}^2&:& \epsilon_2^{(1)}({\bf 1}^{\dag}{\bf 1})_{{\bf 1}}({\bf 5}^{\dag}{\bf 5})_{{\bf 1}} +\epsilon_2^{(2)}({\bf 1}^{\dag}{\bf 5})_{{\bf 5}}({\bf 5}^{\dag}{\bf 1})_{{\bf 5}} +\epsilon_2^{(3)}\left[( ({\bf 1}^{\dag}{\bf 5})_{{\bf 5}})^2+h.c.\right]\nonumber\\ &&+\epsilon_3^{(1)}({\bf 1}^{\dag}{\bf 1})_{{\bf 1}}(\Delta^{\dag}\Delta)_{\bf 1} +\epsilon_3^{(2)}({\bf 1}^{\dag}\cdot{\bf 1})_{{\bf 1}}(\Delta^{\dag}\cdot\Delta)_{\bf 1}, \nonumber\\ {\bf 3}^2{\bf 5}^2&:& \epsilon_4^{(\alpha)}({\bf 3}^{\dag}{\bf 3})_{\alpha}({\bf 5}^{\dag}{\bf 5})_{\alpha} +\epsilon_5^{(\alpha)}\left[(({\bf 3}^{\dag}{\bf 5})_{\alpha})^2+h.c.\right] +\epsilon_6^{(\alpha)}({\bf 3}^{\dag}{\bf 5})_{\alpha}({\bf 5}^{\dag}{\bf 3})_{\alpha}\nonumber\\ &&+\epsilon_{7}^{(\alpha)}({\bf 3}^{\dag}{\bf 3})_{\alpha}(\Delta^{\dag}\Delta)_{\alpha} +\epsilon_{8}^{(\alpha)}({\bf 3}^{\dag}\cdot{\bf 3})_{\alpha}(\Delta^{\dag}\cdot \Delta)_{\alpha}\nonumber\\ {\bf 1}{\bf 3}^3&:&\kappa_1 ({\bf 1}^{\dag}{\bf 3})_{\bf 3}({\bf 3}^{\dag}{\bf 3})_{\bf 3} +h.c.,\nonumber\\ {\bf 1}{\bf 5}^3&:& \kappa_2^{(1)}({\bf 1}^{\dag}{\bf 5})_{\bf 5}({\bf 5}^{\dag}{\bf 5})_{\bf 5} +\kappa_2^{(2)}({\bf 1}^{\dag}{\bf 5})_{\bf 5}(\Delta^{\dag}\Delta)_{\bf 5} +\kappa_2^{(3)}({\bf 1}^{\dag}\cdot{\bf 5})_{\bf 5}(\Delta^{\dag}\cdot\Delta)_{\bf 5} +h.c.,\nonumber\\ {\bf 3}^3{\bf 5}&:& \kappa_3^{(1)}({\bf 3}^{\dag}{\bf 3})_{\bf 3}({\bf 3}^{\dag}{\bf 5})_{\bf 3} +\kappa_3^{(2)}({\bf 3}^{\dag}{\bf 3})_{\bf 5}({\bf 3}^{\dag}{\bf 5})_{\bf 5}+h.c.,\nonumber\\ {\bf 3}{\bf 5}^3&:& \kappa_4^{(\alpha)}({\bf 3}^{\dag}{\bf 5})_{\alpha}({\bf 5}^{\dag}{\bf 5})_{\alpha} +\kappa_5^{(\alpha)}({\bf 3}^{\dag}{\bf 5})_{\alpha}(\Delta^{\dag}\Delta)_{\alpha} +\kappa_6^{(\alpha)}({\bf 3}^{\dag}\cdot{\bf 5})_{\alpha}(\Delta^{\dag}\cdot\Delta)_{\alpha}+h.c.,\nonumber\\ {\bf 1}{\bf 3}{\bf 5}^2&:& \kappa_7^{(1)}({\bf 1}^{\dag}{\bf 3})_{\bf 3}({\bf 5}^{\dag}{\bf 5})_{\bf 3} +\kappa_7^{(2)}({\bf 1}^{\dag}{\bf 5})_{\bf 5}({\bf 3}^{\dag}{\bf 5})_{\bf 5} +\kappa_7^{(3)}({\bf 1}^{\dag}{\bf 5})_{\bf 5}({\bf 5}^{\dag}{\bf 3})_{\bf 5}\nonumber\\ &&+\kappa_7^{(4)}({\bf 1}^{\dag}{\bf 3})_{\bf 3}(\Delta^{\dag}\Delta)_{\bf 3} +\kappa_7^{(5)}({\bf 1}^{\dag}\cdot{\bf 3})_{\bf 3}(\Delta^{\dag}\cdot\Delta)_{\bf 3} +h.c.,\nonumber\\ {\bf 1}{\bf 3}^2{\bf 5}&:& \kappa_{8}^{(1)}({\bf 1}^{\dag}{\bf 3})_{\bf 3}({\bf 3}^{\dag}{\bf 5})_{\bf 3} +\kappa_{8}^{(2)}({\bf 1}^{\dag}{\bf 3})_{\bf 3}({\bf 5}^{\dag}{\bf 3})_{\bf 3} +\kappa_{8}^{(3)}({\bf 1}^{\dag}{\bf 5})_{\bf 5}({\bf 3}^{\dag}{\bf 3})_{\bf 5} +h.c., \end{eqnarray} for the quartic terms. In the above expressions, the sums of index $\alpha$ are defined as \begin{eqnarray} &&{\bf 3}^4~:~ \lambda_2^{(\alpha)}({\bf 3}^{\dag}{\bf 3})_{\alpha}({\bf 3}^{\dag}{\bf 3})_{ \alpha} =\lambda_2^{(1)}({\bf 3}^{\dag}{\bf 3})_{\bf 1}({\bf 3}^{\dag}{\bf 3})_{ \bf 1} +\lambda_2^{(2)}({\bf 3}^{\dag}{\bf 3})_{\bf 3}({\bf 3}^{\dag}{\bf 3})_{ \bf 3} +\lambda_2^{(3)}({\bf 3}^{\dag}{\bf 3})_{\bf 5}({\bf 3}^{\dag}{\bf 3})_{ \bf 5}\nonumber\\&& {\bf 5}^4~:~ \lambda^{(\alpha)}({\bf 5}^{\dag}{\bf 5})_{\alpha}({\bf 5}^{\dag}{\bf 5})_{ \alpha} =\lambda^{(1)}({\bf 5}^{\dag}{\bf 5})_{\bf 1}({\bf 5}^{\dag}{\bf 5})_{\bf 1} +\lambda^{(2)}({\bf 5}^{\dag}{\bf 5})_{\bf 3}({\bf 5}^{\dag}{\bf 5})_{\bf 3} +\lambda^{(3)}({\bf 5}^{\dag}{\bf 5})_{\bf 3'}({\bf 5}^{\dag}{\bf 5})_{\bf 3'}\nonumber\\&& \hspace{4.3cm} +\lambda^{(4)}({\bf 5}^{\dag}{\bf 5})_{\bf 4_1}({\bf 5}^{\dag}{\bf 5})_{\bf 4_1} +\lambda^{(5)}({\bf 5}^{\dag}{\bf 5})_{\bf 4_1}({\bf 5}^{\dag}{\bf 5})_{\bf 4_2} +\lambda^{(6)}({\bf 5}^{\dag}{\bf 5})_{\bf 4_2}({\bf 5}^{\dag}{\bf 5})_{\bf 4_2}\nonumber\\&& \hspace{4.3cm} +\lambda^{(7)}({\bf 5}^{\dag}{\bf 5})_{\bf 5_1}({\bf 5}^{\dag}{\bf 5})_{\bf 5_1} +\lambda^{(8)}({\bf 5}^{\dag}{\bf 5})_{\bf 5_1}({\bf 5}^{\dag}{\bf 5})_{\bf 5_2} +\lambda^{(9)}({\bf 5}^{\dag}{\bf 5})_{\bf 5_2}({\bf 5}^{\dag}{\bf 5})_{\bf 5_2}\nonumber\\&& {\bf 3}^2{\bf 5}^2~:~ \epsilon^{(\alpha)}({\bf 3^{\dag}3})_{\alpha}({\bf 5^{\dag}5})_{\alpha} =\epsilon^{(1)}({\bf 3^{\dag}3})_{\bf 1}({\bf 5^{\dag}5})_{\bf 1} +\epsilon^{(2)}({\bf 3^{\dag}3})_{\bf 3}({\bf 5^{\dag}5})_{\bf 3} +\epsilon^{(3)}({\bf 3^{\dag}3})_{\bf 5}({\bf 5^{\dag}5})_{\bf 5}\nonumber\\&& \hspace{4.6cm} +\epsilon^{(4)}({\bf 3^{\dag}3})_{\bf 3'}({\bf 5^{\dag}5})_{\bf 3'} +\epsilon^{(5)}({\bf 3^{\dag}3})_{\bf 4}({\bf 5^{\dag}5})_{\bf 4}\nonumber\\&& {\bf 3}{\bf 5}^3~:~ \kappa^{(\alpha)}({\bf 3}^{\dag}{\bf 5})_{\alpha}({\bf 5}^{\dag}{\bf 5})_{\alpha} =\kappa^{(1)}({\bf 3}^{\dag}{\bf 5})_{\bf 3}({\bf 5}^{\dag}{\bf 5})_{\bf 3} +\kappa^{(2)}({\bf 3}^{\dag}{\bf 5})_{\bf 3'}({\bf 5}^{\dag}{\bf 5})_{\bf 3'} +\kappa^{(3)}({\bf 3}^{\dag}{\bf 5})_{\bf 4}({\bf 5}^{\dag}{\bf 5})_{\bf 4_1}\nonumber\\&& \hspace{4.6cm} +\kappa^{(4)}({\bf 3}^{\dag}{\bf 5})_{\bf 4}({\bf 5}^{\dag}{\bf 5})_{\bf 4_2} +\kappa^{(5)}({\bf 3}^{\dag}{\bf 5})_{\bf 5}({\bf 5}^{\dag}{\bf 5})_{\bf 5_1} +\kappa^{(6)}({\bf 3}^{\dag}{\bf 5})_{\bf 5}({\bf 5}^{\dag}{\bf 5})_{\bf 5_2}. \end{eqnarray} Here notice that the product ${\bf 55}$ contains two ${\bf 4}(={\bf 4_1,4_2})$ and ${\bf 5}(={\bf 5_1,5_2})$. See Ref.\cite{Hashimoto:2011tn} for details. \end{appendix}
\chapter{TEXT} with actual chapter name \iffalse \begin{enumerate}[1] \item if on an integral manifold $\alpha={\rm d} w_i$ for some locally defined functions $w_i$, then $w_i-w_j=c_{ij}$ and $c_{ij} \in \mathbb C$ is the Czech class. If $[\alpha]$ is integral, then $W= \exp(2\pi \textnormal{i} w_j)=\exp(2\pi \textnormal{i} w_k)$ is a well-defined function. If we use a spectral family on tori, then we have $W(\lambda)=\exp(2\pi \textnormal{i} w(\lambda))$ where$\lambda$ is the spectral parameter. Is this related to those holonomy functions that arise on the spectral curve? Might $W$, or $w$ in the case of spheres, be related to the meromorphic function $g$ that Fernandez has found? \item minimal surfaces in $E^3$ for which $[\omega]$ is trivial- No logarithmic terms at the ends. Delaunay surfaces, sphere with two punctures, has Hopf differential $(1/z^2)(dz)^2$ So this is split, but $[\omega]$ is nontrivial. \item minimal Lagrangian surfaces induced by a linear embedding of the Jacobian \end{enumerate} \fi \section{Introduction}\label{sec:intro} \subsection{Minimal Lagrangian surface} For an immersed Lagrangian submanifold in a K\"ahler manifold, the vanishing of the mean curvature is equivalent to that the associated section of $(n,0)$-form is parallel along the submanifold. From the well known identity in K\"ahler geometry, that the Ricci 2-form is up to constant scale the curvature form of the canonical line bundle, the minimality condition implies that the restriction of Ricci 2-form to the Lagrangian submanifold mush also vanish. When coupled with the Lagrangian condition, they form an over-determined system of differential equations which is generally not compatible. In case the ambient manifold is K\"ahler-Einstein and the Ricci 2-form is a constant multiple of the symplectic form, the minimal Lagrangian equation is involutive and, at least locally, it admits many solutions, \cite{Bryant1987Lag}. \subsubsection{Special Lagrangian submanifold} In relation to the developments in string theory, special Lagrangian submanifolds in Calabi-Yau (or Ricci-flat K\"ahler) manifolds have received much attention recently. As a particular case of calibrated geometry, the various aspects of the geometry of special Lagrangian varieties have been studied from the analytic perspectives, including deformation problem \cite{McLean1998}, gluing constructions \cite{Haskins2007,Haskins2012,Pacini2013}, and singularity analysis \cite{Joyce2003,Butscher2004}, etc. We refer to \cite{Joyce2007,Lee2011} for the further references. From a different perspective, the special Lagrangian submanifolds in $\mathbb C^m$ with nontrivial second order symmetries have been studied by in-depth analyses of the structure equations in \cite{Bryant2006,Ionel2003}. \subsubsection{Minimal Lagrangian surface} In the 2-dimensional case, Schoen and Wolfson gave a variational analysis of area minimizing (Hamiltonian stationary) Lagrangian surfaces, \cite{Schoen2001}. They proved the existence of area minimizer with the particular forms of admissible conical singularities. Haskins and Kapouleas gave a gluing construction of the compact high-genus special Legendrian surfaces in the 5-sphere, \cite{Haskins2007}. Under the Hopf map $\s{5}\to\mathbb C\mathbb P^2$, these surfaces are mapped to minimal Lagrangian surfaces. In \cite{Loftin2013}, the minimal Lagrangian surfaces in the hyperbolic complex space form $\mathbb C\mathbb{H}^2$ were studied in relation to the surface group representations in $\liegroup{SU}(1,2)$. For the integrable system aspects of the theory on minimal Lagrangian tori, we refer to \cite{Carberry2004} and the references therein. \subsection{Formal Killing fields} \subsubsection{Polynomial Killing field} One of the characteristic structural properties of a harmonic torus in a symmetric space is the existence of an associated polynomial Killing field, \cite{Burstall1993}. For the case of a minimal Lagrangian torus in $\mathbb C\mathbb P^2$, a polynomial Killing field can be considered as a higher-order Gau\ss\, map which takes values in the polynomial loop algebra $\liealgebra{sl}(3,\mathbb C)[\lambda]$ (here $\lambda$ denotes the spectral parameter). The corresponding spectral curve is a branched triple covering of $\mathbb C\mathbb P^1$, and a minimal Lagrangian torus linearizes on its Jacobian. From this construction, the relevant spectral curve theory of finite type integration can be applied to the study of a minimal Lagrangian torus. \subsubsection{Formal Killing fields} For a general minimal Lagrangian surface in a $2_{\mathbb C}$-dimensional, non-flat, complex space form, it turns out that the local analytic data can be packaged into a pair of canonical \emph{formal} Killing fields, which take values in the formal loop algebra $\liealgebra{sl}(3,\mathbb C)[[\lambda]].$ For a minimal Lagrangian torus for example, each of these formal Killing fields would factor and reduce to a polynomial Killing field up to scaling by an element in $\mathbb C[[\lambda^6]]$. The original idea of canonical formal Killing fields (for CMC surfaces) is due to Pinkall and Sterling, \cite{Pinkall1989}. \subsection{Results} \subsubsection{A pair of canonical formal Killing fields}\label{sec:results1} We give a systematic derivation of the differential algebraic inductive formulas for a pair of canonical formal Killing fields for the differential system for minimal Lagrangian surfaces in a $2_{\mathbb C}$-dimensional, non-flat, complex space form, Thm.\ref{thm:FKformulaep4}, Thm.\ref{thm:FKformulaea5}. This clarifies the somewhat ad-hoc recursion formulas appeared in \cite{Pinkall1989}\cite{Wang2013}. \subsubsection{Jacobi fields and pseudo-Jacobi fields} In the course of analysis, we find two different kinds of Jacobi fields for the minimal Lagrangian system. Jacobi fields (ordinary), which correspond to the generalized symmetries of the minimal Lagrangian system, are defined by the operator, \eqref{eq:Jacobieq}, $$\partial_{\xi}\partial_{\ol{\xi}}+\frac{3}{2}\gamma^2.$$ Here the notations $\partial_{\xi}, \partial_{\ol{\xi}}$ denote the covariant derivatives with respect to the unitary $(1,0)$-form $\xi$, and its complex conjugate $\ol{\xi}$ respectively, \eqref{eq:delxnotation}, and $4\gamma^2$ is the holomorphic sectional curvature of the ambient complex space form. This shows that, when restricted to a minimal Lagrangian surface, Jacobi fields are the eigenfunctions of Laplacian of the induced Riemannian metric with eigenvalue $6\gamma^2$. A relevant observation is that the Jacobi operator depends only on the induced metric of the surface, similarly as in some of the calibrated geometries, \cite{McLean1998}. Pseudo-Jacobi fields, which correspond to the generalized symmetries of the elliptic Tzitzeica equation underlying the minimal Lagrangian system, are defined by the operator, \eqref{eq:Jacobieq'}, $$\partial_{\xi}\partial_{\ol{\xi}}+\frac{1}{2}(\gamma^2+4\vert{\rm I \negthinspace I}\vert^2).$$ Here $\vert{\rm I \negthinspace I}\vert^2$ is the squared norm of the associated Hopf differential, \S\ref{sec:Hopf}. Applying the previous results from \cite{Fox2011}\cite{Fox2012} on the elliptic Tzitzeica equation, we give a complete classification of the infinite sequence of (pseudo) Jacobi fields, Thm.\ref{thm:classifyJacobi}, Cor.\ref{cor:FKformulaep4}, Cor.\ref{cor:FKformulaea5}. As a corollary, this implies that a minimal Lagrangian torus in $\mathbb C\mathbb P^2$ admits a pair of spectral curves. \subsubsection{Conservation laws} We also obtain an infinite sequence of higher-order conservation laws from the components of the canonical formal Killing fields, Thm.\ref{thm:FKcvlaw}. \subsection{Recursion} We shall give a description of the two 3-step recursion relations embedded in the structure equation for the formal Killing fields. This is the main technical ingredient for our construction of the canonical formal Killing fields. \subsubsection{Previous works} The partial differential equation which locally describes the minimal Lagrangian surfaces in a $2_{\mathbb C}$-dimensional, non-flat, complex space form is the elliptic Tzitzeica equation, \eqref{eq:Tzitzeica}. An infinite sequence of higher-order symmetries and conservation laws were determined in the original works \cite{Fox2011}\cite{Fox2012} via a recursion modeled on the associated formal Killing field equation. On the other hand, this recursion process left the problem of integrating a sequence of exact differential 1-forms. We show that they can be solved differential algebraically without involving integration. \subsubsection{Recursive structure equation} For the case at hand, a formal Killing field is a (twisted) loop algebra $\liealgebra{sl}(3,\mathbb C)[[\lambda]]$-valued function $\textbf{X}_{\lambda}$ on the infinite prolongation space of the minimal Lagrangian system, which satisfies the Killing field equation \eqref{eq:KillingEquation}. In terms of an adapted basis of $\liealgebra{sl}(3,\mathbb C)[[\lambda]]$, the recursive component-wise structure equation \eqref{eq:formalKilling_n} can be summarized in the following infinite schematic diagram of period 6, Fig.\ref{fig:diagram0}. \begin{figure}[t] \begin{equation}\label{recursiondiagram} \begin{split} \xymatrix{ & b^{6n+5} \ar[ld]_{\partial_{\ol{\xi}}} \ar[rd]^{\partial_{\xi}} & && & s^{6n+9} \ar[ld]_{\partial_{\ol{\xi}}} \ar[rd]^{\partial_{\xi}} & \\ ...\quad p^{6n+4} & & f^{6n+6}\ar[r]^{\partial_{\xi}} & a^{6n+7}\ar[r]^{\partial_{\xi}}& g^{6n+8} && p^{6n+10}\quad... \\ & c^{6n+5}\ar[lu]^{\partial_{\ol{\xi}}} \ar[ru]_{\partial_{\xi}} & && & t^{6n+9}\ar[lu]^{\partial_{\ol{\xi}}}\ar[ru]_{\partial_{\xi}} & } \end{split}\notag \end{equation} \caption{Recursion diagram for formal Killing field} \label{fig:diagram0} \end{figure} Here the nodes $\{ \,p^*,b^*,c^*,f^*,a^*,g^*,s^*,t^*\,\}$ are the coefficients of a formal Killing field. Two nodes are connected by an arrow only if there exists a first order differential relation from the structure equation between them. From a node, one moves to the right by applying $\partial_{\xi}$, and to the left by applying $\partial_{\ol{\xi}}$. The upper indices are designated to match (roughly) the jet orders of the coefficients. The structure equation shows that the right-arrows are differential, which means that the coefficients $\{f^*, a^*, g^*, p^*\}$ are obtained from the left-adjacent term(s) by $\partial_{\xi}$ operation. But, the left-arrows decrease the jet order. In addition, note that \begin{align} \partial_{\xi} p^{6n+4}&= \textnormal{i}\gamma b^{6n+5}+2\textnormal{i} h_3 c^{6n+5},\notag\\ \partial_{\xi} g^{6n+8}&= - \textnormal{i} \gamma t^{6n+9}-\textnormal{i} h_3 s^{6n+9}.\notag \end{align} In order to continue the recursion process, one needs to solve for the coefficients $\{b^*, c^*, s^*, t^*\}$. \subsubsection{Differential algebraic inductive formulas} The main idea of construction is to impose the constraint in terms of the characteristic polynomial of $\textbf{X}_{\lambda}$; $$\det(\mu\rm{I}_3+\textbf{X}_{\lambda})=\mu^3+\textnormal{c}\lambda^3,$$ for a constant $\textnormal{c}\in\mathbb C^*$. This allows one to solve for $\{b^*, c^*, s^*, t^*\}$ differential algebraically not just by using the left-adjacent terms, but by using all of the lower-order terms (the left hand side terms in the diagram above). The relevant explicit formulas using the truncated formal Killing fields are given in \S\ref{sec:KFformulae}. \subsubsection{Jacobi fields and conservation laws} The structure equation \eqref{eq:formalKilling_n} implies that the sequence of coefficients $\{ \, a^{6n+7} \}$ are Jacobi fields, and the sequence of coefficients $\{ \, p^{6n+4} \,\}$ are pseudo-Jacobi fields. From this, it follows that the 6-step recursion in the schematic diagram in Fig.\ref{fig:diagram0} can be understood as the union of two 3-step recursions between Jacobi fields and pseudo-Jacobi fields. The structure equation \eqref{eq:formalKilling_n} also implies that an infinite sequence of conservation laws can be assembled from the components of the formal Killing fields, Eq.\eqref{eq:varphin}, Eq.\eqref{eq:varphin'}. \vspace{2mm} In this way, the canonical formal Killing fields provide an efficient method to organize the (infinitely prolonged) local analytic invariants of minimal Lagrangian surfaces. \subsection{Contents} In \S\ref{sec:setup}, the exterior differential system for minimal Lagrangian surfaces is defined as a homogeneous differential system on the bundle of Lagrangian 2-planes, and we record the basic structure equations. In \S\ref{sec:classicallaws}, we compute the space of classical Jacobi fields and conservation laws which arise from the infinitesimal action of the group of K\"ahler isometries of the ambient space form. In hindsight, the recursion relations for the formal Killing fields are already implicit in the structure equation for the classical Killing fields. In \S\ref{sec:prolongation1}, we determine the structure equation for the infinite prolongation of the minimal Lagrangian system. A branched triple cover is introduced for the field extension to accommodate the higher-order Jacobi fields. In \S\ref{sec:prolonglemmas}, we record two useful lemmas on the rigidity property of the associated $\partial_{\ol{\xi}}$-equation. They are applied in \S\ref{sec:Jacobifields} to prove a complete classification of the (pseudo) Jacobi fields. \S\ref{sec:formalKilling} contains the main results of the paper. We use the determinantal identities from the characteristic polynomial to derive the explicit differential algebraic recursion for the canonical formal Killing fields associated with a pair of natural initial data. In \S\ref{sec:highercvlaws1}, we also read off the formal Killing fields an infinite sequence of higher-order conservation laws. \subsection{Remarks} \subsubsection{} This is a continuation of the joint work \cite{Wang2013}. We suspect that the analysis carried out in \cite{Wang2013} can be extended to the primitive harmonic maps in general. \subsubsection{} In the analytic approach, the transition from the minimal surfaces in the 3-sphere to the minimal Lagrangian surfaces in $\mathbb C\mathbb P^2$ is nontrivial, \cite{Schoen2001}. From our point of view, this amounts to replacing the underlying Lie algebra from $\liealgebra{sl}(2,\mathbb C)$ to $\liealgebra{sl}(3,\mathbb C)$. The present work may provide a basis to introduce the further results from the integrable system theory to the study of minimal Lagrangian surfaces in this uniform perspective. \iffalse \subsubsection We suspect that the conservation laws for minimal Lagrangian surfaces (or more generally for Hamiltonian stationary Lagrangian surfaces) would have application to understanding the singularity of the area minimizing Lagrangian surfaces studied in \cite{Schoen2001}. \fi \section{Minimal Lagrangian surfaces in a complex space form}\label{sec:setup} After a brief summary of the structure equation for a $2_{\mathbb C}$-dimensional complex space form, we give an analytic description of the differential equation for minimal Lagrangian surfaces as a homogeneous exterior differential system defined on the $\liegroup{U}(2)/\liegroup{SO}(2)$-bundle of oriented Lagrangian planes. The basic structure equations established in \S\ref{sec:differential system}, together with their infinite prolongation in \S\ref{sec:prolongation1}, will be the basis of our analysis for the minimal Lagrangian system. \subsection{Complex space form}\label{sec:cspaceform} We will summarize the basic formulas of K\"ahlerian geometry for a $2_{\mathbb C}$-dimensional complex space form. We refer to \cite{Bryant2001} for the further related details. In order to avoid repetitions, we agree on the following range for the indices: \[ 1\leq A, B, C \leq 2.\] We use the Einstein summation convention for repeated indices. \subsubsection{$2_{\mathbb C}$-dimensional complex space form}\label{sec:2dcspaceform} Let $M$ be the $2_{\mathbb C}$-dimensional simply connected complex space form of constant holomorphic sectional curvature $4\gamma^2$. In the case $4\gamma^2=0$ and $M=\mathbb C^2$, it turns out that the minimal Lagrangian surfaces in $M$ are equivalent to the holomorphic curves in $\mathbb C^2$ under a different covariant constant complex structure (this is explained by the fact that $\mathbb C^2$ is hyperK\"ahler, \cite[p.148]{Joyce2000}). We shall restrict ourselves to the case $$\fbox{$\quad \gamma^2\ne0,\quad$}$$ where the differential equation for minimal Lagrangian surfaces is genuinely nonlinear. The squared expression $\gamma^2$ is introduced for the sake of convenience, for the quantity $\gamma$ (which is $\frac{1}{2}$-times the square root of the holomorphic sectional curvature) appears frequently in the analysis. We adopt the following convention for $\gamma$: \begin{equation}\label{1deRham1} \gamma= \begin{cases} &+\sqrt{\gamma^2}\\ &+\textnormal{i} \sqrt{- \gamma^2} \end{cases}\quad \textnormal{if}\quad \begin{array}{l} \mbox{$\gamma^2>0$} \\ \mbox{$\gamma^2<0$.} \end{array} \notag \end{equation} Here $\textnormal{i}=\sqrt{-1}$ denotes the unit imaginary number. \subsubsection{Unitary coframe bundle}\label{sec:unitarybundle} Let $\liegroup{U}(2)$ be the group of 2-by-2 unitary matrices. Let\vspace{2mm} \centerline{\xymatrix{\liegroup{U}(2) \ar[r] & \mathcal{F} \ar[d]^{\pi} \\ & M }} \vspace{2mm}\noindent be the principal $\liegroup{U}(2)$-bundle of unitary coframes. An element $\mathfrak{u}\in\mathcal F$ is by definition a Hermitian isometry \[\mathfrak{u}:T_{\pi(\mathfrak{u})}M\to \mathbb C^2,\] where $\mathbb C^2$ is the standard $2_{\mathbb C}$-dimensional Hermitian vector space. The structure group $\liegroup{U}(2)$ acts on $\mathcal F$ on the right by \[\mathfrak{u} \to g^{-1}\circ\mathfrak{u},\quad \mbox{for}\;g\in\liegroup{U}(2). \] Let $(\zeta^1,\,\zeta^2)^t$ be the $\mathbb C^2$-valued tautological 1-form on $\mathcal F$. The structure group $\liegroup{U}(2)$ acts on $(\zeta^1,\,\zeta^2)^t$ on the right by \[ \begin{pmatrix} \zeta^1 \\ \zeta^2 \end{pmatrix} \to g^{-1} \begin{pmatrix} \zeta^1 \\ \zeta^2 \end{pmatrix}, \quad \mbox{for}\;g\in\liegroup{U}(2). \] By definition, the K\"ahler structure on $M$ is given by the pair \begin{equation}\label{1varpi} \begin{array}{rll} \textnormal{g}:=& \zeta^A\circ\ol{\zeta}^A &\textnormal{(Riemannian metric)},\\ \varpi:=& \frac{\textnormal{i}}{2} \zeta^A{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \ol{\zeta}^A &\textnormal{(symplectic form)}. \end{array}\end{equation} \vspace{1mm} Let $\liealgebra{u}(2)$ be the space of 2-by-2 skew-Hermitian matrices, which is the Lie algebra of $\liegroup{U}(2)$. There exists a unique $\liealgebra{u}(2)$-valued connection 1-form $(\zeta^A_B)$ on $\mathcal F$ such that the following structure equations hold: \begin{align}\label{1strt1} {\rm d} \zeta^A &= - \zeta^A_B {\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \zeta^B, \\ \zeta^A_B&=-\ol{\zeta}^B_A. \notag \end{align} The curvature 2-forms $\Omega^A_B$ are then defined by \begin{equation}\label{1strt11} \Omega^A_B={\rm d} \zeta^A_B+\zeta^A_C{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \zeta^C_B. \end{equation} For the case at hand, the curvature forms of the complex space form $M$ are given by \begin{equation}\label{eq:1curv} \Omega^A_B= \gamma^2\left( \zeta^A{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \ol{\zeta}^B +\delta^A_{B}\sum_{C=1}^2 \zeta^C{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \ol{\zeta}^C \right). \end{equation} Here $\delta^A_{B}$ is the Kronecker delta. \subsubsection{Real structure equation}\label{sec:realstrt} The above treatment considers $M$ as a $2_{\mathbb C}$-dimensional complex Hermitian manifold. On the other hand, note the isomorphism $$\liegroup{U}(2)=\liegroup{SO}(4)\cap\liegroup{Sp}(2,\mathbb R). $$ Here the Lie groups $\liegroup{U}(2), \liegroup{SO}(4)$ $\textnormal{(special orthogonal group)}$, $\liegroup{Sp}(2,\mathbb R)$ (symplectic group) are considered as the real subgroups of $\liegroup{SL}(4,\mathbb R)$. Accordingly, it will be convenient for the analysis of Lagrangian surfaces to consider $M$ as a $4_{\mathbb R}$-dimensional real manifold equipped with the pair $(\textnormal{g}, \varpi)$ given by \eqref{1varpi}. \vspace{2mm} To this end, we decompose the structure equations for $\{ \zeta^A, \zeta^A_B\}$ into the real, and imaginary parts as follows. Set \begin{align} \zeta^A&=\omega^A+\textnormal{i} \mu^A, \notag\\ \begin{pmatrix} \zeta^1_1 & \zeta^1_2 \\ \zeta^2_1 & \zeta^2_2 \end{pmatrix} &=\begin{pmatrix} \cdot & \rho \\ -\rho &\cdot \end{pmatrix} +\textnormal{i} \begin{pmatrix} \beta^1-3\gamma^2\theta_0 & \beta^2 \\ \beta^2 & -\beta^1-3\gamma^2\theta_0 \end{pmatrix}, \notag \end{align} for the set of real 1-forms $\{ \omega^A, \mu^A, \rho, \beta^A, \theta_0\}$. In terms of these 1-forms, the structure equations \eqref{1strt1}, \eqref{1strt11}, \eqref{eq:1curv} are written as follows: \begin{align}\label{1strt0} {\rm d}\omega^1&=-\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2+(\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^1 +\beta^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^2)-3 \gamma^{2}\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^1, \\ {\rm d}\omega^2&=+\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1+(\beta^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^1-\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^2)-3 \gamma^{2}\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^2, \notag\\ {\rm d}\mu^1&=-\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^2-(\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1 +\beta^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2)+3 \gamma^{2}\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1, \notag\\ {\rm d}\mu^2&=+\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^1-(\beta^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1-\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2)+3 \gamma^{2}\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2,\notag \\ \notag\\ {\rm d}\rho&= \gamma^2(\omega^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2+\mu^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^2)+2\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\beta^2, \notag\\ \notag\\ {\rm d}\theta_0&= -(\mu^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1+\mu^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2), \notag\\ {\rm d}\beta^1&=-2\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\beta^2+\gamma^2(\mu^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1-\mu^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2),\notag\\ {\rm d}\beta^2&=+2\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\beta^1+\gamma^2(\mu^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1+\mu^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2).\notag \end{align} \subsection{Exterior differential system}\label{sec:differential system} With this preparation, we proceed to describe the differential system for minimal Lagrangian surfaces. \subsubsection{Bundle of Lagrangian 2-planes}\label{sec:Lagbundle} An immersed surface $\textnormal{x}: \Sigma\hookrightarrow M$ is Lagrangian if $$\textnormal{x}^*\varpi=0.$$ This is by definition a first order constraint on $\Sigma$ that its tangent space at each point is a Lagrangian subspace of the tangent bundle $TM$. Under the standard representation, the unitary group $\liegroup{U}(2)$ acts transitively on the set of oriented Lagrangian subspaces ($2$-planes) in $\mathbb C^2$, with $\liegroup{SO}(2)=\liegroup{U}(2)\cap \liegroup{SL}(2,\mathbb R)$ as the stabilizer subgroup. The Grassmannian of Lagrangian subspaces in dimension 2 is the homogeneous space, $$\textnormal{Lag}(\mathbb C^2)=\liegroup{U}(2)/\liegroup{SO}(2).$$ From the general theory of principal bundles, it follows that the $\liegroup{U}(2)/\liegroup{SO}(2)$-bundle of oriented Lagrangian 2-planes in $TM$ is given by $$X:=\mathcal F/\liegroup{SO}(2)\to M.$$ This is summarized by the following diagram: \vspace{2mm}\two \centerline{\xymatrix{ &&\mathcal F \ar[dl]_-{\liegroup{SO}(2)}\ar[dd]^{\liegroup{U}(2)} \\ & X \ar[dr]_-{\liegroup{U}(2)/\liegroup{SO}(2)} & \\ & &M}}\vspace{2mm}\two \subsubsection{Differential system for Lagrangian surfaces}\label{sec:Lagsystem} In view of the diagram, an immersed oriented Lagrangian surface in $M$ admits a unique tangential lift to $X$. Such tangential lifts are characterized as the integral surfaces of the canonical contact differential system on $X$. This is expressed analytically as follows. Consider the real structure equation \eqref{1strt0}. By a standard moving frame analysis, set the contact ideal $\mathcal I_0$ on $X$ generated by \begin{equation}\label{1ideal0} \mathcal I_0:=\langle\, \mu^1, \mu^2, {\rm d}\mu^1, {\rm d}\mu^2 \,\rangle. \end{equation} Note that the differential ideal $\mathcal I_0$ is originally defined on $\mathcal F$. But, it is invariant under the induced action by the structure group $\liegroup{SO}{(2)}$ and $\mathcal I_0$ descends on $X$ as a well defined differential ideal. By construction, an immersed oriented integral surface of $\mathcal I_0$ in $X$ corresponds to a possibly singular oriented Lagrangian surface in $M$. An immersed integral surface of $\mathcal I_0$ in $X$ may become singular under the projection $X\to M$ to the original complex space form. The formulation of Lagrangian surfaces in $M$ in terms of the canonical contact differential system $\mathcal I_0$ on $X$ naturally extends the set of admissible Lagrangian surfaces, see \S\ref{sec:doublecover}. \subsubsection{Differential system for minimal Lagrangian surfaces}\label{sec:mLagsystem} The condition that an immersed Lagrangian surface is minimal (i.e., its mean curvature vector vanishes) is a second order constraint. We therefore need to augment the contact ideal $\mathcal I_0$ to express this additional minimality condition. From Eqs.\eqref{1strt0}, we claim that this is equivalent to the vanishing of the connection 1-form $\theta_0$, $$\theta_0=0.$$ To see this, consider the structure equation \eqref{1strt0} adapted to a Lagrangian surface $\Sigma\hookrightarrow M$. With the 1-forms $\{\mu^1, \mu^2\}$ being set to 0, the induced Riemannian metric on $\Sigma$ is given by \begin{equation} \rm{I}:=(\omega^1)^2+(\omega^2)^2. \notag \end{equation} For the condition of minimality, note that the second fundamental form of $\Sigma$ can be identified with the symmetric cubic differential \begin{equation} \underline{{\rm I \negthinspace I}}:= \omega^A \circ \textnormal{Im}(\zeta^A_B) \circ \omega^B. \notag \end{equation} The mean curvature vector vanishes when the corresponding trace vanishes, $$\rm tr_{\rm{I}}(\underline{{\rm I \negthinspace I}})=0.$$ This is equivalent to the vanishing of the 1-form $\theta_0$, which is up to constant scale the trace of $\textnormal{Im}(\zeta^A_B).$ \begin{prop}\label{prop:mLagdefi Let $X\to M$ be the $\liegroup{U}(2)/\liegroup{SO}(2)$-bundle of oriented Lagrangian 2-planes. Let $\mathcal I_0$ be the canonical contact differential system on $X$, \eqref{1ideal0}. The differential system for minimal Lagrangian surfaces is given by \begin{equation}\label{1ideal1} \mathcal I:=\langle\, \mu^1, \mu^2, \theta_0, \phi^+, \phi^- \,\rangle, \end{equation} where \begin{align \phi^+&=\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1+\beta^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2, \notag \\ \phi^-&=\beta^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^1-\beta^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\omega^2. \notag \end{align} The structure equation \eqref{1strt0} shows that the differential ideal $\mathcal I$, originally defined on $\mathcal F$, is invariant under the induced action by the structure group $\liegroup{SO}{(2)}$. The differential ideal $\mathcal I$ is well defined on $X$. An immersed minimal Lagrangian surface in $M$ admits a unique tangential lift to $X$ as an integral surface of $\mathcal I$. Conversely, an immersed integral surface of $\mathcal I$ in $X$ projects to a possibly singular (branched) minimal Lagrangian surface in $M$. \end{prop} \begin{proof} For the last sentence, see \S\ref{sec:doublecover}. \end{proof} Note from the structure equation that $${\rm d} \theta_0, {\rm d}\mu^A \equiv 0\mod\mathcal I,$$ and $\mathcal I$ is differentially closed. In terms of Cartan's theory of exterior differential systems, the differential system $(X,\mathcal I)$ is involutive and the local moduli space of solutions depends on two arbitrary real functions of 1 variable, \cite{Bryant1991}. \subsubsection{Complexified structure equation}\label{sec:complexify} The differential system for minimal Lagrangian surfaces under consideration is an integrable extension over the elliptic Tzitzeica equation, \cite{Fox2011}. The characteristic directions are complex and they induce a complex structure on a minimal Lagrangian surface. In order to utilize this property, we introduce another set of complex differential forms on $\mathcal F$ adapted to $\mathcal I$. \vspace{2mm} Set \begin{align}\label{eq:cforms} \xi&:=\omega^1+\textnormal{i}\,\omega^2, \\ \theta_1&:=\mu^1-\textnormal{i}\mu^2, \notag\\ \eta_2&:=\beta^1-\textnormal{i}\beta^2.\notag \end{align} Note \[\phi^+ - \textnormal{i} \phi^-=\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi, \] and the ideal \eqref{1ideal1} can be written as \begin{equation}\label{1ideal2} \mathcal I=\langle \theta_0, \theta_1, \eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi \rangle. \end{equation} In terms of these complex 1-forms, the structure equation \eqref{1strt0} simplifies to the following set of equations: \begin{align}\label{eq:strt2} {\rm d} \xi &= \textnormal{i} \rho {\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \xi -3\gamma^2 \theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1-\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\eta}_2,\\ {\rm d} \theta_0&=-\frac{1}{2}\left( \theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi} \right), \notag\\ {\rm d} \theta_1 &=-\textnormal{i}\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_1-\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+3\gamma^2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi},\notag \\ {\rm d} \eta_2 &=-2\textnormal{i}\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\eta_2+\gamma^2\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}, \notag \\ {\rm d} \rho&=\frac{\textnormal{i}}{2}\left(\gamma^2\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}-2\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\eta}_2-\gamma^2\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1\right). \notag \end{align} From now on, the differential analysis for minimal Lagrangian surfaces will be carried out based on this structure equation. \vspace{2mm} Let us mention here a relevant notation which will be frequently used. For a scalar function $f:\mathcal F\to\mathbb C$, the covariant derivatives are written in the upper-index notations, \begin{equation}\label{eq:notation0} {\rm d} f \equiv f^{\xi}\xi+f^{\ol{\xi}}\ol{\xi}+f^0\theta_0+f^1\theta_1+f^{\bar{1}}\ol{\theta}_1+f^2\eta_2+f^{\bar{2}}\ol{\eta}_2 \mod \rho. \end{equation} \subsection{Hopf differential}\label{sec:HopfBonnet} The induced local geometric structures on a minimal Lagrangian surface consists of a triple of data called admissible triple, Defn.\ref{defn:Bonnettriple} in the below. In particular, it contains a holomorphic cubic differential called \emph{Hopf differential} which arises as a complexified version of the second fundamental form. Hopf differential will play an important role in our study of the minimal Lagrangian system. \subsubsection{Hopf differential}\label{sec:Hopf} Let $\textnormal{x}:\Sigma\hookrightarrow X$ be an immersed integral surface of the differential system for minimal Lagrangian surfaces. This is expressed analytically by $$\theta_0, \theta_1=0, \; \eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi=0,$$ on the induced $\liegroup{SO}(2)$-bundle $\textnormal{x}^*\mathcal F\to \Sigma$. It follows that the structure equation \eqref{eq:strt2} restricted to $\textnormal{x}^*\mathcal F$ becomes \begin{align}\label{eq:strtx} {\rm d}\xi&=\textnormal{i} \rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi,\\ 0&=\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi, \notag \\ {\rm d}\eta_2&=-2\textnormal{i} \rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\eta_2, \notag \\ {\rm d}\rho&=\frac{\textnormal{i}}{2}\left(\gamma^2\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi} - 2\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\eta}_2\right).\notag \end{align} In particular, the cubic differential $$\eta_2\circ \xi^2$$ is invariant under the action by the structure group $\liegroup{SO}(2)$ and becomes a well defined holomorphic cubic differential on $\Sigma.$ \vspace{2mm} Let $K\to\Sigma$ denote the canonical line bundle of $(1,0)$-forms. \begin{defn}\label{Hopf} Let $\textnormal{x}: \Sigma \hookrightarrow X$ be an immersed integral surface of the differential system for minimal Lagrangian surfaces. Consider the induced structure equation \eqref{eq:strtx}. The \textbf{Hopf differential} of $\textnormal{x}$ is the holomorphic cubic differential \begin{equation}\label{eq:Hopf} {\rm I \negthinspace I}=\eta_2 \circ \xi^2 \in H^0(\Sigma, K^3). \end{equation} The \textbf{umbilic divisor} $\mathcal U =({\rm I \negthinspace I})_0$ is defined as the zero divisor of ${\rm I \negthinspace I}$. \end{defn} When $\Sigma$ is compact, by Riemann-Roch theorem we have $$\textnormal{deg}(\mathcal U)=6\, \textnormal{genus}(\Sigma)-6.$$ \subsubsection{Admissible triple}\label{sec:Bonnet} Suppose that the integral surface $\Sigma\hookrightarrow X$ is the tangential lift of an immersed oriented minimal Lagrangian surface in $M$. By definition of the tautological forms on $\mathcal F$, this implies the independence condition $$\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}\ne 0,$$ and the induced Riemannian metric on $\Sigma$ is given by $$\rm{I}=\xi\circ\ol{\xi}.$$ It follows that $\textnormal{x}^*\mathcal F\to \Sigma$ can be identified with the principal $\liegroup{SO}(2)$-bundle of the induced metric, and that $\xi$ is the tautological unitary $(1,0)$-form, and $\rho$ is the Levi-Civita connection form. From the second equation of \eqref{eq:strtx}, there exists a coefficient $h_3$ defined on $\textnormal{x}^*\mathcal F$ such that \begin{equation}\label{eq:etah3}\eta_2= h_3 \xi.\end{equation} The Hopf differential is now written as $${\rm I \negthinspace I}=h_3\xi^3.$$ From the fourth equation of \eqref{eq:strtx}, the Gau\ss\, equation, the curvature $R$ of the induced metric $\rm{I}$ satisfies the compatibility equation \begin{equation}\label{eq:Gauss} R=\gamma^2- 2 h_3\bar h_3. \end{equation} \vspace{1mm} Summarizing the analysis so far, we give a definition of the compatible Bonnet data for minimal Lagrangian surfaces. \begin{defn}\label{defn:Bonnettriple} Let $M$ be a $2_{\mathbb C}$-dimensional complex space form of constant holomorphic sectional curvature $4\gamma^2.$ An \textbf{admissible triple} for a minimal Lagrangian surface in $M$ consists of a Riemann surface, a conformal metric, and a holomorphic cubic differential which satisfy the compatibility equation \eqref{eq:Gauss}. \end{defn} An analogue of the classical Bonnet theorem can be stated as follows. The proof is by a standard ODE argument and is omitted. \begin{thm}\label{thm:Bonnet Let $\Sigma$ be a Riemann surface. Let $( \rm{I}, {\rm I \negthinspace I})$ be a pair of a conformal metric and a holomorphic cubic differential on $\Sigma$ such that $(\Sigma, \rm{I}, {\rm I \negthinspace I})$ form an admissible triple for a minimal Lagrangian surface in a complex space form $M$. Let $\pi: \widetilde{\Sigma}\to\Sigma$ be the simply connected universal cover. Then there exists a minimal Lagrangian immersion $\tilde{\textnormal{x}}: \widetilde{\Sigma}\hookrightarrow M$ which realizes $(\pi^*\rm{I}, \pi^*{\rm I \negthinspace I})$ as the induced Riemannian metric and the Hopf differential. Such an immersion $\tilde{\textnormal{x}}$ is unique up to motion by the K\"ahler isometries of the ambient complex space form $M$. \end{thm} \subsubsection{Cube root of Hopf differential} The analysis of the canonical formal Killing fields to be addressed in \S\ref{sec:formalKilling} will inevitably involve the use of the object $$\sqrt[3]{h_3},$$ or, equivalently, the cube root of Hopf differential $$\sqrt[3]{{\rm I \negthinspace I}}.$$ It is generally a multi-valued holomorphic 1-form on a minimal Lagrangian surface. In order to better accommodate this, we introduce the triple cover of a minimal Lagrangian surface defined by Hopf differential. \begin{defn}\label{defn:doublecover} Let $\textnormal{x}:\Sigma\hookrightarrow X$ be an immersed integral surface of the differential system for minimal Lagrangian surfaces. Let ${\rm I \negthinspace I}\in H^0(\Sigma, K^3)$ be the Hopf differential \eqref{eq:Hopf}. The \textbf{triple cover} $$\nu:\hat{\Sigma}\to\Sigma$$ associated with ${\rm I \negthinspace I}$ is the Riemann surface of the complex curve \[ \Sigma'= \left\{ \kappa \in K\; \vert \; \kappa^3={\rm I \negthinspace I} \right\} \subset K. \] The \textbf{cube root $\omega=\sqrt[3]{{\rm I \negthinspace I}}$} of the Hopf differential is the holomorphic 1-form on $\hat{\Sigma}$ obtained by the pull-back of restriction of the tautological 1-form on $K$ to $\Sigma'$. By construction, the projection $\nu$ is a triple covering branched over the umbilics of degree $1$ or $2\mod 3$. \end{defn} \section{Classical Killing fields} \label{sec:classicallaws} Let $$G=\liegroup{SU}(3), \quad \textnormal{or}\quad \liegroup{SU}(1,2)$$ be the group of K\"ahler isometries of the complex space form $M$, depending on the sign of the holomorphic sectional curvature $4\gamma^2.$ The induced action of $G$ on $X$ is transitive, and by construction the differential system $(X,\mathcal I)$ is $G$-invariant, i.e., the group $G$ acts as a symmetry of the differential system for minimal Lagrangian surfaces. In this section, we examine the classical Killing fields generated by the corresponding infinitesimal action of the Lie algebra of $G$, $$\liealgebra{g} = \liealgebra{su}(3), \quad \textnormal{or}\quad\liealgebra{su}(1,2).$$ \vspace{1mm} The structure equation for classical Killing fields reveals its recursive, symmetrical property when it is written in terms of an adapted basis of $\liealgebra{g}$, Eq.\eqref{eq:classicalKF}. From this, we extract two kinds of Jacobi fields called \emph{classical Jacobi fields}, and their variants \emph{classical pseudo-Jacobi fields}, as well as the recursion relations between them, \S\ref{sec:recursionrel}. The structure equation also shows that there exists the eight dimensional space of classical conservation laws. Combining these results, we deduce that Noether's theorem holds and there exists a canonical isomorphism between the classical (pseudo) Jacobi fields and the classical (local) conservation laws, Cor.\ref{cor:Noether}. \iffalse The classical Jacobi fields correspond to the infinitesimal minimal Lagrangian deformation. From the well known fact that the normal bundle of a Lagrangian submanifold in a K\"ahler manifold is canonically isomorphic to its tangent bundle, one finds that the second order elliptic differential equation satisfied by the classical Jacobi fields (see Eq.\eqref{eq:classicalJacobifield}) is intrinsic to the minimal Lagrangian surface depending only on the induced Riemannian metric but not on the Hopf differential. The differential equation for classical pseudo-Jacobi fields on the other hand involves Hopf differential. The analysis of classical Killing fields, Jacobi fields, pseudo-Jacobi fields, conservation laws, and Noether's theorem will be extended to the infinitely prolonged higher-order counterparts in \S\ref{sec:formalKilling}. It will be shown that the minimal Lagrangian system admits an infinite number of their higher-order counterparts. \fi \subsection{Structure equation}\label{sec:classicalstrt} The Killing field equation, \eqref{eq:KillingEquation0} in the below, characterizes the classical Killing fields generated by the infinitesimal action of the Lie algebra $\liealgebra{g}$ on the homogeneous space $X$. When restricted to a minimal Lagrangian surface, it becomes the recursive structure equation \eqref{eq:classicalKF}. We will find that Eq.\eqref{eq:classicalKF} encodes many of the important local structural properties of the minimal Lagrangian system. In particular, a pair of the embedded recursion relations will serve as the model for the construction of the canonical formal Killing fields. \subsubsection{Complexified Maurer-Cartan form} Recall the complexified differential forms \eqref{eq:cforms} and their structure equation \eqref{eq:strt2}. The left invariant, $\liealgebra{g}$-valued Maurer-Cartan form of $G$ can be written in terms of these 1-forms by \begin{equation}\label{eq:psiform} \psi=\psi_{+}+\psi_{0}+\psi_{-}, \end{equation} where \begin{equation} \psi_{-} =\frac{1}{2} \left[ \begin {array}{ccc} \cdot &-\gamma(\xi-\textnormal{i}\theta_1) & \textnormal{i} \gamma( \xi+\textnormal{i} \theta_1) \notag \\ \gamma(\xi+\textnormal{i}\theta_1) &\textnormal{i} \eta_2 & - \eta_2 \\ -\textnormal{i}\gamma(\xi-\textnormal{i}\theta_1) &- \eta_2 &- \textnormal{i} \eta_2 \end {array} \right], \end{equation} \begin{equation} \psi_{0} = \left[ \begin {array}{ccc} 2\textnormal{i}\gamma^2\theta_0 &\cdot & \cdot \notag \\ \cdot &-\textnormal{i}\gamma^2\theta_0& \rho \\ \cdot & -\rho &-\textnormal{i}\gamma^2\theta_0 \end {array} \right], \notag \end{equation} \begin{equation} \psi_{+}=\frac{1}{2} \left[ \begin {array}{ccc} \cdot &-\gamma(\ol{\xi}-\textnormal{i}\ol{\theta}_1) &- \textnormal{i} \gamma( \ol{\xi}+\textnormal{i} \ol{\theta}_1) \notag \\ \gamma(\ol{\xi}+\textnormal{i}\ol{\theta}_1) &\textnormal{i} \ol{\eta}_2 & \ol{\eta}_2 \\ \textnormal{i}\gamma(\ol{\xi}-\textnormal{i}\ol{\theta}_1) & \ol{\eta}_2 &- \textnormal{i} \ol{\eta}_2 \end {array} \right]. \end{equation} $\,$ \noindent It satisfies the structure equation $${\rm d}\psi+\psi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\psi=0.$$ \subsubsection{Classical Killing fields} \begin{defn} Let $\mathcal F/\liegroup{SO}(2)=X\to M$ be the bundle of oriented Lagrangian planes. A \textbf{classical Killing field} is an $\liegroup{SO}(2)$-equivariant, $\liealgebra{g}$-valued function $\textbf{X}$ on $\mathcal F$ which satisfies the Killing field equation \begin{equation}\label{eq:KillingEquation0} {\rm d} \textbf{X} +[\psi, \textbf{X}]=0. \end{equation} \end{defn} By definition, the space of classical Killing fields is isomorphic to $\liealgebra{g}$. \begin{defn} Let $(X,\,\mathcal I)$ be the differential system for minimal Lagrangian surfaces. A vector field $V \in H^0(TX)$ is a \textbf{classical symmetry} if it preserves the differential ideal $\mathcal I$ under the Lie derivative, $$\mathcal L_V \mathcal I \subset \mathcal I.$$ The algebra of classical symmetries is denoted by $\mathfrak{S}^{(0)}.$ \end{defn} It is clear that a classical Killing field generates a classical symmetry. In fact, a classical symmetry is necessarily generated by a classical Killing field. We thus have; \begin{prop}\label{prop:classicalsymmetrytocv} $$ \mathfrak{S}^{(0)}\simeq \liealgebra{g}.$$ \end{prop} The proof follows by a direct computation, and is omitted. \subsubsection{Recursive structure equation} For the ensuing analysis, it will be useful to have Eq.\eqref{eq:KillingEquation0} written component-wise. Consider the explicit decomposition of the Lie algebra $\liealgebra{g}$ in Fig.\ref{fig:Xdecompo0}.\\ \begin{figure}[h] $$ \left[ \begin {array}{ccc} -2\textnormal{i} \textbf{a} &\textbf{b}+\textbf{f}+\textbf{g}-\textbf{s} &\textnormal{i} \textbf{b}- \textnormal{i} \textbf{f}+\textnormal{i} \textbf{g}+\textnormal{i} \textbf{s}\\ \noalign{\medskip}-\textbf{b}+\textbf{f}+\textbf{g}+\textbf{s} & \textnormal{i} \textbf{c}+\textnormal{i} \textbf{a}-\textnormal{i} \textbf{t}&-\textbf{p}+\textbf{c}+\textbf{t}\\ \noalign{\medskip}-\textnormal{i} \textbf{b}-\textnormal{i} \textbf{f} +\textnormal{i} \textbf{g}-\textnormal{i} \textbf{s} &\textbf{p}+\textbf{c}+\textbf{t}&-\textnormal{i} \textbf{c}+\textnormal{i} \textbf{a}+\textnormal{i} \textbf{t}\end {array} \right] $$ \caption{Decomposition of the Lie algebra $\liealgebra{g}=\liealgebra{su}(3),\;\textnormal{or}\;\, \liealgebra{su}(1,2)$} \label{fig:Xdecompo0} \end{figure} \noindent Here $\{ \, \textbf{p}, \textbf{b}, \textbf{c}, \textbf{f}, \textbf{a}, \textbf{g}, \textbf{s}, \textbf{t} \,\}$ are the scalar coefficients which satisfy the following reality conditions: \begin{equation} \liealgebra{g}=\begin{cases} \liealgebra{su}(3) \\ \liealgebra{su}(1,2) \end{cases} \textnormal{then}\quad \begin{aligned} &\textbf{a}=\overline{\textbf{a}},\textbf{p}=\overline{\textbf{p}}, \textbf{t}=-\overline{\textbf{c}}, \textbf{s}=-\overline{\textbf{b}}, \textbf{g}=-\overline{\textbf{f}}, \\ &\textbf{a}=\overline{\textbf{a}},\textbf{p}=\overline{\textbf{p}}, \textbf{t}=-\overline{\textbf{c}}, \textbf{s}=+\overline{\textbf{b}}, \textbf{g}=+\overline{\textbf{f}}. \end{aligned}\notag \end{equation} The Killing field equation \eqref{eq:KillingEquation0} then implies \begin{align}\label{eq:classicalKF} {\rm d} \textbf{p} &\equiv(\textnormal{i}\gamma \textbf{b}+2\textnormal{i} h_3 \textbf{c} )\xi + ( \textnormal{i}\gamma \textbf{s}+2\textnormal{i} \bar h_3 \textbf{t})\ol{\xi}, \\ {\rm d} \textbf{b} +\textnormal{i} \textbf{b}\rho &\equiv\textnormal{i} h_3 \textbf{f} \xi+ \frac{\textnormal{i}}{2}\gamma \textbf{p} \ol{\xi}, \notag\\ {\rm d} \textbf{c} -2 \textnormal{i} \textbf{c}\rho&\equiv \textnormal{i} \gamma \textbf{f}\xi +\textnormal{i} \bar h_3 \textbf{p} \ol{\xi}, \notag\\ {\rm d} \textbf{f} -\textnormal{i} \textbf{f}\rho &\equiv\frac{3\textnormal{i}}{2}\gamma \textbf{a} \xi +(\textnormal{i} \gamma \textbf{c}+\textnormal{i}\bar h_3 \textbf{b})\ol{\xi}, \notag\\ {\rm d} \textbf{a} &\equiv\textnormal{i} \gamma \textbf{g}\xi + \textnormal{i} \gamma \textbf{f}\ol{\xi}, \notag\\ {\rm d} \textbf{g}+\textnormal{i} \textbf{g}\rho &\equiv (-\textnormal{i} \gamma \textbf{t}-\textnormal{i} h_3 \textbf{s})\xi +\frac{3\textnormal{i}}{2}\gamma \textbf{a}\ol{\xi}, \notag\\ {\rm d} \textbf{s} -\textnormal{i} \textbf{s}\rho&\equiv \frac{\textnormal{i}}{2} \gamma \textbf{p}\xi -\textnormal{i}\bar h_3 \textbf{g}\ol{\xi}, \notag\\ {\rm d} \textbf{t}+2 \textnormal{i} \textbf{t}\rho&\equiv\textnormal{i} h_3 \textbf{p}\xi-\textnormal{i} \gamma \textbf{g}\ol{\xi}, \notag\qquad\mod\mathcal I. \end{align} Here we applied the substitution, \eqref{eq:etah3}, $$\eta_2\equiv h_3\xi,\quad \ol{\eta}_2\equiv \bar h_3\ol{\xi} \qquad("\textnormal{mod}\;\mathcal I").$$ The resulting structure equation should be understood as restricted to a minimal Lagrangian surface. \vspace{2mm} Let us introduce the relevant notations for partial derivatives. For a scalar function $u$ on $\mathcal F$, the notations $\partial_{\xi} u, \partial_{\ol{\xi}} u$ would mean \begin{align}\label{eq:delxnotation} \partial_{\xi} u =u_{\xi} &:=\textnormal{ $\xi$-coefficient of ${\rm d} u$}, \\ \partial_{\ol{\xi}} u =u_{\ol{\xi}}&:=\textnormal{ $\ol{\xi}$-coefficient of ${\rm d} u$, \; "\textnormal{mod}\;$\mathcal I$}".\notag \end{align} For instance, the above structure equation shows that $$ \textbf{a}_{\xi}=\textnormal{i}\gamma\textbf{g}, \quad \textbf{a}_{\ol{\xi}}=\textnormal{i}\gamma\textbf{f}.$$ \subsection{Classical Jacobi fields and pseudo-Jacobi fields}\label{sec:classicalJacobi} From the apparent recursive symmetry of Eq.\eqref{eq:classicalKF}, we extract two different kinds of Jacobi fields for the minimal Lagrangian system; classical Jacobi fields and classical pseudo-Jacobi fields. \subsubsection{Classical Jacobi field} Consider the coefficient $\textbf{a}$ in Eq.\eqref{eq:classicalKF}. One finds \begin{align*} \textbf{a}_{\xi}&=\textnormal{i}\gamma\textbf{g}, \\ \textbf{a}_{\xi,\ol{\xi}}&=\textnormal{i}\gamma\textbf{g}_{\ol{\xi}}=-\frac{3}{2}\gamma^2\textbf{a}. \end{align*} Motivated by this, we make the following definition. \begin{defn}\label{defn:classicalJacobifield} A scalar function $A:X\to\mathbb C$ is a \textbf{classical Jacobi field} if it satisfies \begin{equation}\label{eq:classicalJacobifield} A_{\xi,\ol{\xi}}+\frac{3}{2}\gamma^2A=0. \end{equation} The $\mathbb C$-vector space of classical Jacobi fields is denoted by $\mathfrak{J}^{(0)}=\mathfrak{J}^0$. \end{defn} An examination of Eq.\eqref{eq:KillingEquation0} shows that the coefficient $\bba$ generates the classical Killing field $\textbf{X}$ in the sense that if the $\textbf{a}$-component vanishes then the corresponding classical Killing field $\textbf{X}$ vanishes. It follows that the induced map $$\bba:\liealgebra{g}^{\mathbb C} \longrightarrow \mathfrak{J}^{(0)}$$ is injective. Here $\liealgebra{g}^{\mathbb C}=\liealgebra{g}\otimes\mathbb C=\liealgebra{sl}(3,\mathbb C)$ is the complexification of $\liealgebra{g}$. We claim that this is in fact an isomorphism, and the $\textbf{a}$-components of the (complexified) classical Killing fields generate the classical Jacobi fields. \begin{prop}\label{prop:classicalJg} \begin{equation}\label{eq:J0=g} \mathfrak{J}^{(0)}\simeq \liealgebra{g}^{\mathbb C}. \end{equation} \end{prop} The claim follows by a direct computation. Rather than a complete proof, we give a brief sketch of ideas. Let $A$ be a classical Jacobi field, which is defined on $X$. Denote the covariant derivatives of $A$ by $${\rm d} A=A^{\xi}\xi+A^{\ol{\xi}}\ol{\xi}+A^0\theta_0+A^1\theta^1+A^{\overline{1}}\ol{\theta}_1 +A^2\eta_2+A^{\overline{2}}\ol{\eta}_2.$$ We adopt the similar notation for the successive derivatives of $A^{\xi}, A^{\xi, \ol{\xi}}, \, ...$, etc. Applying the covariant derivative operators \eqref{eq:delxnotation}, one finds \begin{align*} \partial_{\xi} A=A_{\xi}&=A^{\xi}+A^2 h_3,\\ \partial_{\ol{\xi}}(A_{\xi})=A_{\xi, \ol{\xi}}&=A^{\xi,\ol{\xi}}+A^{\xi,\overline{2}}\bar h_3+(A^{2,\ol{\xi}}+A^{2,\overline{2}}\bar h_3) h_3. \end{align*} The Jacobi equation \eqref{eq:classicalJacobifield} becomes \begin{equation}\label{eq:separateJB} \Big(A^{\xi,\ol{\xi}}+\frac{3}{2}\gamma^2A\Big) +A^{\xi,\overline{2}}\bar h_3+ A^{2,\ol{\xi}}h_3+A^{2,\overline{2}}\bar h_3 h_3=0.\end{equation} Since $A$ is defined on $X$ while $h_3,\bar h_3$ are the prolongation variables,\footnotemark\footnotetext{See \S\ref{sec:prolongation1}.} each coefficient of $\{ h_3,\bar h_3, h_3\bar h_3\}$ in \eqref{eq:separateJB} must vanish separately. As a result, a classical Jacobi field must satisfy; $$A^{\xi,\ol{\xi}}+\frac{3}{2}\gamma^2A=0, \; A^{\xi,\overline{2}}=A^{2,\ol{\xi}}=A^{2,\overline{2}}=0.$$ From this, a standard over-determined PDE analysis shows that $A$ is the $\bba$-component of a (complexified) classical Killing field. \subsubsection{Classical pseudo-Jacobi fields} In an analogy with the classical Jacobi field case, consider next the coefficient $\textbf{p}$ in Eq.\eqref{eq:classicalKF}. One finds \begin{align*} \textbf{p}_{\xi}&=\textnormal{i}\gamma \textbf{b}+2\textnormal{i} h_3 \textbf{c} , \\ \textbf{p}_{\xi,\ol{\xi}}&=\textnormal{i}\gamma \textbf{b}_{\ol{\xi}}+2\textnormal{i} h_3 \textbf{c}_{\ol{\xi}} =-\frac{1}{2}(\gamma^2+4 h_3\bar h_3)\textbf{p}. \end{align*} \begin{defn}\label{defn:classicalpseudoJacobifield} A scalar function $P:X\to\mathbb C$ is a \textbf{classical pseudo-Jacobi field} if it satisfies \begin{equation}\label{eq:classicalpsJacobifield} P_{\xi,\ol{\xi}}+\frac{1}{2}(\gamma^2+4 h_3\bar h_3)P=0. \end{equation} The $\mathbb C$-vector space of classical pseudo-Jacobi fields is denoted by $\mathfrak{J}'^{(0)} =\mathfrak{J}'^{0}$. \end{defn} By the similar argument as above, the $\textbf{p}$-components of the (complexified) classical Killing fields generate the classical pseudo-Jacobi fields. \begin{prop}\label{prop:classicalpsJ=g} \begin{equation}\label{eq:J0ps=g} \mathfrak{J}'^{(0)}\simeq \liealgebra{g}^{\mathbb C}. \end{equation} \end{prop} As a consequence of the isomorphisms \eqref{eq:J0=g}, \eqref{eq:J0ps=g}, we have \begin{cor}\label{cor:J0J0ps=g} $$\mathfrak{J}^{(0)}\simeq\mathfrak{J}'^{(0)}\simeq \liealgebra{g}^{\mathbb C}.$$ \end{cor} \subsubsection{Recursion relations}\label{sec:recursionrel} The structure equation \eqref{eq:classicalKF} contains two 3-step recursion relations between the classical pseudo-Jacobi field $\bbp$ and the classical Jacobi field $\bba$. We record and emphasize these structures here, for they are the main technical ingredients in the construction of the formal Killing fields later on. \vspace{1mm} [\textbf{From $\bbp$ to $\bba$}] Let the classical pseudo-Jacobi field $\bbp$ be given. From \eqref{eq:classicalKF}, suppose (either of) the equations $$\partial_{\ol{\xi}}\bbb=\frac{\textnormal{i}}{2}\gamma\bbp,\quad \partial_{\ol{\xi}}\bbc=\textnormal{i} \bar h_3\bbp$$ were solved. Differentiating $\bbb, \bbc$, $$\partial_{\xi}\bbb=\textnormal{i} h_3\bbf,\quad\partial_{\xi}\bbc=\textnormal{i}\gamma\bbf,$$ and one gets $\bbf$. Differentiating $\bbf$, $$\partial_{\xi}\bbf=\frac{3\textnormal{i}}{2}\gamma\bba,$$ and one gets the classical Jacobi field $\bba$. The process can be summarized by the following diagram. $$\bbp \xrightarrow{\;\;\;\;\partial_{\ol{\xi}}^{-1}\;\;\;\;} \bbb, \bbc \xrightarrow{\quad\partial_{\xi}\quad} \bbf \xrightarrow{\quad\partial_{\xi}\quad} \bba. $$ \vspace{2mm} [\textbf{From $\bba$ to $\bbp$}] In a similar way, starting from the classical Jacobi field $\bba$, one may reach $\bbp$ by the following process. $$\bba \xrightarrow{\quad\partial_{\xi}\quad} \bbg \xrightarrow{\;\;\;\;\partial_{\ol{\xi}}^{-1}\;\;\;\;} \bbs, \bbt \xrightarrow{\quad\partial_{\xi}\quad} \bbp. $$ \vspace{2mm} Note that by combining the two processes, one gets a 6-step recursion relation for classical (pseudo) Jacobi fields. \subsection{Classical conservation laws}\label{sec:classicalcvlaw} Another important invariants of the minimal Lagrangian system are conservation laws. They also can be read off Eq.\eqref{eq:classicalKF}. \subsubsection{Definition} Let $(\Omega^*(X), {\rm d})$ be the de-Rham complex of $\mathbb C$-valued differential forms on $X$. Since $\mathcal I$ is a differential ideal closed under the exterior derivative, the quotient complex \[(\underline{\Omega}^*,\, \underline{{\rm d}}) \] is well defined, where $\underline{\Omega}^*=\Omega^*(X)/\mathcal I$, and $\underline{{\rm d}}={\rm d}\mod \mathcal I$. Let $H^{q}(\underline{\Omega}^*,\, \underline{{\rm d}})$ be the cohomology at $\underline{\Omega}^q.$ The set $$\{ \;H^{q}(\underline{\Omega}^*,\, \underline{{\rm d}}) \;\}_{q=0}^2$$ is called the \emph{characteristic cohomology} of the differential system $(X,\mathcal I)$. \begin{defn}\label{defn:classicalcvlaw} Let $(X,\mathcal I)$ be the differential system for minimal Lagrangian surfaces. A \textbf{classical conservation law} is an element of the 1-st characteristic cohomology $H^1(\underline{\Omega}^*,\, \underline{{\rm d}})$ of $(X,\mathcal I)$. The $\mathbb C$-vector space of classical conservation laws is denoted by \[ \mathcal C^{(0)}= \mathcal C^{0}:=H^1(\underline{\Omega}^*,\, \underline{{\rm d}}).\] Let $\mathcal C^{(0)}_{loc}$ denote the space of local classical conservation laws of $\mathcal I$ restricted to a small contractible open subset of $X$. \end{defn} \subsubsection{Classical conservation laws from classical Killing fields} From the structure equation \eqref{eq:classicalKF}, consider the 1-form \begin{equation}\label{eq:cvlawphi_a} \varphi_{\textbf{a}}=\textbf{b}\xi+\textbf{s}\ol{\xi}. \end{equation} One finds that $${\rm d} \varphi_{\textbf{a}}\equiv 0\mod\mathcal I,$$ and the 1-form $\varphi_{\textbf{a}}$ represents a classical conservation law. Let $[\varphi_{\textbf{a}}]\in\mathcal C^{(0)}_{loc}$ denote the class represented by $\varphi_{\textbf{a}}$ (which is globally defined on $X$). We claim that the associated map $$\liealgebra{g}^{\mathbb C} \simeq \mathfrak J^{(0)} \longrightarrow \mathcal C^{(0)}_{loc}$$ given by $$\textbf{a} \longrightarrow [\varphi_{\textbf{a}}]$$ is an isomorphism. In order to verify this claim, and for other computational purposes, we introduce a differentiated version of conservation laws. \subsection{Classical differentiated conservation laws}\label{sec:cdcvlaw} From the exact sequence $$0\to \mathcal I \to \Omega^*(X)\to \Omega^*(X)/\mathcal I=\underline{\Omega}^*\to 0,$$ at least locally we have $$\mathcal C^{(0)}_{loc}\simeq H^2(\mathcal I, {\rm d}).$$ Here $H^*(\mathcal I,{\rm d})$ denotes the cohomology of the complex $(\Omega^*(\mathcal I), {\rm d}).$ \begin{defn}\label{defn:classicaldcvlaw} Let $(X,\mathcal I)$ be the differential system for minimal Lagrangian surfaces. A \textbf{classical differentiated conservation law} is an element in the $2$-nd cohomology $H^2(\mathcal I, {\rm d}).$ The $\mathbb C$-vector space of classical differentiated conservation laws is denoted by $$ \mathcal H^{(0)}=\mathcal H^0:= H^2(\mathcal I, {\rm d}).$$ \end{defn} Note from the definition of classical conservation laws that a representative 1-form $\varphi$ of a class $[\varphi]\in\mathcal C^{(0)}_{loc}$ is defined up to exact 1-forms. By taking the exterior derivative ``${\rm d}$" and considering the differentiated conservation laws, we eliminate this ambiguity, which is practically important in the actual computation of conservation laws. Moreover, in a geometric situation as in the present case, the ideal $\mathcal I$ is often equipped with the relevant structures which enable one to determine a subspace of $\Omega^2(\mathcal I)$ in which a differentiated conservation law has the unique representative. In other words, one of the advantages of the differentiated version of conservation laws is that it may lead to an analogue of Hodge theorem; the appropriately normalized differentiated conservation laws can be considered as the \emph{harmonic} representatives. \begin{rem} Consider the symplectic form of $M$, \eqref{1varpi}: \begin{align*} \varpi&= \frac{\textnormal{i}}{2} \zeta^A{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \ol{\zeta}^A \\ &=\omega^1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^1+\omega^2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\mu^2 = {\rm d}\theta_0. \end{align*} Thus the class $[\varpi]$ is trivial in $H^{2}(\mathcal I,\,{\rm d}).$ \end{rem} We shall first establish the isomorphism $$\mathcal H^{(0)}\simeq\liealgebra{g}^{\mathbb C} $$ by a direct differential analysis. Then, by finding a section (up to constant scale) of the natural map \begin{equation}\label{eq:varphiPhimap} [\varphi_{\textbf{a}}]\in\mathcal C^{(0)}_{loc} \longrightarrow [{\rm d} \varphi_{\textbf{a}}]\in\mathcal H^{(0)},\end{equation} we show that there exists an isomorphism $$\mathcal C^{(0)}_{loc}\simeq \mathcal H^{(0)}.$$ \subsubsection{Initial reduction} Let $\Phi\in \Omega^2(\mathcal I)$ be a 2-form in the ideal. Up to addition by ${\rm d}( \Omega^1(\mathcal I))$, an exact 2-form in the ideal of the form ${\rm d}(f_0\theta_0+f_1\theta_1+f_{\bar{1}}\ol{\theta}_1)$ for scalar coefficients $f_0, f_1, f_{\bar{1}}$, a computation shows that one may write \begin{equation}\label{eq:H0} \Phi=A \Psi + \theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\sigma \end{equation} where $A$ is a scalar function, $\sigma\in\Omega^1(X)$, and $$\Psi=\textnormal{Im}(\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi)=-\frac{\textnormal{i}}{2}(\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi-\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}). $$ Note \begin{align}\label{1dPsi} {\rm d}\Psi&= 3\textnormal{i}\gamma^2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}(\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}+\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1) \\ &\equiv 0\mod\theta_0. \notag \end{align} A 2-form $\Phi\in\Omega^2(\mathcal I)$ normalized as in \eqref{eq:H0} is called a \emph{reduced 2-form}. Let $$H^{(0)}\subset \Omega^2(\mathcal I)$$ denote the subspace of such reduced 2-forms. By construction, $H^{(0)}$ is transversal to the subspace of exact 2-forms ${\rm d}( \Omega^1(\mathcal I))\subset \Omega^2(\mathcal I)$. It follows that we have an isomorphism \b \mathcal H^{(0)} \simeq \{ \,\textnormal{closed 2-forms in }\, H^{(0)} \, \}.\notag \end{equation} \iffalse From Eqs.\eqref{eq:strt2}, \eqref{1dPsi}, the equation ${\rm d}\Phi \equiv0\mod\theta_0,\theta_1,\ol{\xi},\ol{\eta}_2$ gives $$F_{1,\bar{1}}=0.$$ The equations $({\rm d}\Phi){\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_0\equiv 0\mod \theta_1,\ol{\eta}_2, \mod \ol{\theta}_1,\eta_2$, and $\mod\eta_2,\ol{\eta}_2$ imply $$\sigma=\textnormal{i}( A^{\xi}\xi-A^{\ol{\xi}}\ol{\xi}+A^{1}\theta_1-A^{\bar{1}}\ol{\theta}_1+A^2\eta_2-A^{\bar{2}}\ol{\eta}_2).$$ With the given $\sigma,$ the equations $({\rm d}\Phi){\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1\equiv 0\mod \xi,\ol{\eta}_2,$ and $\mod \ol{\xi},\eta_2$ give $$B_{1,2}=\frac{\textnormal{i}}{3\gamma^2}\left(A^{\ol{\xi},2}+A^{2,\ol{\xi}}\right),\quad C_{\bar{1},\bar{2}}=-\frac{\textnormal{i}}{3\gamma^2}\left(A^{\xi,\bar{2}}+A^{\bar{2},\xi}\right). $$ \fi \subsubsection{Structure equation} The equation for the classical differentiated conservation laws is now reduced to $${\rm d}\Phi=0,$$ for $\Phi\in H^{(0)}$. From this, a direct differential analysis yields the following closed structure equation.\footnotemark\footnotetext{This part of the analysis is more or less mechanical, and shall be omitted. \begin{equation}\label{eq:classicalJacobi} {\rm d} \begin{pmatrix} A\\ A^{\xi}\\A^{\ol{\xi}} \\ A^1\\A^{\bar{1}}\\A^{1,1}\\A^{\bar{1}, \bar{1}}\\P \end{pmatrix} = \textnormal{i} \begin{pmatrix} \cdot \\-A^{\xi}\\A^{\ol{\xi}}\\A^1\\-A^{\bar{1}} \\ 2A^{1,1}\\-2A^{\bar{1}, \bar{1}}\\ \cdot \end{pmatrix}\rho +\begin{pmatrix} A^{\xi}& A^{\ol{\xi}}&\cdot& A^1& A^{\bar{1}}&\cdot&\cdot\\ A^{\bar{1}, \bar{1}}&-\frac{3}{2}\gamma^2A & -3\gamma^2 A^{\bar{1}} & P& \cdot& A^{1}&\cdot \\ -\frac{3}{2}\gamma^2A & A^{1,1}& -3\gamma^2 A^1 &\cdot & -P & \cdot & A^{\bar{1}}\\ P&\cdot& 3\gamma^2 A^{\ol{\xi}} &A^{{1,1}}&-\frac{3}{2}\gamma^2A& \cdot &-A^{\xi} \\ \cdot& -P & 3\gamma^2 A^{\xi} &-\frac{3}{2}\gamma^2A& A^{{\bar{1},\bar{1}}}& -A^{\ol{\xi}} &\cdot \\ -\gamma^2 A^{\ol{\xi}}& \cdot&\cdot&\cdot&-\gamma^2 A^1 & \cdot &-2P \\ \cdot&-\gamma^2 A^{\xi}&\cdot&-\gamma^2 A^{\bar{1}} &\cdot& 2P & \cdot \\ \frac{\gamma^2}{2}A^{\bar{1}} &-\frac{\gamma^2}{2}A^{1}&\cdot &\frac{\gamma^2}{2}A^{\ol{\xi}}&-\frac{\gamma^2}{2}A^{\xi} & A^{1,1} & -A^{\bar{1},\bar{1}} \end{pmatrix} \begin{pmatrix} \xi \\ \ol{\xi} \\ \theta_0\\\theta_1\\ \ol{\theta}_1 \\ \eta_2 \\ \ol{\eta}_2 \end{pmatrix}. \end{equation} \noindent Here the upper-script $A^{*,*}$ denotes the covariant derivative as before. One finds that the closed linear differential system for $\{A,A^{\xi},A^{\ol{\xi}},A^1,A^{\bar{1}}, A^{1,1}, A^{\bar{1},\bar{1}}, P\}$ is compatible, i.e., ${\rm d}^2=0$ is a formal identity. By the existence and uniqueness theorem of ODE, the space of (local) classical differentiated conservation laws is eight dimensional. Note that, at this stage, the reduced 2-form $\Phi$ is normalized to \begin{equation}\label{eq:H0AA} \Phi=A \Psi +\textnormal{i} \theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\left(A^{\xi}\xi-A^{\ol{\xi}}\ol{\xi}+A^{1}\theta_1-A^{\bar{1}}\ol{\theta}_1\right). \end{equation} \subsection{Noether's theorem} Observe that the coefficients of $\Phi$ are determined by the coefficient $A$ (or $P$) and its successive derivatives. Moreover, Eq.\eqref{eq:classicalJacobi} shows that $$\begin{cases} &\textnormal{$A$ is a classical Jacobi field},\\ &\textnormal{$P$ is a classical pseudo-Jacobi field.} \end{cases}$$ It follows from this that we have the isomorphisms $$ \mathcal H^{(0)}\simeq \mathfrak{J}^{(0)}\simeq \mathfrak{J}'^{(0)}\simeq \liealgebra{g}^{\mathbb C}.$$ In particular, this implies that the classical differentiated conservation laws are globally defined on $X$. \begin{cor}[\textbf{Noether's theorem} for classical conservation laws]\label{cor:Noether} Let $M$ be the simply connected $2_{\mathbb C}$-dimensional complex space form of constant holomorphic sectional curvature $4\gamma^2.$ Let $G=\liegroup{SU}(3),\,\textnormal{or}\; \liegroup{SU}(1,2)$ be the group of K\"ahler isometries of $M$. Let $\liealgebra{g}$ be its Lie algebra. Then there exist the isomorphisms $$ (\mathfrak{S}^{(0)})^{\mathbb C}\simeq \mathfrak{J}^{(0)}\simeq \mathfrak{J}'^{(0)}\simeq\mathcal H^{(0)}\simeq\liealgebra{g}^{\mathbb C}\simeq\mathcal C^{(0)}_{loc}.$$ Here $\mathcal C^{(0)}_{loc}$ denotes the space of local conservation laws of $\mathcal I$ restricted to a small contractible open subset of $X$. \end{cor} \begin{proof} For a classical Jacobi field $A$, set $$\varphi_A=\frac{\textnormal{i}}{6\gamma^2}\left( A^{\xi}\ol{\theta}_1-A^{\ol{\xi}}\theta_1+A^{\bar{1}}\xi-A^1\ol{\xi} \right). $$ Then $ {\rm d}\varphi_A=\Phi$, where $\Phi$ is given by \eqref{eq:H0AA}. This establishes the isomorphism $\mathcal H^{(0)}\simeq\mathcal C^{(0)}_{loc}.$ \end{proof} For the global consideration of conservation laws including the contribution from the cohomology of the ambient space $X$, we refer to \cite[p.584, Theorem 1]{Bryant1995}. \begin{rem}[Moment conditions] Consider a closed integral curve $\Gamma\subset X$ of $\mathcal I$. Such $\Gamma$ corresponds to a pair $(\gamma, \Pi)$ of a closed curve $\gamma\subset M$ equipped with a field of oriented Lagrangian 2-planes $\Pi$ tangent to $\gamma$ such that the associated $(2,0)$-vector field is covariant constant. The integrals of the classical conservation laws on $\Gamma$ yield the moment conditions for $(\gamma, \Pi)$ to bound a minimal Lagrangian surface, with the given boundary $\gamma$ and the prescribed tangent planes $\Pi$ along $\gamma$. Compare \cite{Fu1995}. \end{rem} \iffals \begin{rem} We remark that a classical symmetry $V$ satisfies the stronger relations \b \mathcal L_V\theta_0=0,\;\mathcal L_V\Psi=0.\notag \end{equation} Define the \textbf{Poincar\'e-Cartan form} \b \Upsilon=\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\Psi\in F^2\Omega^3(X).\notag \end{equation} Then since ${\rm d} \Upsilon=0$, we have $\mathcal L_{V_A}\Upsilon=0={\rm d} (V_A \lrcorner \Upsilon).$ It follows that the map \b A\mapsto V_A\lrcorner \Upsilon_0.\notag \end{equation} directly establishes the isomorphism between the classical Jacobi fields and the classical differentiated conservation laws. \end{rem} \f \iffalse The classical conservation laws were used by Kusner \cite{Kusner1991} to define certain global invariants for minimal Lagrangian surfaces that became important for the structure theorems of properly embedded minimal Lagrangian surfaces in 3-dimensional space forms. Let us give a discussion of some of the main results in \cite{Korevaar1989}, \cite{Korevaar1992} adapted to our framework. Let $\liegroup{G}$ be the six dimensional group of isometries of $M$. Let $\liealgebra{g}$ be the Lie algebra of $\liegroup{G}$. We identify an element $Y \in \liealgebra{g}$ with the corresponding Killing vector field on $M$. It is a standard fact that the divergence of a Killing field of any Riemannian manifold vanishes, \begin{equation}\label{eq:Ydivergence} \nabla \cdot Y=0. \end{equation} Since $H^2(M, \mathbb R)=0$, for a given Killing field $Y$ there exists a vector field $Z^Y\in H^0(TM)$ such that \begin{equation}\label{eq:Ycurl} Y=\nabla\times Z^Y \quad\textnormal{(curl)}. \end{equation} Such a vector field $Z^Y$ is not unique, but by choosing a basis of $\liealgebra{g}$ let us assume we have an injective homomorphism from $\liealgebra{g}$ to $H^0(TM)$ that maps $Y\to Z^Y$. Let $\textnormal{x}:\Sigma\hookrightarrow M$ be a minimal Lagrangian-$\delta$ surface. Let $e_3\in H^0(\textnormal{x}^*TM)$ be the unit normal vector field to $\Sigma$. For a Killing field $Y$ of the ambient space $M$, let $Y=Y^T+Y^N$ be the orthogonal decomposition into the tangential and normal components along $\Sigma$. A computation shows that \[ \nabla^{\Sigma}\cdot Y^T=2\delta \,\langle e_3,Y \rangle. \] Here $\nabla^{\Sigma}\cdot Y^T$ is the divergence of $Y^T$ as a vector field on $\Sigma$. Let $\Gamma\hookrightarrow\Sigma$ be an oriented closed curve representing a homology class $[\Gamma] \in H_1(\Sigma,\mathbb Z)$. Let $e_1$ be the tangent vector field of $\Gamma$, and $e_2$ be the unit conormal field along $\Gamma$ such that the frame $\{e_1, e_2, e_3\}$ is positively oriented. Let $D \hookrightarrow M$ be a surface (2-simplex) such that ${\partial} D = \Gamma$. Let $\mathfrak{n}$ the oriented unit normal field to $D$ which is in the direction of $e_2$ along $\Gamma$. Kusner defined the \textbf{momentum class} \[ \mu \in H^1(\Sigma,\mathbb R) \otimes \liealgebra{g}^* \] by the integral formula \begin{equation}\label{eq:momentumclass} \langle \mu([\Gamma]), Y \rangle= \int_{\Gamma} \langle e_2,Y \rangle + 2\delta \int_{D} \langle \mathfrak{n},Y \rangle. \end{equation} Under the action of $g\in\liegroup{G}$ on $\textnormal{x}$, the momentum class $\mu$ transforms as \[ g^*(\mu)=\Ad^*(g) \cdot \mu. \] In order to see momentum class is well defined, let $(\Gamma', D')$ be another set of data representing the given homology class $[\Gamma]$. There exists a 3-simplex $U\subset M$ and a subset $S\subset\Sigma$ such that \[\qquad\quad {\partial} U = D-D'+S,\quad\; {\partial} S=\Gamma-\Gamma'\quad\textnormal{(outward normal)}. \] Applying the divergence theorem to $Y$ and $Y^T$ in turn, we find that \begin{align}\label{eq:divergence} 2\delta \int_{D} \langle \mathfrak{n},Y \rangle -2\delta \int_{D'} \langle \mathfrak{n'},Y \rangle &=-\int_S 2\delta \langle e_3,Y \rangle,\\ &=- \int_{\Gamma} \langle e_2,Y \rangle + \int_{\Gamma'} \langle e_2',Y \rangle, \notag \end{align} and the claim follows. \vspace{2mm} We wish to show that the momentum class is equivalent to the classical conservation laws. Note from the identity \eqref{eq:Ycurl} and Stokes' theorem that the momentum class can be expressed entirely in terms of line integrals \[ \langle \mu([\Gamma]), Y \rangle= \int_{\Gamma} \langle e_2,Y \rangle + 2\delta \int_{\Gamma} \langle e_1, Z^Y \rangle. \] Analytically this can be interpreted as follows. From the structure equation \eqref{eq:classicalJacobi} we have the identity \begin{equation}\label{eq:classicalcvlawidentity} {\rm d} \Big(\textnormal{i}(A^{\bar{1}}\xi-A^1\ol{\xi})\Big)=\Phi_A+\delta\textnormal{i} \Big(A\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}-2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}(A^{\bar{1}}\xi-A^1\ol{\xi})\Big). \end{equation} Here $\Phi_A$ denotes the differentiated conservation law \eqref{eq:H0} generated by a Jacobi field $A$. One may check that $\Xi_A=\textnormal{i} \Big(A\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}-2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}(A^{\bar{1}}\xi-A^1\ol{\xi})\Big)$ is a closed 2-form defined on the ambient space form $M$ (this is equivalent to \eqref{eq:Ydivergence}). Since $H^2(M,\mathbb R)=0$, there exists a 1-form $\chi_A$ on $M$ such that ${\rm d}\chi_A=\Xi_A$. Consequently we have \[ {\rm d} \Big(\textnormal{i}(A^{\bar{1}}\xi-A^1\ol{\xi})-\delta\chi_A\Big)=\Phi_A. \] By the defining properties of the contact 1-form $\theta_0$, Eq.\eqref{eq:divergence} is recovered by first integrating $\Xi_A$ on the 2-cycle ${\partial} U$, and then integrating \eqref{eq:classicalcvlawidentity} on the 2-simplex $S$. As a result of the momentum class analysis, we conclude that for a classical differentiated conservation law there exists an un-differentiated representative which is globally defined on $X$. In particular, Kusner's momentum class accounts for all the classical conservation laws. It is known that there exists an infinite sequence of non-classical higher-order symmetries and conservation laws for the elliptic $\sinh$-Gordon equation which locally describes the minimal Lagrangian surfaces for the case $\gamma^2>0$, \cite{Fox2011}. We shall show that Noether's theorem can be extended to the higher-order symmetries and conservation laws for minimal Lagrangian surfaces. See \S\ref{sec:highercvlaws1}, \ref{sec:recursion}. In \S\ref{sec:families} we will return to the adjoint action of the isometry group G on $\mu$ and its higher-order counterparts, and formulate these in terms of the action of symmetry vector fields on the characteristic cohomology. We put it off so that we may deal with the classical and higher-order conservation laws at the same time. \fi \subsection{Higher-order extension} In the remainder of the paper, the analysis of the classical objects in this section will be extended to their higher-order analogues in the setting of the infinite prolongation of the differential system $(X,\mathcal I)$. \vspace{2mm} The extension process relies on the familiar extension of the Maurer-Cartan form $$\psi \longrightarrow \psi_{\lambda}=\lambda \psi_{+}+\psi_0+\lambda^{-1}\psi_{-},$$ obtained by inserting the spectral parameter $\lambda$. This leads to the generalization of the $\liealgebra{g}$-valued classical Killing field $\textbf{X}$ to the loop algebra $\liealgebra{g}^{\mathbb C}[[\lambda]]$-valued formal Killing field $\textbf{X}_{\lambda},$ $$\textbf{X}\longrightarrow \textbf{X}_{\lambda}.$$ In practice, this amounts to \emph{expanding} the scalar coefficients $\{ \textbf{p}, \textbf{b}, \textbf{c}, \textbf{f}, \textbf{a}, \textbf{g}, \textbf{s}, \textbf{t} \}$ of $\textbf{X}$ to the appropriate formal series in $\lambda$. We will find that this extension is also adapted so that the structure equation for the formal Killing field contains the higher-order version of the recursion relations discussed in \S\ref{sec:recursionrel}. We will be able to read off the higher-order (pseudo) Jacobi fields and conservation laws from the structure equation for $\textbf{X}_{\lambda}$. \vspace{2mm} As a first step to the higher-order analysis, we begin by introducing the infinite prolongation of the differential system $(X,\mathcal I)$, and record the basic structure equations. \section{Prolongation}\label{sec:prolongation1} The minimal Lagrangian surfaces under consideration are locally described by the elliptic Tzitzeica equation, which is a well known example of an integrable elliptic equation. It possesses an infinite sequence of higher-order symmetries and conservation laws, \cite{Fox2012}. In order to access the corresponding structures for the minimal Lagrangian system, it becomes necessary to introduce the infinite prolongation. \vspace{2mm} In \S\ref{sec:prolonginf}, we determine the basic structure equations for the infinite prolongation. The commutation relations for the dual frame of vector fields will be used in the next sections for the classification of Jacobi fields. In \S\ref{sec:doublecover}, we define the triple cover of the infinite prolongation to support the triple cover of a minimal Lagrangian surface defined by Hopf differential. In \S\ref{sec:balanced}, we introduce a sequence of adapted functions called \emph{balanced coordinates} and record their basic structure equations. \vspace{2mm} In hindsight, the hidden symmetries of the minimal Lagrangian system begin to emerge in the form of a weighted homogeneity with respect to the balanced coordinates. To be precise, this is a non-local symmetry associated with an integrable extension. We do not pursue to formulate this properly in this paper, and the notion of weighted homogeneity would suffice for our purposes. We mention that this is a part of an infinite hierarchy of non-local symmetries called spectral symmetry. \subsection{Infinite prolongation}\label{sec:prolonginf} \subsubsection{Infinite sequence of $\mathbb P^1$-bundles}\label{sec:infP1} Set $(X^{(0)},\,\I{0})=(X,\,\mathcal I)$ be the original differential system \eqref{1ideal2}. Inductively define $(\X{k+1},\,\I{k+1})$ as the prolongation of $(\X{k},\,\I{k})$ such that $$\pi_{k+1, k}: \X{k+1}\to \X{k}$$ is the bundle of oriented integral 2-planes of $(\X{k},\,\I{k})$, and that the differential ideal $\I{k+1}$ is generated by $\pi_{k+1,k}^*\I{k}$ and the restriction of the canonical contact ideal to $\X{k+1}\subset\Gr^+(2,T\X{k})$. For each $k\geq 0$, the prolongation space $\X{k+1}$ is a smooth manifold. The projection $\pi_{k+1,k}$ is a smooth submersion with two dimensional fibers isomorphic to $\mathbb C\mathbb{P}^1$, see \S\ref{sec:doublecover} for further details. \begin{defn} The \textbf{infinite prolongation} $(X^{(\infty)},\I{\infty})$ of the differential system $(X,\mathcal I)$ for minimal Lagrangian surfaces is defined as the projective limit \begin{align X^{(\infty)} &=\lim_{\longleftarrow} \X{k}, \notag \\ \I{\infty}&=\cup_{k\geq 0} \I{k}. \notag \end{align} Let $\pi_{\infty,k}:X^{(\infty)}\to\X{k}$ be the associated projection. Here we identity $\I{k}$ with its image $\pi_{\infty,k}^*\I{k}\subset\Omega^*(X^{(\infty)}).$ By construction, the sequence of Pfaffian systems $\I{k}$ satisfy the inductive closure conditions $${\rm d}\I{k}\equiv 0\mod \I{k+1}, \; k\geq 1.$$ \end{defn} \subsubsection{Associated principal $\liegroup{SO}(2)$-bundles}\label{eq:associatedbundle} For the purpose of computation, we also introduce the following associated bundles. Consider the principal $\liegroup{SO}(2)$-bundle $\Pi:\mathcal F\to X=\mathcal F/\liegroup{SO}(2)$. Set $\F{k}=\Pi^*\X{k}$ for $k\geq 1$, and define \[\F{\infty}= \lim_{\longleftarrow} \,\F{k}.\] The higher-order differential analysis will be practically carried out on the principal $\liegroup{SO}(2)$-bundle $$ \F{\infty}\to\X{\infty}$$ in an $\liegroup{SO}(2)$-equivariant manner so that it has a well defined meaning on $\X{\infty}$. For simplicity, we continue to use $\I{k}, {\rm I}^{(\infty)}$ to denote the corresponding differential ideals on $\F{k}, \F{\infty}.$ \subsubsection{Zariski open sets}\label{sec:opensets} Let $\X{1}_0\subset\X{1}$ be the open subset defined by the independence condition \begin{equation}\label{eq:X10} \X{1}_0=\{ \;\, \textnormal{E}\in \X{1} \;\vert\; \; (\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi})_{\vert_{\textnormal{E}}}\ne 0\,\}. \end{equation} For $k\geq 1$, inductively define the corresponding sequence of open subsets $$\X{k+1}_0=\pi_{k+1,k}^{-1}(\X{k}_0).$$ Let $$\lim_{\longleftarrow} \X{k}_0=X^{(\infty)}_0\subsetX^{(\infty)}$$ denote their projective limit. Let $\F{k}_0\subset\F{k}, \F{\infty}_0\subset\F{\infty}$ denote the associated open subsets. Dually, consider the complementary open subset \begin{equation}\label{eq:X1infty} \X{1}_{\infty}=\{ \;\, \textnormal{E}\in \X{1} \;\vert\; \; (\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\eta}_2)_{\vert_{\textnormal{E}}}\ne 0\,\}. \end{equation} Its associated open subsets $\X{k}_{\infty}, X^{(\infty)}_{\infty}, \F{k}_{\infty}, \F{\infty}_{\infty}$ are defined similarly as above. Note that the pair $\{ X^{(\infty)}_{0}, X^{(\infty)}_{\infty}\}$ form an open covering of $X^{(\infty)}$ (and similarly for $\{ \F{\infty}_{0}, \F{\infty}_{\infty}\}$ and $\F{\infty}$). \vspace{2mm} The analysis of the infinitely prolonged differential system for minimal Lagrangian surfaces will be carried out on $\F{\infty}_0$, which is one half of the open covering of $\F{\infty}.$ But we remark that it can be equally well carried out on the other half of the open covering $\F{\infty}_{\infty}$. Since the formulation of the structure equation on $\F{\infty}_{\infty}$ is almost identical to that on $\F{\infty}_0$, we shall present only the $\F{\infty}_{0}$-part of the analysis and omit the $\F{\infty}_{\infty}$-part. Let us give instead an indication on how the dual prolongation process on $\F{1}_{\infty}$ would proceed. Given the 2-form $\eta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi$ in the ideal $\mathcal I$, recall the prolongation variable $h_3$ defined by \eqref{eq:etah3}. Introduce the new prolongation variable $p_3$ such that $$\xi=p_3\eta_2$$ on the integral 2-planes in $\F{1}_{\infty}$ (and hence $p_3=h_3^{-1}$ on $\F{1}_{0}\cap\F{1}_{\infty}$). By switching from $h_3$ to $p_3$, it is clear that one may proceed analogously to the infinite prolongation on $\F{\infty}_{\infty}$, such that the associated formulas agree on the intersection $\F{\infty}_{0}\cap\F{\infty}_{\infty}.$ \subsubsection{Infinitely prolonged structure equation}\label{sec:infstrt} Recall the induced structure equation \eqref{eq:strtx} on an immersed minimal Lagrangian surface. By the standard prolongation process of the exterior differential system theory, \cite[Chapter \rm{VI}]{Bryant1991}, set $$\eta_2=h_3\xi$$ for a prolongation variable $h_3.$ From the third equation of \eqref {eq:strtx}, we get $${\rm d} h_3+3\textnormal{i} h_3\rho\equiv 0\mod\xi.$$ Inductively define the higher-order derivatives of $h_3$ by the equations \begin{equation}\label{eq:dhj} {\rm d} h_j + \textnormal{i} j h_j \rho = h_{j+1} \xi + T_{j} \ol{\xi}, \; \; j = 3, \, 4, \, ... \, , \end{equation} where \begin{align}\label{eq:Tj} T_{3}&=0, \, \\ T_{j+1}&= \sum_{s=0}^{j-3} a_{j,s} \, h_{j-s} \, \partial^{s}_{\xi} R, \; \; \; \mbox{for} \; \; j\geq 3, \notag \\ & \quad \;\;a_{j,s} =\frac{(j+2s+3) }{2(j-1)} {j-1 \choose s+2}, \notag\\ & \quad \partial^{s}_{\xi} R=\delta_{0s}\gamma^2 - 2 h_{3+s} \bar h_3.\notag \end{align} The formula for the sequence $\{ \,T_{j+1}\,\}$ is uniquely determined by requiring that ${\rm d}({\rm d} h_j)=0$ for $j=3,\,4,\,...\,$. This implies the recursive relation \begin{equation} \label{eq:Trecurs} T_{j+1} = \partial_{\xi} T_j + \frac{j}{2}R\, h_j. \end{equation} For example, $$T_3=0, \; T_4=\frac{3}{2}h_3(\gamma^2-2h_3\bar h_3),\; T_5=\frac{7}{2}\gamma^2h_4 -10h_3\bar h_3h_4.$$ \begin{rem} Similarly as in \eqref{eq:delxnotation}, we use the subscript notation $f_{\xi}$ or $\,{\partial}_{\xi}f$ to denote the $\xi$-coefficient of ${\rm d} f$ (and similarly for $f_{\ol{\xi}}$ or $ {\partial}_{\ol{\xi}}f$). \end{rem} Based on Eqs.\eqref{eq:dhj}, \eqref{eq:Tj}, we define the following set of differential forms and vector fields on $\F{\infty}_0$: \vspace{2mm} \begin{enumerate}[\qquad a)] \item Differential forms: \begin{equation}\label{eq:thetaj} \begin{array}{rlllll} \eta_j&={\rm d} h_j&+\textnormal{i} j h_j \rho&&-T_j \ol{\xi},\,\quad&\textnormal{for}\;j\geq 3,\\ \theta_j&=\eta_j &&-h_{j+1}\xi, &&\textnormal{for}\; j\geq 2. \\ \end{array} \end{equation} On the open subset $\,\F{k}_0$, the set of 1-forms $$\{ \rho; \, \xi, \ol{\xi}, \theta_0, \theta_1 , \ol{\theta}_{1}, \ldots, \,\theta_{k+1}, \ol{\theta}_{k+1}; \,\eta_{k+2}, \ol{\eta}_{k+2}\}$$ form a coframe. On the open subset $\,\F{\infty}_0$, the set of 1-forms \begin{equation}\label{Finfframe} \{ \rho; \, \xi, \ol{\xi}, \theta_0, \theta_1 , \ol{\theta}_{1}, \,\theta_2, \ol{\theta}_{2}, \ldots \} \end{equation} form a coframe. By construction, $${\rm I}^{(\infty)}=\langle \theta_0, \theta_1 , \ol{\theta}_{1}, \,\theta_2, \ol{\theta}_{2}, \ldots \rangle.$$ \vspace{1mm} \item Vector fields: \begin{align} \partial_{\xi}&:= \textnormal{total derivative with respect to}\;\xi\mod{\rm I}^{(\infty)},\notag\\ \partial_{\ol{\xi}}&:= \textnormal{total derivative with respect to}\;\ol{\xi}\mod{\rm I}^{(\infty)}.\notag \end{align} On the open subset $\,\F{\infty}_0$, let \begin{equation}\label{Finfframev} \{ E_{\rho}; \, \partial_{\xi}=E_{\xi}, \partial_{\ol{\xi}}=E_{\ol{\xi}}, E_0, E_1 , E_{\bar{1}}, E_2,E_{\bar{2}}, \ldots \} \end{equation} denote the dual frame of \eqref{Finfframe}. \end{enumerate} \vspace{2mm} In terms of the 1-forms introduced above, we denote the covariant derivatives of $T_{j}$ by \b {\rm d} T_{j} +\textnormal{i} \, (j-1) T_{j} \rho=({\partial}_{\xi} T_{j}) \xi + ({\partial}_{\ol{\xi}} T_{j}) \ol{\xi} +T_{j}^{\bar{3}}\ol{\theta}_{3} +\sum_{s=3}^{j-1}T_{j}^s\theta_s,\quad \,j\geq 3,\notag \end{equation} where $T_j^s=E_s(T_j)$. Recall from \eqref{eq:strt2}, \begin{align}\label{eq:strt22} {\rm d} \xi &= \textnormal{i} \rho {\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \xi -3\gamma^2 \theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1-\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_2-\bar h_3\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}, \\ {\rm d} \theta_0&=-\frac{1}{2}\left( \theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi} \right), \notag\\ {\rm d} \theta_1 +\textnormal{i}\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_1&=-\theta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+3\gamma^2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi},\notag \\ {\rm d} \theta_2+2\textnormal{i}\rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_2 &=-\theta_3{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi +3h_3\gamma^2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1 +h_3\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_2+(\gamma^2+h_3\bar h_3)\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}, \notag \\ {\rm d} \rho&=\frac{\textnormal{i}}{2}\left(R\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}-2\theta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_2-\gamma^2\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1 -2\bar h_3\theta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}+2h_3\ol{\theta}_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi \right). \notag \end{align} Extending this, we record the structure equations satisfied by the 1-forms $\theta_j$ on $\,\mathcal F^{(\infty)}_0$. \begin{lem}\label{lem:prolongedstrt} For $\,j\geq 3$, \begin{align}\label{eq:strt2inf} {\rm d} \theta_j + \textnormal{i} j \rho{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_j&=-\theta_{j+1}{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi +3\gamma^2 \theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}(T_j\theta_1+h_{j+1}\ol{\theta}_1) +\frac{jh_j}{2}\left( \gamma^2 \theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1+2\theta_2{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_2 \right) \\ &\quad+T_j\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\theta_2+h_{j+1}\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_2 +\tau_j'{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+\tau_j''{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}, \quad\textnormal{where} \notag \\ &\quad\tau_j'= h_3T_j \ol{\theta}_1- j h_3h_j \ol{\theta}_2,\notag\\ &\quad\tau_j''= \bar h_3 h_{j+1}\theta_1+j \bar h_3h_j \theta_2 -(\,T_{j}^{\bar{3}}\ol{\theta}_{3}+\sum_{s=3}^{j-1}\,T_{j}^s\theta_s).\notag \end{align} \end{lem} \begin{proof} Direct calculation. We omit the details. \end{proof} Eqs.\eqref{eq:strt22}, \eqref{eq:strt2inf} imply the following commutation relation for the dual frame. \begin{cor}\label{lem:commutation} The dual frame \eqref{Finfframev} satisfies the following commutation relations. \begin{equation}\label{eq:lxib} [E_{\ell},\,E_{\ol{\xi}} ] =\sum_{j\geq {\ell}+1}T^{\ell}_j E_j, \qquad \ell\geq 3. \end{equation} \end{cor} \begin{proof} Let $\theta$ be a differential 1-form, and let $E_a,\,E_b$ be vector fields. The corollary follows from Cartan's formula \[ {\rm d}\theta(E_a,\,E_b)=E_a(\theta(E_b))-E_b(\theta(E_a))-\theta([E_a,\,E_b]). \] \end{proof} We introduce the following notations in order to keep track of orders ($k\geq 2$): \begin{itemize} \item $\mathcal O(k)$: functions on $X^{(\infty)}$ that do not depend on $h_j$ for $j>k$. \item $\mathcal O(-k)$: functions on $X^{(\infty)}$ that do not depend on $\bar h_j$ for $j>k$. \end{itemize} \noindent Note for example that $$ T_j\in\mathcal O(j-1). $$ \noindent \subsection{Triple cover}\label{sec:doublecover} \subsubsection{Set up}\label{sec:triplemoti} The triple cover $\hat{\Sigma}\to\Sigma$ of an immersed integral surface of $(X,\mathcal I)$ defined in Defn.~ \ref{defn:doublecover} prompts the definition of a global triple cover $\hat{X}^{(\infty)}\toX^{(\infty)}$ such that the following commutative diagram holds: \vspace{1mm}\\ \centerline{\xymatrix{\hat{\Sigma} \ar[d] \; \ar@{^{}.>}[r] &\;\;\hat{X}^{(\infty)} \ar[d] \\ \Sigma \; \ar[r] &\;\;X^{(\infty)} }} \\ \vspace{2mm}\noindent It turns out that the triple cover $\hat{X}^{(\infty)}$ to be constructed also supports the splitting field for the (pseudo) Jacobi equation for the minimal Lagrangian system, see \S\ref{sec:Jacobifields}. \subsubsection{Definition}\label{sec:tripledefi} Let $$K^{\mathcal F} \to \mathcal F, \quad K^{\mathcal F}_{\eta_2} \to \mathcal F,$$ be the (trivial) complex line bundles generated by the 1-forms $\xi, \eta_2$ respectively. Recall the principle $\liegroup{SO}(2)$-bundle $\Pi:\mathcal F \to X$, and the structure equation \eqref{eq:strt2}. Let $$K\to X, \quad K_{\eta_2} \to X,$$ be the induced line bundles respectively. Note from Eq.\eqref{eq:strt2} that $$ K_{\eta_2} =K^{-2}.$$ Consider $\mathcal F^{(1)}=\Pi^*(\X{1})$. Since an element in $\mathcal F^{(1)}$ is defined by the equation $$\eta_2-h_3\xi=0$$ for a coefficient $h_3$, we have $$\mathcal F^{(1)}\cong {\mathbb{P}}( K^{\mathcal F} \oplus K^{\mathcal F}_{\eta_2}) \cong \mathcal F \times \cp{1}.$$ It follows that $$ \X{1}\cong {\mathbb{P}}(K^{p+1} \oplus K^{p-2})$$ for any integer $p$. In order to define the triple cover, we choose $\X{1} \cong {\mathbb{P}}(K^{3} \oplus \mathbb C)$. \begin{defn} The \textbf{triple cover} $\Xh{1}$ of the first prolongation $\X{1}$ is defined by \[ \Xh{1}= {\mathbb{P}}(K \oplus \mathbb C). \] Define similarly $ \Fh{1}= {\mathbb{P}}( K^{\mathcal F} \oplus \mathbb C).$ \end{defn} Denote the triple covering map \begin{equation} \textbf{c}:\Fh{1} \to \F{1}, \notag \end{equation} which is the standard branched triple cover when restricted to the fibers; \begin{align} \cp{1} &\to \cp{1},\notag\\ [x,y] & \mapsto [x^3,y^3].\notag \end{align} It is branched at the two points $0 = [0,1]$ and $\infty = [1,0]$. By construction, $\textbf{c}$ descends to the triple covering map $$ \textbf{c}:\Xh{1} \to \X{1}. $$ Let $\Ih{1}=\textbf{c}^*\I{1}$ be the pulled back ideal, and define the new differential system $$(\Xh{1},\,\Ih{1}).$$ It is clear that the present construction has the naturality so that the triple cover of a minimal Lagrangian surface $\hat{\Sigma}\to\Sigma$ admits a unique lift to $\Xh{1}$ as an integral surface of $\Ih{1}$. \begin{defn} For each $2\leq k \leq \infty$, define the \textbf{triple cover} of the prolongations \begin{align*} \Xh{k}&:=\textbf{c}^*(\X{k}), \\ \Fh{k}&:= \textbf{c}^*(\F{k}), \end{align*} such that we have the commutative diagram:\vspace{1mm} \\ \centerline{ \xymatrix{ \Xh{k} \ar[d]^{\hat{\pi}_{k,1}} \ar[r]^{\textbf{c}_k} & \X{k} \ar[d]^{\pi_{k,1}} \\ \Xh{1} \ar[r]^{\textbf{c}} & \X{1} }}\\ \vspace{1mm}\noindent (and similarly for $\Fh{k}, \Fh{1}$, etc). Set the ideals $\Ih{k}:=\textbf{c}_k^*(\I{k})$. Denote the pulled-back canonical bundles by \begin{align*} \pi_{k,0}^*K&:=K\to \X{k},\\ \textbf{c}_k^*K&:=\hat{K} \to \Xh{k}.\notag \end{align*} \end{defn} The following proposition summarizes the construction so far. \begin{prop} Let $\textnormal{x}:\Sigma\hookrightarrow X$ be an immersed integral surface of the differential system for minimal Lagrangian surfaces. Let $$\textnormal{x}^{(k)}:\Sigma\hookrightarrow\X{k},\,1\leq k \leq\infty,$$ be the prolongation of $\textnormal{x}$. Let $\nu:\hat{\Sigma}\to\Sigma$ be the triple cover defined by the Hopf differential of $\textnormal{x}$, Definition \ref{defn:doublecover}. There exists a lift $\hat{\textnormal{x}}^{(1)}:\hat{\Sigma}\hookrightarrow\Xh{1}$ and the associated sequence of prolongations \[ \hat{\textnormal{x}}^{(k)}:\hat{\Sigma}\hookrightarrow\Xh{k},\quad 2\leq k \leq\infty, \] such that \begin{enumerate}[\qquad a)] \item each $\hat{\textnormal{x}}^{(k)}$ is integral to $\Ih{k}$, \item $\textnormal{x}^{(k)}\circ \nu =\textbf{c}_k\circ\hat{\textnormal{x}}^{(k)}$. \end{enumerate} The lift $\,\hat{\textnormal{x}}^{(1)}$ and its prolongation sequence $\{\,\hat{\textnormal{x}}^{(k)}\,\}$ are uniquely determined by these properties. \end{prop} \subsection{Balanced coordinates}\label{sec:balanced} As is often the case with integrable differential equations, the infinite prolongation of the minimal Lagrangian system supports a preferred ring of functions called \emph{balanced coordinates}.\footnotemark \footnotetext{They form a coordinate system when restricted to a fiber of the projection $\hat{X}^{(\infty)}_{**}\to\Xh{1}.$} We will find that many of the higher-order objects we are interested in, such as higher-order Jacobi fields and conservation laws, admit the weighted homogeneous expressions in terms of these functions. \subsubsection{Set up}\label{sec:balancedmoti} Recall the open subsets $\X{k}_0, \X{k}_{\infty},$ $1\leq k \leq \infty$, from \eqref{eq:X10} and below. Let $\X{1}_*\subset \X{1}$ be their intersection, \b \X{1}_*:=\X{1}_0\cap\X{1}_{\infty}=\{ \, \textnormal{E}\in\X{1}\;\vert\; \;h_3\vert_E\ne0,\infty\,\}.\notag\end{equation} Inductively define the sequence of open subsets $\X{k+1}_*:=\pi_{k+1,k}^{-1}(\X{k}_*)\subset \X{k+1}.$ Let \b \X{k}_{**}:=\X{k}_*\cap \{ \, h_j \ne \infty,\; 4\leq j \leq k+2\, \}\subset \X{k}_*, \quad k\geq 2.\notag \end{equation} Denote the corresponding open subsets on the triple cover by $$\Xh{k}_0, \Xh{k}_{\infty}, \Xh{k}_*, \Xh{k}_{**}, \;\textnormal{etc}. $$ The balanced coordinates to be constructed will be defined on $\Xh{\infty}_{**}.$ Let $\{\F{k}_0, \F{k}_{\infty}, \F{k}_*, \F{k}_{**}, \Fh{k}_0, \Fh{k}_{\infty}, \Fh{k}_*, \Fh{k}_{**} \}$ denote the associated open sets. \subsubsection{Definition}\label{sec:balanceddefi} With this preparation, we introduce a preferred set of functions on $\hat{X}^{(\infty)}_{**}$ which are adapted for the prolongation structure of the minimal Lagrangian system. \begin{defn}\label{defn:balancedz} A sequence of \textbf{balanced coordinates} $z_j : \Xh{\infty}_{**} \to \mathbb C, j \geq 4,$ are defined by $$ z_j:=h_3^{-\frac{j}{3}}h_j. $$ \end{defn} It is clear that these functions, originally defined on $\Fh{\infty}_{**}$, are invariant under the action of the structure group $\liegroup{SO}(2)$. They are well defined on $\hat{X}^{(\infty)}_{**}$. \iffalse \begin{lem}\label{lem:zj} Suppose $\hat{\textnormal{x}}^{\infty}:\hat{\Sigma} \hookrightarrow \Xh{\infty}$ is a smooth integral surface of the differential ideal $\Ih{\infty}$. Then the induced function $(\hat{\textnormal{x}}^{\infty})^*(z_j)$ on $\hat{\Sigma}$ can be expressed locally as the quotient of a smooth function and a holomorphic function. \end{lem} \begin{proof} Let $\omega=\sqrt[3]{{\rm I \negthinspace I}}$ be the cube root of the Hopf differential for $\hat{\textnormal{x}}^{\infty}$ on $\hat{\Sigma}$. Then \[ (\hat{\textnormal{x}}^{\infty})^*(z_j)= \frac{\textbf{c}^*(h_j\xi^j)}{\omega^j}. \] \end{proof}\fi \subsubsection{Structure equation}\label{sec:balancedstrt} We record the structure equation for the balanced coordinates. Set \begin{equation}\left\{ \begin{array}{rl} \textnormal{r}&:=|h_3|=(h_3\bar h_3)^{\frac{1}{2}}, \notag \\ \omega&:=h_3^{\frac{1}{3}} \xi, \notag \\ \zeta_j&:=h_3^{-\frac{j}{3}}\theta_j,\qquad\quad j\geq 0, \notag \\ \hat{T}_j &:=h_3^{-\frac{j-1}{3}}T_j, \;\;\quad\quad j\geq 3.\notag\\ \end{array}\right. \end{equation} They are all invariant under the action of the structure group $\liegroup{SO}(2)$ and hence well defined on $\Xh{\infty}_{**}$. This enables to express the ideal $\hat{\rm I}^{(\infty)}$ on $\Xh{\infty}_{**}$ by $$\hat{\rm I}^{(\infty)}=\langle \theta_0, \zeta_j, \ol{\zeta}_j \rangle,$$ and the generators are defined on $\Xh{\infty}_{**}$ itself. \vspace{2mm} Note the following structure equations: \begin{align}\label{eq:zetastruct} {\rm d} \textnormal{r}&\equiv\frac{\textnormal{r}}{2}(z_4\omega+\ol{z}_4\ol{\omega}), \\ {\rm d} z_j&\equiv \left(z_{j+1}-\frac{j}{3} z_4 z_j \right) \omega + \hat{T}_{j} \textnormal{r}^{-\frac{2}{3}}\ol{\omega} \;\mod \hat{\rm I}^{(\infty)}, \quad{\rm for} \; j \geq 4,\notag\\ \hat{T}_{j+1}&= {\partial}_{\omega} \hat{T}_{j}+\frac{j-1}{3}z_4 \hat{T}_j +\frac{j}{2}(\gamma^2- 2 \textnormal{r}^2)z_j, \quad{\rm for} \; j \geq 3.\notag \end{align} Here we set $z_3=1$ for convenience, and denote by ${\partial}_{\omega}$ the operator $${\partial}_{\omega}:=h_3^{-\frac{1}{3}}\partial_{\xi}.$$ From these formulas, we note an important property of $\hat{T}_j$. \begin{defn}\label{defn:spectralweights} The \textbf{spectral weights} of the balanced coordinates are \begin{equation}\label{eq:spectralweight} \textnormal{weight}(z_j)=j-3, \quad \textnormal{weight}(\ol{z}_j)=-(j-3). \end{equation} Set $\textnormal{weight}(\textnormal{r})=0.$ Assign the weights for the 1-forms $$ \textnormal{weight}(\omega)=-1, \quad \textnormal{weight}(\overline{\omega})=+1. $$ \end{defn} \begin{lem}\label{lem:Tjh} By definition, \b \hat{T}_j \in \mathbb C[\textnormal{r}^2,z_4,z_5,z_6,\ldots]. \notag \end{equation} It is weighted homogeneous of weight $j-4$ under the spectral weight \eqref{eq:spectralweight}. \end{lem} \begin{proof} From the identities \begin{align} {\partial}_{\omega} z_j&=z_{j+1}-\frac{j}{3} z_4 z_j, \notag\\ {\partial}_{\omega} ( \textnormal{r}^2 )&= \textnormal{r}^2 z_4,\notag \end{align} the operator ${\partial}_{\omega}$ increases the spectral weight by +1 when acting on $\mathbb C[\textnormal{r}^2,z_4,z_5,z_6,\ldots]$. Note the initial term $\hat{T}_4=\frac{3}{2}(\gamma^2-2\textnormal{r}^2)$ is of weight $\,0$. The rest follows from the inductive formula for $\hat{T}_j$ in \eqref{eq:zetastruct}. \end{proof} \iffalse Consider the functions $$ z_5-\frac{5}{3} z_4^2, \quad z_4.$$ One finds that they are weighted homogeneous Jacobi field and pseudo-Jacobi field respectively. Generalizing this, we will find an infinite sequence of (pseudo) Jacobi fields which are weighted homogeneous polynomials in the ring of balanced coordinates $\mathbb C[z_4,z_5,z_6,\ldots].$ \fi \subsection{Order vs. spectral weight} The prolongation variable $h_3$ represents the second fundamental form of the minimal Lagrangian surface, and hence it is a second order object. Consequently, the jet order of the balanced coordinate $z_j$ is $j-1$. However, for the sake of convenience, we re-define the \textbf{order} of the functions $z_j, \ol{z}_j$ as follows: \begin{equation}\label{table:order} \begin{array}{r|r|r} &\textnormal{order}&\textnormal{spectral weight}\\ \hline z_j & j & j-3\\ \ol{z}_j& j &-(j-3) \end{array}\end{equation} \section{Two lemmas}\label{sec:prolonglemmas} We record two useful lemmas regarding the $\partial_{\ol{\xi}}$-equation on $\hat{X}^{(\infty)}$. Among other things, they will be applied to give a classification of the higher-order Jacobi fields in \S\ref{sec:Jacobifields}. Lemma~\ref{lem:lemma5.4'} is a variant of the fact that the minimal Lagrangian system is not Darboux integrable at any order. Lemma~\ref{lem:delbpoly0} is a rigid property of the polynomial ring $\mathbb C[z_4,z_5,z_6, \, ... \, ]$ under the $\partial_{\ol{\xi}}$-operator. \subsection{Lemma \ref{lem:lemma5.4'} } \begin{lem}\label{lem:lemma5.4'} Let $f: U\subset \hat{X}^{(\infty)}\to\mathbb C$ be a scalar function on an open subset $U\subset\hat{X}^{(\infty)}$ such that $$\partial_{\ol{\xi}} f= c h_3^{-\frac{1}{3}}$$ for a constant $c$. From the structure equation, this implies that for some $k>0$, \begin{equation}\label{eq:Lemma5.4'} {\rm d} f \equiv c h_3^{-\frac{1}{3}}\ol{\xi} \mod\; \xi, \theta_0, \theta_1, \ol{\theta}_1, \theta_2,\ol{\theta}_2, \theta_3, \theta_4,\,...\,\theta_{k}, \end{equation} i.e., such $f$ does not depend on any of the conjugate variables $\bar h_j$, $j\geq 3.$ Then $c=0$ necessarily and $f$ is a constant function. \end{lem} \begin{cor}\label{cor:lemma5.4} Let $f: U\subset \hat{X}^{(\infty)}\to\mathbb C$ be a scalar function on an open subset $U\subset\hat{X}^{(\infty)}$ such that \begin{equation}\label{eq:Lemma5.4} \partial_{\ol{\xi}} f =0. \end{equation} Then $f$ is a constant function. \end{cor} The corollary states that the infinitely prolonged minimal Lagrangian differential system is not Darboux integrable, see \cite{Bryant1995b} for a discussion of Darboux integrability (in the hyperbolic case). Roughly, this means that no matter how many times one differentiates, the minimal Lagrangian system under consideration, $\gamma^2\ne 0$, does not admit a Weierstra\ss\, type of holomorphic representation formula. This is in contrast with the case $\gamma^2=0$ where the minimal Lagrangian system is equivalent to the holomorphic differential system for complex curves in $\mathbb C^2.$ \vspace{2mm} A proof of Lem.\ref{lem:lemma5.4'} can be obtained by a direct adaptation of the proof of the corresponding lemma for the differential system for constant mean curvature surfaces given in \cite{Wang2013}. Although the analysis goes by straightforward computations, it is a little involved and, in order to avoid repetition, let us content ourselves with a brief description of the relevant ideas. We shall apply induction on $k$. Assume $k\geq 5$. The case $k\leq 4$ can be checked by a direct computation. Let $\mathtt{I}$ be the Pfaffian system generated by \[ \mathtt{I}=\langle \, \ol{\theta}_2, \ol{\theta}_1,\,\theta_0,\,\xi,\,\theta_1,\,\theta_2,\, ... \,\theta_{k}\,\rangle. \] Our claim is that there is no nonzero closed 1-form $\alpha$ of the form $${\rm d} f=\alpha= c h_3^{-\frac{1}{3}}\ol{\xi} + \boldsymbol{\theta}, \quad \boldsymbol{\theta}\in\mathtt{I}. $$ Denote a 1-form in $\texttt{I}$ by $$\boldsymbol{\theta}=a_{\bar{2}}\ol{\theta}_2+a_{\bar{1}}\ol{\theta}_1+a_0\theta_0 +a_{\xi}\xi+a_1\theta_1+\sum_{j=2}^{k}a_j\theta_j.$$ From the equation (which follows from the identity ${\rm d}^2=0$) $${\rm d}\alpha\equiv 0\mod\mathtt{I},$$ collect $\ol{\theta}_3{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}$-terms and one gets \[ a_{\bar{2}}=-\sum_{j=4}^ka_j T^{\bar{3}}_j. \] Let $\mathtt{I}^{(\bar{1})}$ be the Pfaffian system generated by \[ \mathtt{I}^{(\bar{1})}=\langle \,\overline{\boldsymbol{\theta}}_1, \boldsymbol{\theta}_0,\,\xi,\,\boldsymbol{\theta}_1,\,\boldsymbol{\theta}_2,\, ... \,\boldsymbol{\theta}_{k}\,\rangle, \] where we now set \begin{align} \overline{\boldsymbol{\theta}}_{1}&=\ol{\theta}_{1}, \boldsymbol{\theta}_{0}=\theta_0, \;\boldsymbol{\theta}_1=\theta_1, \;\boldsymbol{\theta}_2=\theta_2, \;\boldsymbol{\theta}_3=\theta_3, \notag\\ \boldsymbol{\theta}_j&=\theta_j-T^{\bar{3}}_j\ol{\theta}_2, \quad \textnormal{for}\;j\geq 4.\notag \end{align} Denote a (new) 1-form in $\mathtt{I}^{(\bar{1})}$ by $$\boldsymbol{\theta}=a_{\bar{1}}\overline{\boldsymbol{\theta}}_{1}+a_0\boldsymbol{\theta}_0 +a_{\xi}\xi+a_1\boldsymbol{\theta}_1+\sum_{j=2}^{k}a_j\boldsymbol{\theta}_j.$$ Now our claim is that there is no nonzero closed 1-form $\alpha$ of the form $$ \alpha= c h_3^{-\frac{1}{3}}\ol{\xi} + \boldsymbol{\theta}, \quad \boldsymbol{\theta}\in\mathtt{I}^{(\bar{1})}. $$ Repeat the similar computations and, by induction argument, solve for the sequence of coefficients $\{\, a_{\bar{1}}, a_0, a_{\xi}, a_1, a_2, \, ... \, \}$. Continuing this process, one arrives at the following normal form for the closed 1-form $\alpha$: $${\rm d} f= \alpha= c \left(h_3^{-\frac{1}{3}}\ol{\xi} +f^{\xi}\xi+\sum_{j=-2}^{k}f^j\theta_j\right), \qquad ( \theta_{-2}=\overline{\theta}_{2}, \; \textnormal{etc}), $$ for the given constant $c$, where $$f^j\in\mathcal O(k-1)\quad\textnormal{for}\;\; j\geq 4.$$ \vspace{2mm} \begin{itemize} \item Suppose $c=0$. Then ${\rm d} f =0$ and $f$ is a constant. \item Suppose $c\ne 0$. We show that this leads to a contradiction. Applying the commutation relation \eqref{eq:lxib}, $$[E_k, \partial_{\ol{\xi}}]=\sum_{j\geq k+1}T_j^k E_j,$$ to the given function $f$, one gets $$ \partial_{\ol{\xi}}(f^k)= E_k(ch_3^{-\frac{1}{3}})=0. $$ By the induction hypothesis, this implies that $f^k$ is a constant multiple of $h_3^{-\frac{k}{3}}$. \begin{itemize} \item Suppose $f^k=0$. Then $f\in\mathcal O(k-1)$ and, again by the induction hypothesis, $f\equiv \textnormal{constant}$. Hence $c=0$, a contradiction. \item Suppose $f^k\ne 0$. Applying the commutation relation $$ [E_{k-1}, \partial_{\ol{\xi}}]=\sum_{j\geq k}T_j^{k-1} E_j$$ to the given function $f$, one gets (since $k\geq 5$) $$ -\partial_{\ol{\xi}}(f^{k-1})=T^{k-1}_k f^k. $$ It is easily checked that, using the induction hypothesis, this forces $f^k=0$, a contradiction. \end{itemize}\end{itemize} \subsection{Lemma \ref{lem:delbpoly0}} Consider the polynomial rings in the balanced coordinates, $$\mathcal R:=\mathbb C[z_4, z_5, z_6, \, ... \, ], \quad \overline{\mathcal R}:=\mathbb C[\ol{z}_4, \ol{z}_5, \ol{z}_6, \, ... \, ]. $$ Recall the spectral weights in Defn.\ref{defn:spectralweights}. Define the associated sequences of polynomial vector spaces filtered by the spectral weight as follows: \begin{equation} \begin{cases} \quad P_d&=\{ \textnormal{weighted homogeneous polynomials of degree $d\geq 0$ in $\mathcal R$} \}, \notag\\ \quad \mathcal P_d&=\oplus_{i=0}^{d} P_i\subset\mathcal R, \notag\\ \quad \mathcal P_d(k)&=\mathcal P_d \cap \mathcal O(k)\subset\mathcal R\cap \mathcal O(k), \notag\\ \quad Q_d&=P_d\oplus (h_3\bar h_3) P_d, \notag\\ \quad \mathcal Q_d&= \oplus_{i=0}^{d} Q_{i},\notag\\ \quad \mathcal Q_d(k)&=\mathcal Q_d \cap \mathcal O(k).\notag \end{cases} \end{equation} \vspace{1mm} The following lemma, which is an application of Lem.\ref{lem:lemma5.4'}, records a characteristic rigidity property of the subspace $\mathcal P_d(k)$ under the differential operator $h_3^{\frac{1}{3}}\partial_{\ol{\xi}}$. \begin{lem}\label{lem:delbpoly0} Let $v\in\mathcal O(k), k\geq 4$. Suppose $$h_3^{\frac{1}{3}}\partial_{\ol{\xi}} v\in\mathcal Q_{d}(k). $$ Then $$v\in\mathcal P_{d+1}(k).$$ \end{lem} \begin{cor}\label{cor:delbpoly} Let $u\in\mathcal O(k), k\geq 4$. Let $u^{k}=E_{k} (u)=\frac{{\partial} u}{{\partial} h_{k }}$. Suppose $$h_3^{\frac{1}{3}}\partial_{\ol{\xi}} (h_3^{\frac{k}{3}} u^{k }) \in\mathcal Q_{d}(k). $$ Then $h_3^{\frac{k}{3}}u^k\in\mathcal P_{d+1}(k)$, and hence $$u\in\mathcal P_{d+(k-2)}(k)\mod \mathcal O(k-1).$$ \end{cor} \begin{proof} Substitute $v=h_3^{\frac{k}{3}} u^{k }$ from Lem.\ref{lem:delbpoly0}. \end{proof} \noindent\emph{Proof of Lem.~ \ref{lem:delbpoly0}}.\, Consider the case $k=4.$ For functions in $\mathcal O(4)$, we have the commutation relation $[E_4,\partial_{\ol{\xi}}]=0.$ Hence by applying $E_4$ repeatedly to $\partial_{\ol{\xi}} v$, we get $$h_3^{\frac{1}{3}}\partial_{\ol{\xi}} E^{m}_4(v)=0$$ for some $m\leq d+1$. By Cor~.\ref{cor:lemma5.4}, $v$ is a polynomial in $z_4$ and one may write $$v= v_m z_4^m+v_{m-1}z_4^{m-1}+\, ... \, +v_1 z_4+v_0, $$ where $v_j\in\mathcal O(3)$ with $v_m$ being a constant. Substitute this to the given equation for $h_3^{\frac{1}{3}}\partial_{\ol{\xi}} v$, and one gets the recursive equations $$h_3^{\frac{1}{3}}\partial_{\ol{\xi}} v_j+(j+1)v_{j+1}\frac{3}{2}R =c'_j\gamma^2+c_j^{"}h_3\bar h_3, \quad j=m, m-1, \, ... \,,$$ for constants $c'_j,c_j^{"}$ (set $v_{m+1}=0$). Since $v_m$ is a constant and $$h_3^{\frac{1}{3}}\partial_{\ol{\xi}} z_4=\frac{3}{2}R=\frac{3}{2}(\gamma^2-2h_3\bar h_3),$$ whereas $v_j\in\mathcal O(3)$, an inductive argument using Lem.\ref{lem:lemma5.4'} for $j$ from $m-1$ to $0$ shows that all the coefficients $v_j$ must be constant. This shows $$v\in\mathcal P_{d+1}(4).$$ Applying the induction argument, suppose the claim is true up to $\mathcal O(k-1)$. Let $v\in\mathcal O(k)$. For functions in $\mathcal O(k)$, we have the commutation relation $[E_k,\partial_{\ol{\xi}}]=0.$ Hence, similarly as above, $v$ is a polynomial in $z_k$ and one may write $$v= v_m z_k^m+v_{m-1}z_k^{m-1}+\, ... \, +v_1 z_k+v_0, $$ where $v_j\in\mathcal O(k-1)$, $m(k-3)\leq d+1$, and $v_m$ being a constant. Substitute this to the given equation for $h_3^{\frac{1}{3}}\partial_{\ol{\xi}} v$, and one gets the recursive equations $$h_3^{\frac{1}{3}}\partial_{\ol{\xi}} v_j+(j+1)v_{j+1}\hat{T}_k\in \mathcal Q_{d-j(k-3)}(k-1), \quad j=m, m-1, \, ... \,.$$ By Lem.\ref{lem:Tjh}, $\hat{T}_k\in \mathcal Q_{k-4}(k-1)$. It follows by the similar inductive argument as above, with decreasing $j$ from $m-1$ to $0$, that $$v_j\in \mathcal P_{d-j(k-3)+1}(k-1), \quad m-1\geq j\geq 0.$$ This shows $v\in\mathcal P_{d+1}(k).$ \hfill $\square$ \section{Jacobi fields}\label{sec:Jacobifields} From the general theory of differential equations, \cite{Vinogradov1989}, a generating function of symmetry of a differential equation is characterized as an element in the kernel of its linearization. For the minimal Lagrangian system, it turns out that the generating functions of symmetries are Jacobi fields. Since the normal bundle of a Lagrangian surface is canonically isomorphic to the cotangent bundle, the corresponding Jacobi operator (i.e., the defining equation for Jacobi fields) reduces to the second order operator $$\partial_{\xi}\partial_{\ol{\xi}}+\frac{3}{2}\gamma^2.$$ In particular, Jacobi fields correspond to the eigenfunctions of Laplacian of the induced metric. In this section, we give a complete classification of the (pseudo) Jacobi fields for the minimal Lagrangian system based on the results of \S\ref{sec:prolonglemmas}. An infinite sequence of higher-order (pseudo) Jacobi fields will be constructed in the next section by the similar recursion procedure as discussed in \S\ref{sec:recursionrel}. \vspace{2mm} In \S\ref{sec:JacobiSymmetry}, we show that a Jacobi field is a generating function of symmetry for the minimal Lagrangian system. In \S\ref{sec:splitting}, we prove a splitting theorem that a Jacobi field decomposes as the sum of a classical Jacobi field and a higher-order Jacobi field. Combining this with the results from \cite{Fox2011,Fox2012}, we obtain a complete classification of Jacobi fields in \S\ref{sec:classification}. We find that a higher-order ($\geq5$) Jacobi field exists at each odd order $5,$ or $1$ mod $6.$ The similar analysis applies to the classification of pseudo-Jacobi fields. A pseudo-Jacobi field corresponds to a symmetry of the elliptic Tzitzeica equation underlying the minimal Lagrangian system. We find that a higher-order ($\geq 4$) pseudo-Jacobi field exists at each even order $4,$ or $2$ mod $6.$ \subsection{Jacobi fields and pseudo-Jacobi fields}\label{sec:JacobiSymmetry} \subsubsection{Definition} Motivated by Defns.\ref{defn:classicalJacobifield}, \ref{defn:classicalpseudoJacobifield}, we give a general definition of (pseudo) Jacobi fields on $\hat{X}^{(\infty)}.$\footnotemark\footnotetext{It turns out that the infinite sequence of higher-order (pseudo) Jacobi fields belong to $\mathcal R\oplus\overline{\mathcal R}\subset C^{\infty}(\hat{X}^{(\infty)}_{**}).$} \begin{defn}\label{defn:Jacobifield} A scalar function $A:\hat{X}^{(\infty)}\to\mathbb C $ is a \textbf{Jacobi field} if it satisfies the linear Jacobi equation \begin{equation}\label{eq:Jacobieq} \mathcal E(A):=\partial_{\xi}\partial_{\ol{\xi}} A+\frac{3}{2} \gamma^2A=0. \end{equation} The $\mathbb C$-vector space of Jacobi fields is denoted by $\mathfrak J^{(\infty)}$. A scalar function $P:\hat{X}^{(\infty)}\to\mathbb C$ is a \textbf{pseudo-Jacobi field} if it satisfies the linear pseudo-Jacobi equation \begin{equation}\label{eq:Jacobieq'} \mathcal E'(P):=\partial_{\xi}\partial_{\ol{\xi}} P+\frac{1}{2} ( \gamma^2+4h_3\bar h_3)P=0. \end{equation} The $\mathbb C$-vector space of pseudo-Jacobi fields is denoted by $\mathfrak J'^{(\infty)}$. A \textbf{higher-order Jacobi field} is a Jacobi field in the ring of the balanced coordinates $\mathcal R\oplus\overline{\mathcal R}$. The subspace of higher-order Jacobi fields is denoted by $\mathfrak J^{(\infty)}_h$. Let $\mathfrak J^{(k)}\subset\mathfrak J^{(\infty)}$ be the subspace of \textbf{Jacobi fields of order $\leq k+2$} defined on $\Xh{k}$. Let $\mathfrak J^{(k)}_h=\mathfrak J^{(k)}\cap\mathfrak J^{(\infty)}_h.$ The space of \textbf{Jacobi fields of order $k+2$} is defined as the quotient space \begin{align} \mathfrak J^k&=\mathfrak J^{(k)}/\mathfrak J^{(k-1)},\quad k\geq 1, \notag \\ \mathfrak J^0&=\mathfrak J^{(0)}=\{\textnormal{\,classical Jacobi fields }\}.\notag \end{align} The corresponding subspaces of pseudo-Jacobi fields are denoted by $\mathfrak J'^{(\infty)}_h$, $\mathfrak J'^{(k)}$, $\mathfrak J'^{(k)}_h$, $\mathfrak J'^k,\mathfrak J'^{(0)}$. \end{defn} \subsubsection{Stability lemma} Before we proceed to the analysis of Jacobi fields and symmetries, let us record the useful formulas related to the stability of the ring of balanced coordinates under the Jacobi operators $\mathcal E,\mathcal E'$. \begin{lem}\label{lem:mcez} Consider the polynomial ring $\mathcal R=\mathbb C[z_4, z_5, \, ... \, ]$ of balanced coordinates. We have, \begin{align} \mathcal E(z_j)&=\hat{T}_{j+1}-\frac{j}{3}(z_j \hat{T}_4+z_4 \hat{T}_j)+\frac{3}{2} \gamma^2z_j, \notag\\ \mathcal E'(z_j)&=\hat{T}_{j+1}-\frac{j}{3}(z_j \hat{T}_4+z_4 \hat{T}_j)+\frac{1}{2}(\gamma^2+4\textnormal{r}^2) z_j.\notag \end{align} It follows that $$\mathcal E(\mathcal R), \mathcal E'(\mathcal R)\subset \mathcal R\oplus (h_3\bar h_3)\mathcal R. $$ In particular, the operators $\mathcal E, \mathcal E'$ preserve the spectral weight when acting on $\mathcal R$. \end{lem} Note $\hat{T}_4=\frac{3}{2}(\gamma^2-2 \textnormal{r}^2).$ \subsubsection{Jacobi field and symmetry} It was observed in \S\ref{sec:classicallaws} that there exists an isomorphism between the classical Jacobi fields and the classical symmetries. The higher-order analogue of this isomorphism is true, and a Jacobi field uniquely determines a (vertical) symmetry vector field of $(\hat{X}^{(\infty)}, \hat{\rm I}^{(\infty)})$. \vspace{2mm} Consider the coframe of $\Fh{\infty}$, \[\{\,\rho,\,\xi,\,\ol{\xi},\,\theta_0,\,\theta_1,\,\ol{\theta}_1,\,\theta_2,\, \ol{\theta}_2,\,... \,\},\] and its dual frame \[\{\,E_\rho,\,E_{\xi},\,E_{\ol{\xi}},\,E_0,\,E_1,\,E_{\bar{1}},\,E_2,\, E_{\bar{2}},\,... \,\}.\] By a vector field on $\hat{X}^{(\infty)}$, we mean a vector field (derivation) on $\Fh{\infty}$ of the form $$ V=V_{\xi}E_{\xi} +V_{\ol{\xi}}E_{\ol{\xi}}+V_0 E_0+\sum_{j=1}^{\infty} ( V_j E_j + V_{\bar{j}}E_{\bar{j}})$$ which is invariant under the action of the structure group $\liegroup{SO}(2)$ of the principal bundle $\Fh{\infty}\to\hat{X}^{(\infty)}$. We denote the set of vector fields on $\hat{X}^{(\infty)}$ by $H^0(T\hat{X}^{(\infty)})$. \begin{defn} A vector field $V\in H^0(T\hat{X}^{(\infty)})$ is a \textbf{symmetry} of the differential system $(\hat{X}^{(\infty)},\Ih{\infty})$ if the formal Lie derivative $ \mathcal L_V$ preserves the ideal $\Ih{\infty}$, \[ \mathcal L_V \Ih{\infty} \subset \Ih{\infty}.\] The $\mathbb C$-algebra of symmetry vector fields is denoted by $\mathfrak{S}$. A symmetry $V\in\mathfrak{S}$ is \textbf{vertical} when $V_{\xi}=V_{\ol{\xi}}=0$ and it has no (horizontal) $E_{\xi}, \,E_{\ol{\xi}}$ components. The subspace of vertical symmetries is denoted by $\mathfrak{S}_v$. \end{defn} \vspace{1mm} We wish to give an analytic characterization of symmetry. Consider a vertical symmetry \begin{equation}\label{eq:symvf} V=V_0 E_0+\sum_{j=1}^{\infty} ( V_j E_j + V_{\bar{j}}E_{\bar{j}}). \end{equation} We claim that $V_0$ is necessarily a Jacobi field, and that it is the generating function of symmetry in the sense that $V_j, V_{\bar{j}}$'s are determined by $V_0$ and its successive derivatives. \vspace{2mm} \textbf{Step 0}. The condition that the Lie derivative $\mathcal L_V\theta_0\equiv 0\mod\Ih{\infty}$ shows that \begin{align} {\rm d} V_0-V\lrcorner \left(\frac{1}{2}(\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi})\right)&\equiv {\rm d} V_0 - \frac{1}{2}(V_1 \xi+V_{\bar{1}} \ol{\xi}) \notag\\ &\equiv 0 \mod \hat{\rm I}^{(\infty)}.\notag \end{align} One gets \begin{equation}\label{eq:symstep0} {\rm d} V_0\equiv \frac{1}{2}(V_1 \xi+V_{\bar{1}} \ol{\xi})\mod \hat{\rm I}^{(\infty)}. \end{equation} \textbf{Step 1}. By a similar computation, the condition that the Lie derivatives $\mathcal L_V\theta_1, \, \mathcal L_V\ol{\theta}_1\equiv 0\mod\I{\infty}$ shows that \begin{equation}\label{eq:symstep1}\begin{array}{rcrll} {\rm d} V_1 + \textnormal{i} V_1\rho&\equiv& V_2 \xi&-3\gamma^2 V_0\ol{\xi},& \\ {\rm d} V_{\bar{1}} - \textnormal{i} V_{\bar{1}}\rho&\equiv& V_{\bar{2}}\ol{\xi}&-3\gamma^2 V_0\xi,&\mod\;\hat{\rm I}^{(\infty)}. \end{array} \end{equation} Note that Eq.\eqref{eq:symstep0} and Eq.\eqref{eq:symstep1} imply that $V_0$ is a Jacobi field. \textbf{Step j}. The rest of the coefficients $V_j, V_{\overline{j}}, j\geq 2,$ are similarly determined by $V_0$ and its successive derivatives by the condition that the Lie derivatives $\mathcal L_V\theta_j, \, \mathcal L_V\ol{\theta}_j\equiv 0\mod\Ih{\infty}$. These equations imply that \begin{equation}\label{eq:symstepj}\begin{array}{rcll} {\rm d} V_j +j \textnormal{i} V_j\rho&\equiv& \left( V_{j+1} +\{ V_j, V_{\bar{j}}, \, ... \, V_0\} \right) \xi+\{ V_{j-1}, V_{\overline{j-1}}, \, ... \, V_0\}\ol{\xi}, \\ {\rm d} V_{\bar{j}} -j \textnormal{i} V_{\bar{j}}\rho&\equiv& \left(V_{\overline{j+1}} +\{ V_j, V_{\bar{j}}, \, ... \, V_0\} \right) \ol{\xi}+\{ V_{j-1}, V_{\overline{j-1}}, \, ... \, V_0\}\xi, \mod\;\hat{\rm I}^{(\infty)}.&\\ \end{array} \end{equation} Here $\{ V_{j}, V_{\overline{j}}, \, ... \, V_0\}$, etc, is a generic notation for an expression which is linear in $\{ V_k, V_{\bar{k}}\}_{k=0}^{j}$ with coefficients in the ring $\mathbb C[h_3, \bar h_3,h_4,\bar h_4, \, ... \, ]$. \vspace{2mm} In fact, the following converse of this analysis is true. \begin{prop}\label{prop:symmetry} The generating function of a vertical symmetry of the differential system for minimal Lagrangian surfaces is a Jacobi field. Conversely, a Jacobi field $A$ uniquely determines a vertical symmetry $V$ of the form \eqref{eq:symvf} with the generating function $V_0=A$. As a consequence, there exists a canonical isomorphism \[ \mathfrak{J}^{(\infty)}\simeq \mathfrak{S}_v.\] \end{prop} \begin{proof} The compatibility of the recursive defining equations \eqref{eq:symstep0}, \eqref{eq:symstep1}, and \eqref{eq:symstepj} can be checked from the formula for $T_j$ and its differential consequences. We omit the details. \end{proof} \subsubsection{Interpretation of pseudo-Jacobi field}\label{sec:interpseudo} The pseudo-Jacobi fields correspond to the symmetries of the elliptic Tzitzeica equation underlying the minimal Lagrangian system. The symmetries of the elliptic Tzitzeica equation are classified in \cite{Fox2011,Fox2012}. Away from the umbilic divisor on a minimal Lagrangian surface, take a local holomorphic coordinate $z$ such that $${\rm d} z =h_3^{\frac{1}{3}}\xi,$$ and the Hopf differential is normalized to ${\rm I \negthinspace I}=({\rm d} z)^3.$ Set accordingly, $$\xi=e^{\frac{u}{2}}{\rm d} z, \quad h_3=e^{-\frac{3}{2}u}$$ for a real scalar function $u$. The connection 1-form $\rho$ is given by $\rho=\frac{\textnormal{i}}{2}(u_z{\rm d} z -u_{\overline{z}}{\rm d}\overline{z}),$ and it follows that the curvature $R$ is, $$R=-2 e^{-u} u_{z\overline{z}}.$$ The compatibility equation \eqref{eq:Gauss} then translates to the elliptic Tzitzeica equation \begin{equation}\label{eq:Tzitzeica} u_{z\overline{z}}+\frac{1}{2}\left(\gamma^2 e^u-2 e^{-2u}\right)=0. \end{equation} On the other hand, consider the pseudo-Jacobi operator \eqref{eq:Jacobieq'}. From ${\rm d} z=h_3^{\frac{1}{3}}\xi$, one gets ${\partial}_z=h_3^{-\frac{1}{3}}\partial_{\xi}$. Hence $$\partial_{\xi}\partial_{\ol{\xi}}=e^{-u} {\partial}_z{\partial}_{\overline{z}},$$ and the pseudo-Jacobi operator translates to $$\mathcal E'=e^{-u}\left({\partial}_z{\partial}_{\overline{z}}+\frac{1}{2}(\gamma^2 e^u+4 e^{-2u})\right).$$ Up to scaling by $e^{-u}$, this is the linearization of \eqref{eq:Tzitzeica} and the claim follows. \subsubsection{Examples} \begin{exam}\label{exam:initialdata} A direct computation show that $$z_4$$ is a pseudo-Jacobi field (of order 4 and degree 1), and $$z_5-\frac{5}{3}z_4^2$$ is a Jacobi field (of order 5 and degree 2). \end{exam} This example provides a hint for the existence of higher-order (pseudo) Jacobi fields. \subsection{Splitting theorem}\label{sec:splitting} For a nonlinear differential equation, it is a stringent condition to admit a higher-order symmetry, not to mention an infinite number of them. The corresponding Jacobi equation would imply a variety of compatibility conditions for the given differential equation, and generically one expects that the space of Jacobi fields is trivial. In this section, we explore the constraints given by the Jacobi equation of the minimal Lagrangian system by a repeated application of Lem.\ref{lem:lemma5.4'}, Lem.\ref{lem:delbpoly0}, and Cor.\ref{cor:delbpoly}. As a result, we obtain a rough normal form for Jacobi fields and this implies that the space of Jacobi fields splits into the direct sum of the classical Jacobi fields and the higher-order Jacobi fields. A similar splitting theorem holds for the pseudo-Jacobi fields. \subsubsection{Symbol lemma} We start with a lemma on the normal form for the highest order term of a Jacobi field. \begin{lem}\label{lem:Jacobinormal} Let $A\in \mathfrak J^{(k)}\subset\mathcal O(k+2)$ be a (pseudo) Jacobi field. Suppose $A^{k+2}=E_{k+2}(A)\ne 0$. Then $A$ is at most linear in the highest order variable $z_{k+2}=h_{3}^{-\frac{k+2}{3}}h_{k+2}$, and, up to constant scale, it admits the normal form $$ A=z_{k+2}+\mathcal O(k+1). $$ \end{lem} \begin{proof} Since $A\in\mathcal O(k+2)$, we have \begin{align} A_{\xi}&\equiv h_{k+3} A^{k+2} \mod \mathcal O(k+2), \notag\\ A_{\xi,\ol{\xi}}&\equiv h_{k+3} \partial_{\ol{\xi}}(A^{k+2})\mod \mathcal O(k+2),\notag \\ &\equiv 0\mod\mathcal O(k+2), \quad \textnormal{for \;$\mathcal E(A)=0$\, (\textnormal{or}\; $\mathcal E'(A)=0$).}\notag \end{align} This forces $\partial_{\ol{\xi}}(A^{k+2})=0$. By Cor.\ref{cor:lemma5.4}, $A^{k+2}$ is a constant multiple of $h_3^{-\frac{k+2}{3}}$. \end{proof} \subsubsection{Splitting theorem} We refine Lem.\ref{lem:Jacobinormal} to the splitting theorem for (pseudo) Jacobi fields with the help of Cor.\ref{cor:delbpoly}. Recall $\mathcal R=\mathbb C[z_4, z_5, z_6,\,... \, ]$, $\overline{\mathcal R}=\mathbb C[\ol{z}_4, \ol{z}_5, \ol{z}_6,\,... \, ].$ \begin{prop}\label{prop:splitting} The space of Jacobi fields splits into the direct sum $$ \mathfrak{J}^{(\infty)}= \mathfrak{J}^{(\infty)}_h\oplus \mathfrak{J}^{(0)} $$ of the higher-order Jacobi fields and the classical Jacobi fields. By definition, $$ \mathfrak{J}^{(\infty)}_h \subset \mathcal R\oplus \overline{\mathcal R} $$ and the space of higher-order Jacobi fields is generated by the un-mixed weighted homogeneous polynomial Jacobi fields. A similar splitting theorem holds for the space of pseudo-Jacobi fields. \end{prop} \begin{proof} Consider first the Jacobi field case. Let $A\in\mathcal O(k)$ be a Jacobi field for $k\geq 5$. By Lem.\ref{lem:Jacobinormal} above, we may set $$ A=z_{k}+u_{(1)},\quad u_{(1)}\in\mathcal O(k-1).$$ Applying the Jacobi operator, one finds $$-\mathcal E(z_{k})\equiv h_3^{\frac{1}{3}}\partial_{\ol{\xi}}(h_3^{\frac{(k-1)}{3}}u_{(1)}^{k-1})z_{k}\mod\mathcal O(k-1).$$ By Lem.\ref{lem:mcez}, $\mathcal E(z_{k})\in\mathcal Q_{k-3}(k)$. By Cor.\ref{cor:delbpoly}, one gets $$ u_{(1)} = p_{(1)}+\mathcal O(k-2)$$ for some $p_{(1)}\in \mathcal P_{k-3}(k-1).$ Suppose by induction we arrive at the formula \begin{equation}\label{eq:Az2kj} A=z_{k}+p_{(1)}+p_{(2)}+\, ... \, +p_{(j)}+u_{(j+1)},\quad u_{(j+1)}\in\mathcal O(k-(j+1)), \end{equation} where each $p_{(i)}\in \mathcal P_{k-3}(k-i)$ such that $$ q_{(i)}:=\mathcal E\left(z_{k}+p_{(1)}+p_{(2)}+\, ... \, +p_{(i)}\right)\in\mathcal O(k-i). $$ By Lem.\ref{lem:mcez}, one finds $q_{(j)}\in\mathcal Q_{k-3}(k-j)$. Applying the Jacobi operator to the refined normal form \eqref{eq:Az2kj}, one gets $$ -q_{(j)}\equiv h_3^{\frac{1}{3}}\partial_{\ol{\xi}}(h_3^{-\frac{k-(j+1)}{3}}u_{(j+1)}^{k-(j+1)})z_{k-j}\mod\mathcal O(k-(j+1)).$$ By Cor.\ref{cor:delbpoly}, one may write \begin{align*} u_{(j+1)}&=p_{(j+1)}+u_{(j+2)}, \\ p_{(j+1)}&\in\mathcal P_{k-3}(k-(j+1)),\\ u_{(j+2)}&\in\mathcal O(k-(j+2)). \end{align*} Continuing this process, we arrive at the normal form \begin{equation}\begin{array}{rll} A&=p+u_{(k-3)}, &u_{(k-3)}\in\mathcal O(3),\notag \\ p&=z_{k}+p_{(1)}+p_{(2)}+\, ... \, +p_{(k-4)},& \end{array} \end{equation} where $p_{(k-4)}\in \mathcal P_{k-3}(4)$ such that $$ q_{(k-4)}:=\mathcal E(p)\in\mathcal Q_{k-3}(4). $$ Since $\mathcal E$ is a real operator, the complex conjugate of this argument implies that the Jacobi field $A$ decomposes into $$A =f+g,$$ where $f$ is an un-mixed pure polynomial in $z_j, \ol{z}_j$, and $g$ is a function on $\Xh{1}$. Now, the Jacobi operator $\mathcal E$ preserves the spectral weight. Since $\mathcal E(g)\in\mathcal O(4)\cap\mathcal O(-4)$ is at most linear in $z_4,\ol{z}_4$, this implies (note $\mathcal E(z_4)=R z_4$), \begin{equation}\label{eq:efeg} \mathcal E(c'z_4+c''\ol{z}_4)=R(c'z_4+c''\ol{z}_4)=-\mathcal E(g), \end{equation} for some constants $c', c"$, where $c'z_4+c''\ol{z}_4$ denotes the terms of spectral wight $\pm 1$ in the polynomial $f$. A short computation shows that this forces $c', c"=0$, and $f\in\mathcal R\oplus\overline{\mathcal R}$ is a higher-order Jacobi field. Hence, Eq.\eqref{eq:efeg} also implies that $\mathcal E(g)=0$ and consequently, $$h_3 E_3(g), \;\bar h_3 E_{\overline{3}}(g)\equiv\textnormal{const}.$$ It is easily checked from this that $E_3(g), E_{\overline{3}}(g)=0$ and $g$ is necessarily a classical Jacobi field defined on $X$. \vspace{2mm} Consider next the pseudo-Jacobi field case. By the similar argument as above, a pseudo-Jacobi field $P$ decomposes into $$P=f+g, $$ where $f$ is an un-mixed pure polynomial in $z_j, \ol{z}_j$, and $g$ is a function on $\Xh{1}$. Since $\mathcal E'(g)\in\mathcal O(4)\cap\mathcal O(-4)$ is at most linear in $z_4,\ol{z}_4$ and $z_4,\ol{z}_4$ are pseudo-Jacobi fields, it follows that $$\mathcal E'(f)=0, \quad \textnormal{and hence }\; \mathcal E'(g)=0. $$ By the similar argument as above, this implies that $g$ is necessarily a classical pseudo-Jacobi field. \end{proof} \subsection{Classification}\label{sec:classification} Combining the results of \S\S\ref{sec:JacobiSymmetry}, \ref{sec:splitting}, we state a complete classification of (pseudo) Jacobi fields. The classification consists of the following steps. We first show, by a direct computation, that there exist no even order Jacobi fields in $\mathcal R\oplus\overline{\mathcal R}.$ This enables to apply the classification results for elliptic Tzitzeica equation in \cite{Fox2011,Fox2012}, and we find that a nontrivial higher-order Jacobi field exists at order $5$, or $1$ mod $6$ only. The infinite sequence of higher-order Jacobi fields of the admissible orders will be constructed in \S\ref{sec:formalKilling}. The classification of pseudo-Jacobi fields follows from \S\ref{sec:interpseudo}, Prop.\ref{prop:splitting}, and \cite{Fox2011,Fox2012}. \begin{thm}\label{thm:classifyJacobi} The infinitely prolonged differential system $(\hat{X}^{(\infty)},\hat{\rm I}^{(\infty)})$ for minimal Lagrangian surfaces in a $2_{\mathbb C}$-dimensional, non-flat, complex space form admits an infinite sequence of higher-order (pseudo) Jacobi fields as follows. \begin{enumerate}[\qquad a)] \item There exists a unique (up to constant scale) nontrivial weighted homogeneous polynomial Jacobi field in $\mathcal R$ of degree $d\geq 2$ for each $$d\equiv 2, \, 4\mod 6,$$ (hence of order $\equiv 5, \,1\mod 6$). The classical Jacobi fields, and these higher-order Jacobi fields and their complex conjugates generate the space of Jacobi fields $\mathfrak{J}^{(\infty)}.$ \item There exists a unique (up to constant scale) nontrivial weighted homogeneous polynomial pseudo-Jacobi field in $\mathcal R$ of degree $d\geq 2$ for each $$d\equiv 1, \, 5\mod 6,$$ (hence of order $\equiv 4, \,2\mod 6$). The classical pseudo-Jacobi fields, and these higher-order pseudo-Jacobi fields and their complex conjugates generate the space of pseudo-Jacobi fields $\mathfrak{J}'^{(\infty)}.$ \end{enumerate} \end{thm} We present the proof of the theorem in the following two subsections. \subsubsection{No even order Jacobi fields}\label{sec:noeven}\; Recall $\mathcal E(\mathcal R)\subset\mathcal R\oplus(h_3\bar h_3)\mathcal R.$ Let $F=\frac{3}{2}x_0 z_{2k}+\, ... \, \in\mathcal R$ be an even order weighted homogeneous polynomial Jacobi field of weight $2k-3$ (for a constant $\frac{3}{2}x_0$). Consider the expansion, \begin{equation}\begin{array}{rlll} F= \frac{3}{2}x_0z_{2k}&+z_{2k-1} (&x_1z_4 &) \\ &+z_{2k-2} (&x_2z_5 &+y_2z_4^2) \\ &+z_{2k-3} (&x_3z_6 &+y_3z_5z_4 +\, ... \, ) \\ &+z_{2k-4} (&x_4z_7 &+y_4z_6z_4 +\, ... \, ) \\ & ... && \\ &+z_{2k-(i-1)} (&x_{i-1}z_{i+2} &+y_{i-1}z_{i+1}z_4 +\, ... \, ) \\ &+z_{2k-i} (&x_i z_{i+3} &+y_iz_{i+2}z_4 +\, ... \, ) \\ &+z_{2k-(i+1)} (&x_{i+1}z_{i+4} &+y_{i+1}z_{i+3}z_4 +\, ... \, ) \\ &...&& \\ &+z_{k+3} (&x_{k-3}z_k &+y_{k-3}z_{k-1}z_4 +\, ... \, ) \\ &+z_{k+2} (&x_{k-2}z_{k+1} &+y_{k-2}z_k z_4 +\, ... \, ) \\ &&&+y_{k-1} z_{k+1}^2z_4 +\, ... \, . \end{array} \end{equation} Here $\{ x_i, y_i\}$ are constant coefficients. First, by considering the $z_{2k}$ term in $\mathcal E(F)$, we get \begin{equation}\label{eq:x10} x_1+(a_{2k,0}+\frac{3}{2} -\frac{2k}{3}a_{3,0} )x_0=0. \end{equation} In order to extract the compatibility equations imposed on the $x_i$-coefficients only, from now on we compute modulo the curvature $R=\gamma^2-2 h_3\bar h_3$. It means that we identify $$h_3\bar h_3\equiv\frac{\gamma^2}{2}.$$ Consider the Jacobi equation $$\mathcal E(F)\mod R,$$ and collect the equations from the coefficients of the monomials $z_{2k}, z_{2k-1}z_4, z_{2k-2}z_5, \, ... $ up to $z_{k+2} z_{k+1} .$ i.e. the terms with the coefficient $x_j$'s (let's call them the principal terms). It is easily checked that when acted on by the Jacobi operator, the terms not appearing in the above expansion do not have any contribution to the principal terms. Also, since ${\partial}_{\omega}z_4\equiv 0\mod R$, by computing mod $R$ we eliminate the contributions from the $y_j$ coefficients, as claimed above. It follows that one may evaluate $$\mathcal E(\{\textnormal{principal terms}\})\mod R,$$ and check only the principal terms in the image. This would yield a set of linear equations among the coefficients $x_j$'s. Recall the following formulas: \begin{align*} T_{j+1}&= \sum_{s=0}^{j-3} a_{j,s} \, h_{j-s} \, \partial^{s}_{\xi} R, \; \; \; \mbox{for} \; \; j\geq 3, \notag \\ & \quad \;\;a_{j,s} =\frac{(j+2s+3) }{2(j-1)} {j-1 \choose s+2}, \notag\\ & \quad \partial^{s}_{\xi} R=\delta_{0s}\gamma^2 - 2 h_{3+s} \bar h_3, \notag\\ {\rm d} z_j&\equiv \left(z_{j+1}-\frac{j}{3} z_4 z_j \right) \omega + \hat{T}_{j} \textnormal{r}^{-\frac{2}{3}}\ol{\omega} \;\mod \hat{\rm I}^{(\infty)}, \quad{\rm for} \; j \geq 4.\notag \end{align*} Note that, for $j\geq 4$, \begin{align} \partial_{\ol{\xi}}\partial_{\xi}(z_j)&\equiv\hat{T}_{j+1}-\frac{j}{3}( z_4 \hat{T}_j) \mod R, \notag\\ \hat{T}_{j+1}&\equiv-\gamma^2 a_{j,j-3} z_j +\textnormal{ (quadratic terms in $z_i$'s) }\mod R.\notag \end{align} Since the principal terms except $z_{2k}$ are quadratic in the balanced coordinates $z_j$'s, the term $-\gamma^2 a_{j,j-3} z_j $ is the only form of contribution to the principal terms from $\partial_{\ol{\xi}} z_{j+1}$ when $j+1<2k$. \vspace{2mm} With this preparation, a direct computation yields the following formulas for the principal terms. We record only the relevant terms (here we set the scaling factor $\gamma^2=1$ temporarily for simplicity). \begin{align*} -\mathcal E(z_{2k})&\equiv ( a_{2k,2k-3} -\frac{3}{2}) \, z_{2k} + ( a_{2k,2k-4} + a_{2k,1} - \frac{2k}{3}a_{2k-1,2k-4}) \,z_4 z_{2k-1} \\ &\qquad+\sum_{j=2}^{k-2} ( a_{2k,2k-j-3} +a_{2k,j}) \, z_{j+3} z_{2k-j} ,\\ -\mathcal E(z_4z_{2k-1})&\equiv \begin{cases} &\quad\cdot \\ &(a_{4,1}+a_{2k-1,2k-4}-\frac{3}{2}) \,z_{4}z_{2k-1}\\ &(a_{2k-2,2k-5}) \,z_{5}z_{2k-2} \end{cases}, \\ &... \\ -\mathcal E(z_{j+3}z_{2k-j})&\equiv \begin{cases} &( a_{j+2,j-1} ) \,z_{j+2}z_{2k-(j-1)} \\ &(a_{j+3,j}+a_{2k-j,2k-(j+3)}-\frac{3}{2}) \,z_{j+3}z_{2k-j}\\ &(a_{2k-(j+1),2k-(j+4)}) \,z_{j+4}z_{2k-(j+1)} \end{cases}, \\ &... \\ -\mathcal E(z_{k+1}z_{k+2})&\equiv \begin{cases} & (a_{k,k-3}) \,z_{k}z_{k+3} \\ & (2 a_{k+1,k-2}+a_{k+2,k-1}-\frac{3}{2}) \,z_{k+1}z_{k+2} \\ &\quad\cdot \\ \end{cases}, \mod R. \end{align*} Note $a_{j+3,j}=\frac{3}{2}$ for all $j\geq0.$ Hence, except for the terms from $\mathcal E(z_{2k})$ and the last term $z_{k+1}z_{k+2}$, all the terms have the equal coefficient $\frac{3}{2}$. The resulting set of linear equations on the coefficients $x_j$'s are the following system of three term relations, including the initial equation \eqref{eq:x10}. \begin{equation}\begin{array}{rrrl} \cdot&\cdot& x_1 &+ (a_{2k,2k-3}+a_{2k,0} -k )x_0=0, \\ \cdot &x_1&+x_2 &+ (a_{2k,2k-4} + a_{2k,1} - k )x_0=0, \\ x_1 &+x_2&+x_3 &+ (a_{2k,2k-5} + a_{2k,2})x_0=0, \\ &&&...\\ x_{j-1}&+x_j&+x_{j+1}&+ (a_{2k,2k-j-3} +a_{2k,j})x_0=0,\\ &&&... \\ x_{k-4}&+x_{k-3}&+x_{k-2} &+ (a_{2k,k} +a_{2k,k-3})x_0=0,\\ x_{k-3}&+x_{k-2}&+x_{k-2} &+ (a_{2k,k-1} +a_{2k,k-2})x_0=0. \end{array}\end{equation} \vspace{1mm} It is left to show that this system of $k-1$ linear equations on the set of $k-1$ coefficient $\{ x_0, x_1, \, ... \, x_{k-2}\}$ has full rank. Set $t_j$ be the coefficient of $x_0$ of the $j$-th equation above, i.e., \begin{align*} t_0&=a_{2k,2k-3}+a_{2k,0} -k , \notag \\ t_1&=a_{2k,2k-4} + a_{2k,1} - k, \notag \\ t_j&=a_{2k,2k-j-3} +a_{2k,j}, \quad \textnormal{for} \;j\geq 2.\notag \end{align*} A direct computation shows that the determinant $\chi_k$ of the $(k-1)$-by-$(k-1)$ matrix for the set of linear equations is given by $$\pm \chi_k =\sum_{j=0}^{k-2} \epsilon^j_k t_j, $$ where $$\epsilon^j_k= \begin{cases} &-2\;\;\textnormal{when}\; j\equiv k \mod 3 \\ &+1\;\;\textnormal{otherwise}. \end{cases}$$ A computation mod 3 shows tha \begin{equation}\pm \chi_k = \begin{cases} & \frac{1}{2}\;\;\textnormal{when}\; k\equiv 0 \mod 3 \\ &1\;\;\textnormal{otherwise}. \end{cases} \end{equation} \subsubsection{Proof of Thm.\ref{thm:classifyJacobi}}\label{sec:proof} $\,$ \quad b)\; By Prop.\ref{prop:splitting} and the analysis in \S\ref{sec:interpseudo}, the higher-order pseudo-Jacobi fields correspond to the `Jacobi fields' for the elliptic Tzitzeica equation studied in \cite{Fox2011,Fox2012}. It follows from the classification result in \cite[Theorem 8.1]{Fox2012}. \quad a)\; From the analysis above, there exist no even order Jacobi fields. We remark, without giving the proofs, that the higher-order analogues of the results in \S\S\ref{sec:classicalcvlaw}, \ref{sec:cdcvlaw} are true: the space of (differentiated) conservation laws injects into the space of Jacobi fields under the natural \emph{symbol map} given by the differential of the associated spectral sequence; see \S\ref{sec:highercvlaws1}. Thus, Prop.\ref{prop:splitting} also implies the corresponding splitting theorem for the higher-order conservation laws. In particular, a higher-order conservation law of the minimal Lagrangian system corresponds to a higher-order conservation law of the elliptic Tzitzeica equation. From this, the induction argument using the recursion operators $\mathcal P, \mathcal N$ in \cite{Fox2012} shows that there are no Jacobi fields (for the minimal Lagrangian system) of degree $0\mod 6$ (or of order $3\mod 6$).\footnotemark\footnotetext{The obstruction to the application of the recursion operators $\mathcal P, \mathcal N$ lies in the space of higher-order conservation laws of even weight. This vanishes by the results from \cite{Fox2011,Fox2012}.} The sequence of Jacobi fields of the given degree $d\equiv 2, 4\mod 6$ will be constructed in \S\ref{sec:formalKilling}. \hfill$\square$ \section{Formal Killing fields}\label{sec:formalKilling} Recall that the original differential system for minimal Lagrangian surfaces is defined on the bundle of Lagrangian planes $X\to M$. It is a 6-symmetric space associated with the Lie group $\liegroup{SL}(3,\mathbb C)$, and the minimal Lagrangian surfaces arise as the primitive harmonic maps. From the theory of integrable systems, this implies that the $\liealgebra{g}$-valued Maurer-Cartan form $\psi$, \eqref{eq:psiform}, admits an extension to the $\liealgebra{g}^{\mathbb C}[\lambda^{-1},\lambda]$-valued ($\liealgebra{g}^{\mathbb C}=\liealgebra{sl}(3,\mathbb C)$) extended Maurer-Cartan form $\psi_{\lambda}$, \eqref{eq:primitiveform} in the below, by inserting the spectral parameter $\lambda.$ The structure equation for the minimal Lagrangian system shows that $\psi_{\lambda}$ is compatible and satisfies the Maurer-Cartan equation. In this section, we give a construction of the corresponding $\liealgebra{g}^{\mathbb C}[[\lambda]]$-valued canonical formal Killing fields to utilize this aspect of symmetry of the minimal Lagrangian system. The construction relies on the pair of 3-step recursions between Jacobi fields and pseudo-Jacobi fields which are embedded in the structure equation for the formal Killing fields. We give the differential algebraic inductive formulas for the pair of formal Killing fields that correspond to two particular sets of initial data. As a consequence, we will be able to read off the infinite sequence of higher-order (pseudo) Jacobi fields and conservation laws from the components of the formal Killing fields. In hindsight, the recursion relations were anticipated from the structure equation for the classical Killing field \eqref{eq:classicalKF}. Note that the 6-step recursion introduced in \cite{Fox2012} is the union of these two 3-step recursions when translated to our setting. \vspace{2mm} In \S\ref{sec:KFstrt}, we record the structure equation for the formal Killing fields with respect to the extended Maurer-Cartan form. In \S\ref{sec:KFinitial}, we determine the first few terms of the formal Killing fields for the two initial ans\"atze given by Exam.\ref{exam:initialdata}. The relevant observation is that the coefficient $h_3$ of Hopf differential provides the lower-end terms for the formal Killing fields, which allow one to truncate the terms of negative $\lambda$-degrees. With this preparation, we give in \S\ref{sec:KFformulae} the inductive formula for the formal Killing field for each set of the initial data. \subsection{Structure equation}\label{sec:KFstrt} In this section, we introduce the extended Maurer-Cartan form $\psi_{\lambda}$ and record the recursive structure equation (mod $\hat{\rm I}^{(\infty)}$) for the coefficients of the $\liealgebra{g}^{\mathbb C}[[\lambda]]$-valued formal Killing field with respect to $\psi_{\lambda}$. \subsubsection{Extended Maurer-Cartan form} Consider the $\liealgebra{g}$-valued 1-form $\psi$, Eq.\eqref{eq:psiform}. Evaluating mod $\hat{\rm I}^{(\infty)}$ (i.e., $\eta_2=h_3\xi, \ol{\eta}_2=\bar h_3\ol{\xi}$), it becomes \[\psi=\psi_{+}+\psi_{0}+\psi_{-}, \] where \begin{equation} \psi_{-} =\frac{1}{2} \left[ \begin {array}{ccc} \cdot &- \gamma &\textnormal{i} \gamma \notag \\ \noalign{\medskip} \gamma &\textnormal{i} h_{{3}} & - h_{{3}} \\ \noalign{\medskip} -\textnormal{i} \gamma &- h_{{3}} &- \textnormal{i} h_{{3}} \end {array} \right]\xi, \end{equation} \begin{equation} \psi_{0} = \left[ \begin {array}{ccc} \cdot &\cdot & \cdot \notag \\ \cdot &\cdot& \rho \\ \cdot & -\rho &\cdot \end {array} \right], \notag \end{equation} \begin{equation} \psi_{+} =\frac{1}{2} \left[ \begin {array}{ccc} 0&- \gamma &-\textnormal{i} \gamma \\ \noalign{\medskip} \gamma & \textnormal{i} \bar h_3 & \bar h_3 \\ \noalign{\medskip}\textnormal{i} \gamma &\bar h_3 &-\textnormal{i} \bar h_3 \end {array} \right]\ol{\xi}. \notag \end{equation} \vspace{1mm} Let $\lambda\in\mathbb C^*$ be the auxiliary spectral parameter. \begin{defn} The \textbf{extended Maurer-Cartan form} is the $\liealgebra{g}^{\mathbb C}[\lambda^{-1},\lambda]$-valued 1-form on $\Fh{\infty}$ given by \begin{equation}\label{eq:primitiveform} \psi_{\lambda}:=\lambda \psi_{+}+\psi_0+\lambda^{-1} \psi_-. \end{equation} \end{defn} The extended 1-form $\psi_{\lambda}$ takes values in the Lie algebra $\liealgebra{g}$ when $\lambda$ is a unit complex number. It satisfies the structure equation \begin{equation}\label{eq:primitiveMC} {\rm d}\psi_{\lambda}+\psi_{\lambda}{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\psi_{\lambda}\equiv 0\mod\hat{\rm I}^{(\infty)}. \end{equation} \subsubsection{Formal Killing field} Recall the decomposition of the Lie algebra $\liealgebra{g}^{\mathbb C}$ given in Fig.\ref{fig:Xdecompo0}. By expanding each of the scalar coefficients $\{ \, \textbf{p}, \textbf{b}, \textbf{c}, \textbf{f}, \textbf{a}, \textbf{g}, \textbf{s}, \textbf{t} \,\}$ as a series in $\mathbb C[[\lambda]]$, we give an abridged definition of the formal Killing fields associated with the extended Maurer-Cartan form. \begin{defn}\label{defn:FK} Let $\psi_{\lambda}$, \eqref{eq:primitiveform}, be the extended Maurer-Cartan form. A \textbf{formal Killing field} is a function\footnotemark\footnotetext{We will find that the canonical formal Killing fields to be constructed are defined on the open subset $\Fh{\infty}_{**}=\{ h_3\ne 0, \infty, \; h_j\ne\infty\;\forall j\geq 4\}\subset\Fh{\infty}_{*}.$} \begin{equation} \textbf{X}_{\lambda}: \Fh{\infty} \to\liealgebra{g}^{\mathbb C}[[ \lambda]],\notag \end{equation} such that; \begin{enumerate}[\qquad a)] \item it satisfies the Killing field equation \begin{equation}\label{eq:KillingEquation} {\rm d} \textbf{X}_{\lambda}+[\psi_{\lambda}, \textbf{X}_{\lambda}]\equiv 0\mod\hat{\rm I}^{(\infty)}, \end{equation} \item its components are the formal series in $ \lambda$ given explicitly by \begin{equation}\label{eq:KFcomponents} \begin{array}{rlrl} \textbf{p}&=\sum p^{6k+4}\lambda^{6k+2}, & \textbf{a}&= \sum a^{6k+7}\lambda^{6k+5}, \\ \textbf{b}&=\sum b^{6k+5}\lambda^{6k+3}, & \textbf{g}&= \sum g^{6k+2}\lambda^{6k}, \\ \textbf{c}&=\sum c^{6k+5}\lambda^{6k+3}, &\textbf{s}&=\sum s^{6k+3}\lambda^{6k+1}, \\ \textbf{f}&=\sum f^{6k+6}\lambda^{6k+4}, &\textbf{t}&= \sum t^{6k+3}\lambda^{6k+1}. \end{array} \end{equation} Here the sums are over the integer index $k$ from $0$ to $\infty$. \end{enumerate} \end{defn} \subsubsection{Recursive structure equation} When the Killing field equation \eqref{eq:KillingEquation} is expanded as a series in $\lambda$, it implies the following recursive structure equation. \vspace{1mm} [\textbf{$n$-th equation}] \begin{align}\label{eq:formalKilling_n} {\rm d} p^{6n+4} &=(\textnormal{i}\gamma b^{6n+5}+2\textnormal{i} h_3 c^{6n+5} )\xi + ( \textnormal{i}\gamma s^{6n+3}+2\textnormal{i} \bar h_3 t^{6n+3})\ol{\xi}, \\ {\rm d} b^{6n+5} +\textnormal{i} b^{6n+5}\rho &=\textnormal{i} h_3 f^{6n+6} \xi+ \frac{\textnormal{i}}{2}\gamma p^{6n+4} \ol{\xi}, \notag\\ {\rm d} c^{6n+5} -2 \textnormal{i} c ^{6n+5}\rho&= \textnormal{i} \gamma f^{6n+6}\xi +\textnormal{i} \bar h_3 p^{6n+4} \ol{\xi}, \notag\\ {\rm d} f^{6n+6} -\textnormal{i} f^{6n+6}\rho &=\frac{3\textnormal{i}}{2}\gamma a^{6n+7} \xi +(\textnormal{i} \gamma c^{6n+5}+\textnormal{i}\bar h_3b^{6n+5})\ol{\xi}, \notag\\ {\rm d} a^{6n+7} &=\textnormal{i} \gamma g^{6n+8}\xi + \textnormal{i} \gamma f^{6n+6}\ol{\xi}, \notag\\ {\rm d} g^{6n+8}+\textnormal{i} g^{6n+8}\rho &= (-\textnormal{i} \gamma t^{6n+9}-\textnormal{i} h_3 s^{6n+9})\xi +\frac{3\textnormal{i}}{2}\gamma a^{6n+7}\ol{\xi}, \notag\\ {\rm d} s^{6n+9} -\textnormal{i} s^{6n+9}\rho&= \frac{\textnormal{i}}{2} \gamma p^{6n+10}\xi -\textnormal{i}\bar h_3 g^{6n+8}\ol{\xi}, \notag\\ {\rm d} t^{6n+9}+2 \textnormal{i} t^{6n+9}\rho&=\textnormal{i} h_3 p^{6n+10}\xi-\textnormal{i} \gamma g^{6n+8}\ol{\xi}, \qquad\qquad \quad (\textnormal{mod}\;\hat{\rm I}^{(\infty)}). \notag \end{align} \vspace{2mm} The relevance of this formal structure equation for the analysis of the minimal Lagrangian system lies in the following observation. \begin{lem}\label{lem:formalJacobi} Suppose the coefficients $\{p^{6n+4}, b^{6n+5},\,c^{6n+5}, f^{6n+6}, a^{6n+7}, g^{6n+8}, s^{6n+9}, t^{6n+9} \}$ satisfy the recursive structure equation \eqref{eq:formalKilling_n}. Then, \begin{enumerate}[\qquad a)] \item $p^{6n+4}$ is a pseudo-Jacobi field. \item $a^{6n+7}$ is a Jacobi field. \end{enumerate} \end{lem} The lemma indicates that one may obtain a canonical sequence of (pseudo) Jacobi fields by \emph{solving} the structure equation \eqref{eq:formalKilling_n}. \subsection{Initial analysis}\label{sec:KFinitial} Recall from Exam.\ref{exam:initialdata} that $z_4$ is a pseudo-Jacobi field, and $z_5-\frac{5}{3}z_4^2$ is a Jacobi field. We start the process of solving for the canonical formal Killing fields by determining the first few terms generated by these initial data. By Lem.\ref{lem:delbpoly0}, the coefficients of the resulting formal Killing fields are the elements in the polynomial ring $\mathbb C[z_4, z_5, \, ... \, ]$, up to scaling by the appropriate powers of $h_3^{\frac{1}{3}}$. It turns out that the pair of formal Killing fields generated by $z_4$, and $z_5-\frac{5}{3}z_4^2$ are sufficient to cover all of the infinite sequence of higher-order (pseudo) Jacobi fields. \subsubsection{Case $p^4=z_4$}\label{sec:521} Set $g^2=0.$ By inspection, set $$s^3=-\frac{3\textnormal{i}}{2}\gamma h_3^{-\frac{1}{3}}, \quad t^3= \frac{3\textnormal{i}}{2}h_3^{\frac{2}{3}}.$$ Differentiating this, we get $$p^4=z_4$$ as expected. Solving the equation $\partial_{\ol{\xi}} b^5= \frac{\textnormal{i}}{2}\gamma z_4$, we get $$b^5= -\frac{\textnormal{i}}{3\gamma} h_3^{\frac{1}{3}} (z_5 -\frac{5}{3}z_4^2). $$ From the equation $\partial_{\xi} p^4= \textnormal{i}\gamma b^{5}+2\textnormal{i} h_3 c^{5}$, this implies $$c^5=-\frac{\textnormal{i}}{3}h_3^{-\frac{2}{3}}(z_5 -\frac{7}{6}z_4^2). $$ Successive derivatives of $b^5$ give \begin{align} f^6&= -\frac{1}{3\gamma} h_3^{-\frac{1}{3}} (z_6-\frac{14}{3}z_5 z_4+\frac{35}{9}z_4^3), \notag\\ a^7&= \frac{2\textnormal{i}}{9\gamma^2} (z_7-7 z_6 z_4 -\frac{14}{3} z_5^2+\frac{245}{9}z_5 z_4^2 - \frac{455}{27}z_4^4), \notag\\ g^8&= \frac{2}{9\gamma^3} h_3^{\frac{1}{3}} ( z_{{8}}-{\frac {28}{3}}\,z_{{7}}z_{{4}}-{\frac {49}{3}}\,z_{{6}}z_{{5}} +{\frac {455}{9}}\,z_{{6}} z_{{4}}^{2} +70 z_{{5}}^{2}z_{{4}}-{\frac {5005}{27}}z_{{5}} z_{{4}}^{3} +{\frac {7280}{81}}\, z_{{4}}^{5}). \notag \end{align} Proceed with the similar computation as above by solving (by inspection) the associated $\partial_{\ol{\xi}}$-equation, and we get \begin{align} s^9&=\frac{2\textnormal{i}}{27\gamma^3}h_3^{-\frac{1}{3}}\Big( z_{{9}} -11z_{{8}}z_{{4}}-{\frac {79}{3}}z_{{7}}z_{{5}} +{\frac {689}{9}}z_{{7}}z_{{4}}^{2} -16z_{{6}}^{2} +286z_{{6}}z_{{5}}z_{{4}} -{\frac{3380}{9}}z_{{6}}z_{{4}}^{3} \notag\\ &\qquad\qquad\;\; +{\frac {1976}{27}}z_{{5}}^{3} -{\frac {22360}{27}}z_{{5}}^{2}z_{{4}}^{2}+{\frac {108680}{81}}z_{{5}}z_{{4}}^{4} -{\frac {380380}{729}}z_{{4}}^{6} \Big), \notag \\ t^9&=\frac{4\textnormal{i}}{27\gamma^4}h_3^{\frac{2}{3}} \Big(z_{{9}}-12 z_{{8}}z_{{4}}-{\frac {76}{3}}z_{{7}}z_{{5}}+{\frac {758}{9}}z_{{7}}z_{{4}}^{2} -{\frac {33}{2}}z_{{6}}^{2}+{\frac {901}{3}}z_{{6}}z_{{5}}z_{{4}} -{\frac {3770}{9}}z_{{6}}z_{{4}}^{3}\notag\\ & \qquad\qquad\;\; +{\frac {1847}{27}}z_{{5}}^{3}-{\frac {47255}{54}} z_{{5}}^{2}z_{{4}}^{2} +{\frac {120380}{81}}z_{{5}}z_{{4}}^{4}-{\frac {432250}{729}}z_{{4}}^{6} \Big), \notag \\ p^{10}&=\frac{4}{27\gamma^4}\Big( z_{{10}} -{\frac {43}{3}}z_{{9}}z_{{4}} -{\frac {112}{3}}z_{{8}}z_{{5}}+{\frac {1118}{9}}z_8 z_{{4}}^{2} -{\frac {175}{3}}z_{{7}}z_{{6}}+{\frac {4979}{9}}z_{{7}}z_{{5}}z_{{4}}-{\frac {21164}{27}}z_{{7}}z_{{4}}^{3} \notag\\ & \qquad\quad\;\; +{\frac {1066}{3}}z_{{6}}^{2}z_{{4}} +{\frac {4550}{9}}z_{{6}}z_{{5}}^{2} -{\frac {116324}{27}}z_{{6}}z_{{5}}z_{{4}}^{2} +{\frac {301340}{81}}z_{{6}}z_{{4}}^{4} \notag\\ & \qquad\quad\;\; -{\frac {165776}{81}}z_{{5}}^{3}z_{{4}} +{\frac {286520}{27}}z_{{5}}^{2}z_{{4}}^{3} -{\frac {3151720}{243}}z_{{5}}z_{{4}}^{5} +{\frac {9509500}{2187}}z_{{4}}^{7} \Big). \notag \end{align} By Lem.\ref{lem:formalJacobi}, $a^7$ is a Jacobi field, and $p^{10}$ is a pseudo-Jacobi field. \subsubsection{Case $a^5=z_5-\frac{5}{3}z_4^2$}\label{sec:522} For the formal Killing field generated by the Jacobi field $z_5-\frac{5}{3}z_4^2$, it is convenient to lower the upper indices of the formal Killing field coefficients by 2 to match the order. The resulting structure equation is recorded as follows. \begin{equation}\label{eq:KFcomponents'} \begin{array}{rlrl} \textbf{p}&=\sum p^{6k+2}\lambda^{6k}, & \textbf{a}&= \sum a^{6k+5}\lambda^{6k+3}, \\ \textbf{b}&=\sum b^{6k+3}\lambda^{6k+1}, & \textbf{g}&= \sum g^{6k+6}\lambda^{6k+4}, \\ \textbf{c}&=\sum c^{6k+3}\lambda^{6k+1}, &\textbf{s}&=\sum s^{6k+7}\lambda^{6k+5}, \\ \textbf{f}&=\sum f^{6k+4}\lambda^{6k+2}, &\textbf{t}&= \sum t^{6k+7}\lambda^{6k+5}. \end{array} \end{equation} [\textbf{$n$-th equation}'] \begin{align}\label{eq:formalKilling_n'} {\rm d} p^{6n+2} &=(\textnormal{i}\gamma b^{6n+3}+2\textnormal{i} h_3 c^{6n+3} )\xi + ( \textnormal{i}\gamma s^{6n+1}+2\textnormal{i} \bar h_3 t^{6n+1})\ol{\xi}, \\ {\rm d} b^{6n+3} +\textnormal{i} b^{6n+3}\rho &=\textnormal{i} h_3 f^{6n+4} \xi+ \frac{\textnormal{i}}{2}\gamma p^{6n+2} \ol{\xi}, \notag\\ {\rm d} c^{6n+3} -2 \textnormal{i} c ^{6n+3}\rho&= \textnormal{i} \gamma f^{6n+4}\xi +\textnormal{i} \bar h_3 p^{6n+2} \ol{\xi}, \notag\\ {\rm d} f^{6n+4} -\textnormal{i} f^{6n+4}\rho &=\frac{3\textnormal{i}}{2}\gamma a^{6n+5} \xi +(\textnormal{i} \gamma c^{6n+3}+\textnormal{i}\bar h_3b^{6n+3})\ol{\xi}, \notag\\ {\rm d} a^{6n+5} &=\textnormal{i} \gamma g^{6n+6}\xi + \textnormal{i} \gamma f^{6n+4}\ol{\xi}, \notag\\ {\rm d} g^{6n+6}+\textnormal{i} g^{6n+6}\rho &= (-\textnormal{i} \gamma t^{6n+7}-\textnormal{i} h_3 s^{6n+7})\xi +\frac{3\textnormal{i}}{2}\gamma a^{6n+5}\ol{\xi}, \notag\\ {\rm d} s^{6n+7} -\textnormal{i} s^{6n+7}\rho&= \frac{\textnormal{i}}{2} \gamma p^{6n+8}\xi -\textnormal{i}\bar h_3 g^{6n+6}\ol{\xi}, \notag\\ {\rm d} t^{6n+7}+2 \textnormal{i} t^{6n+7}\rho&=\textnormal{i} h_3 p^{6n+8}\xi-\textnormal{i} \gamma g^{6n+6}\ol{\xi}, \qquad\qquad \quad (\textnormal{mod}\;\hat{\rm I}^{(\infty)}). \notag \end{align} We proceed to solve for the first few terms. Let $p^2=0.$ By inspection, set $$b^3=-\frac{9}{2}\gamma h_3^{\frac{1}{3}}, \quad c^3= \frac{9}{4}\gamma^2 h_3^{-\frac{2}{3}}.$$ Differentiating these equations successively, one gets $$f^4=\frac{3\textnormal{i}}{2}\gamma h_3^{-\frac{1}{3}}z_4,\quad a^5=z_5-\frac{5}{3}z_4^2,\quad g^6 =-\frac{\textnormal{i}}{\gamma} h_3^{\frac{1}{3}} \left( z_6 -5 z_5z_4+\frac{40}{9} z_4^3 \right).$$ Note that $a^5$ is as expected. \iffalse and this implies $$a^5=z_5-\frac{5}{3}z_4^2$$ as expected. Differentiating $a^5$, one gets $$ g^6 =-\frac{\textnormal{i}}{\gamma} h_3^{\frac{1}{3}} \left( z_6 -5 z_5z_4+\frac{40}{9} z_4^3 \right). $$ \fi A similar computation as in the previous case yields, \begin{align} s^7&=\frac{1}{3\gamma}h_3^{-\frac{1}{3}} \Big( z_7 -6z_6z_4-\frac{16}{3} z_5^2+\frac{220}{9} z_5z_4^2-\frac{385}{27} z_4^4 \Big), \notag \\ t^7&=\frac{2}{3\gamma^2}h_3^{\frac{2}{3}}\Big( z_{{7}}-7z_{{6}}z_{{4}}-{\frac {29}{6}}z_{{5}}^{2} +{\frac {250}{9}}z_{{5}}z_{{4}}^{2} -{\frac {935}{54}}z_{{4}}^{4} \Big), \notag \\ p^8&=-\frac{2\textnormal{i}}{3\gamma^2}\Big(z_{{8}}-{\frac {26}{3}}z_{{7}}z_{{4}} -{\frac {50}{3}}z_{{6}}z_{{5}}+{\frac {418}{9}}z_{{6}}z_{{4}}^{2}+{\frac {616}{9}}z_{{5}}^{2}z_{{4}} -{\frac {1540}{9}}z_{{5}}z_{{4}}^{3}+{\frac {6545}{81}}z_{{4}}^{5} \Big),\notag\\ b^9&=-\frac{2}{9\gamma^3}h_3^{\frac{1}{3}} \Big( z_{{9}}-12z_{{8}}z_{{4}} -{\frac {74}{3}}z_{{7}}z_{{5}}+{\frac {748}{9}}z_{{7}}z_{{4}}^{2} -17z_{{6}}^{2}+{\frac {902}{3}}z_{{6}}z_{{5}}z_{{4}} -{\frac {3740}{9}}z_{{6}}z_{{4}}^{3} \notag\\ &\qquad\qquad\;\;\; +{\frac {1760}{27}}z_{{5}}^{3} -{\frac {23320}{27}}z_{{5}}^{2}z_{{4}}^{2} +{\frac {118745}{81}}z_{{5}}z_{{4}}^{4} -{\frac {425425}{729}}z_{{4}}^{6}\Big), \notag \\ c^9&=-\frac{2}{9\gamma^2}h_3^{-\frac{2}{3}} \Big( z_{{9}}-11z_{{8}}z_{{4}}-{\frac {77}{3}}z_{{7}}z_{{5}}+{\frac {682}{9}}z_{{7}}z_{{4}}^{2} -{\frac {33}{2}}z_{{6}}^{2} +286z_{{6}}z_{{5}}z_{{4}} -374z_{{6}}z_{{4}}^{3} \notag \\ &\qquad\qquad\; +{\frac {1892}{27}}z_{{5}}^{3} -{\frac {22066}{27}} z_{{5}}^{2}z_{{4}}^{2} +{\frac {107525}{81}}z_{{5}}z_{{4}}^{4} -{\frac {752675}{1458}}z_{{4}}^{6}\Big).\notag \end{align} Successively differentiating $b^9, c^9$, one finally gets, \begin{align*} f^{10}&=\frac{2\textnormal{i}}{9\gamma^3}h_3^{-\frac{1}{3}}\Big( z_{10}-{\frac {44}{3}}z_{{9}}z_{{4}}-{\frac {110}{3}}z_{{8}}z_{{5}} +{\frac {1144}{9}}z_{{8}}z_{{4}}^{2} -{\frac {176}{3}}z_{{7}}z_{{6}} +{\frac {1672}{3}}z_{{7}}z_{{5}}z_{{4}} -{\frac {21692}{27}}z_{{7}}z_{{4}}^{3} \notag\\ &\qquad\qquad\; +363z_{{6}}^{2}z_{{4}} +{\frac {4466}{9}}z_{{6}}z_{{5}}^{2} -{\frac {118184}{27}}z_{{6}}z_{{5}}z_{{4}}^{2} +{\frac {309485}{81}}z_{{6}}z_{{4}}^{4} -{\frac {164560}{81}}z_{{5}}^{3}z_{{4}} \notag\\ &\qquad\qquad\; +{\frac {871420}{81}}z_{{5}}^{2}z_{{4}}^{3} -{\frac {1075250}{81}}z_{{5}}z_{{4}}^{5} +{\frac {9784775}{2187}}z_{{4}}^{7} \Big), \notag\\ a^{11}&=\frac{4}{27\gamma^4} \Big( z_{11}-{\frac {55}{3}}z_{{10}}z_{{4}}-{\frac {154}{3}}z_{{9}}z_{{5}}+{\frac {1760}{9}}z_{{9}}z_{{4}}^{2} -{\frac {286}{3}}z_{{8}}z_{{6}}+{\frac {2948}{3}}z_{{8}}z_{{5}}z_4 -{\frac {41140}{27}}z_{{8}}z_{{4}}^{3} \\ &\qquad\quad\; -{\frac {176}{3}}z_{{7}}^{2} +{\frac {14014}{9}}z_{{7}}z_{{6}}z_{{4}} +{\frac {9482}{9}}z_{{7}}z_{{5}}^{2} -{\frac {268532}{27}}z_{{7}}z_{{5}}z_{{4}}^{2} +{\frac {247775}{27}}z_{{7}}z_{{4}}^{4} \\ &\qquad\quad\; +{\frac {12199}{9}}z_{{6}}^{2}z_{{5}} -{\frac {173723}{27}}z_{{6}}^{2}z_{{4}}^{2} -{\frac {158950}{9}}z_{{6}}z_{{5}}^{2}z_{{4}} +{\frac {5344460}{81}}z_{{6}}z_{{5}}z_{{4}}^{3} -{\frac {10343905}{243}}z_{{6}}z_{{4}}^{5} \\ &\qquad\quad\; -{\frac {164560}{81}}z_{{5}}^{4} +{\frac {11133980}{243}}z_{{5}}^{3}z_{{4}}^{2} -{\frac {36171410}{243}}z_{{5}}^{2}z_{{4}}^{4} +{\frac {320101925}{2187}}z_{{5}}z_{{4}}^{6} -{\frac {283758475}{6561}}z_{{4}}^{8} \Big).\notag \end{align*} By Lem.\ref{lem:formalJacobi}, $a^5, a^{11}$ are Jacobi fields, and $p^{8}$ is a pseudo-Jacobi field. \subsection{Inductive formulas}\label{sec:KFformulae} Based on the initial analyses given above, we give the differential algebraic inductive formulas for the respective formal Killing fields: \vspace{1mm} \qquad \ref{sec:p4z4})\quad Case $p^4=z_4$, \qquad \ref{sec:a5z5})\quad Case $a^5=z_5-\frac{5}{3}z_4^2$. \vspace{1mm} Note from Eqs.\eqref{eq:formalKilling_n}, \eqref{eq:formalKilling_n'} that one needs to solve for the coefficients $\{ s^*, t^*\}$, and $\{b^*, c^*\}$. \iffalse For the sake of notation, let $$\widetilde{\mathcal R}=\mathbb C[ p^*, b^*, c^*, f^*, a^*, g^*, s^*, t^* ] $$ denote the polynomial ring. The construction shows that in fact, up to scaling by powers of $h_3^{\frac{1}{3}}$, the resulting formal Killing field coefficients are elements in the polynomial ring $\mathbb{Q}[z_4, z_5, \, ... ]$ with rational coefficients. \fi \subsubsection{Case $p^4=z_4$}\label{sec:p4z4} Assume the initial data from \S\ref{sec:521}. \vspace{2mm}\noindent \textbf{[Formulas for $\,s^{6n+3}, t^{6n+3}$]}.\, Suppose all the coefficients up to $g^{6n+2}$ are known, $n\geq 1$. We give a formula for $\{s^{6n+3}, t^{6n+3}\}.$ Set the truncated formal Killing field $$ \textbf{X}_{6n+2}:= \left[ \begin {array}{ccc} -2\textnormal{i} \textbf{a} &\textbf{b}+\textbf{f}+\textbf{g}-\textbf{s} &\textnormal{i} \textbf{b}- \textnormal{i} \textbf{f}+\textnormal{i} \textbf{g}+\textnormal{i} \textbf{s}\\ \noalign{\medskip}-\textbf{b}+\textbf{f}+\textbf{g}+\textbf{s} & \textnormal{i} \textbf{c}+\textnormal{i} \textbf{a}-\textnormal{i} \textbf{t}&-\textbf{p}+\textbf{c}+\textbf{t}\\ \noalign{\medskip}-\textnormal{i} \textbf{b}-\textnormal{i} \textbf{f} +\textnormal{i} \textbf{g}-\textnormal{i} \textbf{s} &\textbf{p}+\textbf{c}+\textbf{t}&-\textnormal{i} \textbf{c}+\textnormal{i} \textbf{a}+\textnormal{i} \textbf{t}\end {array} \right], $$ where \begin{equation}\label{eq:components1} \begin{array}{rlrl} \textbf{p}&=\sum_{k=0}^{n} p^{6k+4}\lambda^{6k+2}, & \textbf{a}&= \sum_{k=0}^{n-1} a^{6k+7}\lambda^{6k+5}, \\ \textbf{b}&=\sum_{k=0}^{n-1} b^{6k+5}\lambda^{6k+3}, & \textbf{g}&= \sum_{k=0}^{n} g^{6k+2}\lambda^{6k}, \\ \textbf{c}&=\sum_{k=0}^{n-1} c^{6k+5}\lambda^{6k+3}, &\textbf{s}&=\sum_{k=0}^{n} s^{6k+3}\lambda^{6k+1}, \\ \textbf{f}&=\sum_{k=0}^{n-1} f^{6k+6}\lambda^{6k+4}, &\textbf{t}&= \sum_{k=0}^{n} t^{6k+3}\lambda^{6k+1}. \end{array} \end{equation} Here the unknown coefficients are $s^{6n+3}, t^{6n+3}, p^{6n+4}.$ The determinant is given by $$ \det(\textbf{X}_{6n+2}) =\textnormal{i}( 4\textbf{gsp} -4 \textbf{fga}-4\textbf{b$^2$c} -4\textbf{f$^2$t}+4\textbf{g$^2$c} +4\textbf{s$^2$t}+2\textbf{a$^3$}-2\textbf{ap$^2$}+8\textbf{act}-4\textbf{bsa}+4\textbf{bfp}) . $$ Expanding as a series in $\lambda$, let us denote $$\det(\textbf{X}_{6n+2}):=\sum_{j=0}^{3n} \textbf{x}_{6n+2}^{6j+3} \lambda^{6j+3}. $$ Consider now the derivative $$\partial_{\ol{\xi}}(\det(\textbf{X}_{6n+2})).$$ The structure equation shows that this term stems from the absence of $b^{6n+5}, c^{6n+5}$-terms in $\textbf{X}_{6n+2}$. Hence only the terms that contain $p^{6n+4}$ contribute to $\partial_{\ol{\xi}}(\det(\textbf{X}_{6n+2}))$. From the determinant formula above, one finds by checking the $\lambda$-degree that $$\partial_{\ol{\xi}} \textbf{x}_{6n+2}^{6j+3}=0, \quad \textnormal{for}\; j\leq n. $$ By Cor.\ref{cor:lemma5.4} and weighted homogeneity, this implies that $$ \textbf{x}_{6n+2}^{6j+3}=0, \quad \textnormal{for}\; j\leq n. $$ Consider the term $\textbf{x}_{6n+2}^{6n+3}$ of the highest $\lambda$-degree among these. We have \begin{align} \label{eq:x6n+43} \textbf{x}_{6n+2}^{6n+3} &= 4\textnormal{i}(2s^3 t^3 s^{6n+3}+(s^3)^2t^{6n+3})+\textbf{y}_{6n+2}^{6n+3}, \end{align} where $\textbf{y}_{6n+2}^{6n+3} \in\mathcal O(6n+2)$. On the other hand, we have \begin{equation}\label{eq:delxg6n+2} \partial_{\xi} g^{6n+2}=-\textnormal{i} h_3 s^{6n+3}-\textnormal{i}\gamma t^{6n+3}. \end{equation} Combining \eqref{eq:x6n+43}, \eqref{eq:delxg6n+2}, one gets \begin{align}\label{eq:st6n+3} s^{6n+3}&= \frac{\textnormal{i}}{27\gamma}h_3^{-1}\left(9\gamma \partial_{\xi} g^{6n+2} + h_3^{\frac{2}{3}} \textbf{y}_{6n+2}^{6n+3} \right), \\ t^{6n+3}&= \frac{\textnormal{i}}{27\gamma^2}\left(18\gamma \partial_{\xi} g^{6n+2} - h_3^{ \frac{2}{3}} \textbf{y}_{6n+2}^{6n+3} \right). \notag \end{align} \vspace{2mm}\noindent \textbf{[Formulas for $b^{6n-1}, c^{6n-1}$]}.\, Suppose all the coefficients up to $p^{6n-2}$ are known, $n\geq 1$. We give a formula for $\{b^{6n-1}, c^{6n-1}\}.$ \vspace{1mm} Given the truncated formal Killing field $\textbf{X}_{6n+2}$ as above, let $$ \det(\mu \textnormal{I}_3+\textbf{X}_{6n+2})=\mu^3+\sigma_2(\textbf{X}_{6n+2})\mu+\det(\textbf{X}_{6n+2}) $$ be the characteristic polynomial. For the case at hand, we utilize $\sigma_2(\textbf{X}_{6n+2}).$ It is given by the formula $$\sigma_2(\textbf{X}_{6n+2}) = 3\textbf{a}^2+\textbf{p}^2-4\textbf{c}\textbf{t}-4\textbf{b}\textbf{s}-4\textbf{f}\textbf{g}. $$ Expanding as a series in $\lambda$, let us denote $$\sigma_2(\textbf{X}_{6n+2}):=\sum_{j=0}^{2n} \textbf{x}_{6n+2}^{6j+4} \lambda^{6j+4}. $$ Consider the derivative $\partial_{\ol{\xi}}( \sigma_2(\textbf{X}_{6n+2}) )$. By the similar argument as above, one finds that $$\partial_{\ol{\xi}} \textbf{x}_{6n+2}^{6j-2}=0, \quad \textnormal{for}\; j\leq n, $$ and hence $$ \textbf{x}_{6n+2}^{6j-2}=0, \quad \textnormal{for}\; j\leq n. $$ \vspace{1mm} Consider the term $\textbf{x}_{6n+2}^{6n-2}$. Then \begin{align} \label{eq:x6n-6} \textbf{x}_{6n+2}^{6n-2}&= -4 s^3 b^{6n-1} -4 t^3c^{6n-1} +\textbf{y}_{6n+2}^{6n-2}. \end{align} Here $\textbf{y}_{6n+2}^{6n-2} \in\mathcal O(6n-2)$. On the other hand, we have \begin{equation}\label{eq:delxp6n+4} \partial_{\xi} p^{6n-2}= \textnormal{i} \gamma b^{6n-1}+2\textnormal{i} h_3c^{6n-1}. \end{equation} Combining \eqref{eq:x6n-6}, \eqref{eq:delxp6n+4}, one gets \begin{align}\label{eq:bc6n-1} b^{6n-1}&= \frac{\textnormal{i}}{9\gamma}\left(-3\partial_{\xi} p^{6n-2} + h_3^{\frac{1}{3}} \textbf{y}^{6n-2}_{6n+2} \right), \\ c^{6n-1}&= -\frac{\textnormal{i}}{18}h_3^{-1}\left( 6\partial_{\xi} p^{6n-2} + h_3^{ \frac{1}{3}} \textbf{y}^{6n-2}_{6n+2} \right). \notag \end{align} \begin{thm}\label{thm:FKformulaep4} Given the ansatz $p^4=z_4$ and the initial data described in \S\ref{sec:521}; \begin{enumerate}[\qquad a)] \item there exists a $\liealgebra{g}^{\mathbb C}[[\lambda]]$-valued canonical formal Killing field $\textbf{X}(p^4)$ which extends these data. The coefficients of its components are generated by the structure equation \eqref{eq:formalKilling_n}, and the differential algebraic inductive formulas \eqref{eq:st6n+3},\eqref{eq:bc6n-1}. Equivalently, $\textbf{X}(p^4)$ is determined by the constraint, $$\det(\mu \textnormal{I}_3+\textbf{X}(p^4))=\mu^3+\left( \frac{27}{2}\gamma^2\right)\lambda^3.$$ Here $\textnormal{I}_3$ denotes the 3-by-3 identity matrix. \item each coefficient of $\textbf{X}(p^4)$ is an element in the polynomial ring $\mathbb C[z_4, z_5, \, ... \, ]$ up to scaling by appropriate powers of $h_3^{\frac{1}{3}}$. \end{enumerate} \end{thm} \begin{cor}\label{cor:FKformulaep4} Given the formal Killing field $\textbf{X}(p^4)$, the sequence of coefficients $$\{\, p^{6n+4}, \overline{p}^{6n+4} \,\}_{n=0}^{\infty}$$ are distinct higher-order pseudo-Jacobi fields, and the sequence of coefficients $$\{\, a^{6n+7}, \overline{a}^{6n+7} \,\}_{n=0}^{\infty}$$ are distinct higher-order Jacobi fields. \end{cor} \subsubsection{Case $a^5=z_5-\frac{5}{3}z_4^2$}\label{sec:a5z5} Recall that we follow \eqref{eq:KFcomponents'}, and \eqref{eq:formalKilling_n'}. Assume the initial data from \S\ref{sec:522}. By the same analysis as in \S\ref{sec:p4z4}, we obtain the corresponding formal Killing field $\textbf{X}(a^5)$. \iffalse \vspace{2mm}\noindent \textbf{[Formulae for $b^{6n+3}, c^{6n+3}$]}.\; Suppose all the coefficients up to $p^{6n+2}$ are known, $n\geq 1$. We give a formula for $\{b^{6n+3}, c^{6n+3}\}.$ Define the truncated formal Killing field $$ \textbf{X}_{6n+5}= \left[ \begin {array}{ccc} -2\textnormal{i} \textbf{a} &\textbf{b}+\textbf{f}+\textbf{g}-\textbf{s} &\textnormal{i} \textbf{b}- \textnormal{i} \textbf{f}+\textnormal{i} \textbf{g}+\textnormal{i} \textbf{s}\\ \noalign{\medskip}-\textbf{b}+\textbf{f}+\textbf{g}+\textbf{s} & \textnormal{i} \textbf{c}+\textnormal{i} \textbf{a}-\textnormal{i} \textbf{t}&-\textbf{p}+\textbf{c}+\textbf{t}\\ \noalign{\medskip}-\textnormal{i} \textbf{b}-\textnormal{i} \textbf{f} +\textnormal{i} \textbf{g}-\textnormal{i} \textbf{s} &\textbf{p}+\textbf{c}+\textbf{t}&-\textnormal{i} \textbf{c}+\textnormal{i} \textbf{a}+\textnormal{i} \textbf{t}\end {array} \right], $$ where \begin{equation}\label{eq:components2} \begin{array}{rlrl} \textbf{p}&=\sum_{k=1}^{n} p^{6k+2}\lambda^{6k}, & \textbf{a}&= \sum_{k=0}^{n} a^{6k+5}\lambda^{6k+3}, \\ \textbf{b}&=\sum_{k=0}^{n} b^{6k+3}\lambda^{6k+1}, & \textbf{g}&= \sum_{k=0}^{n-1} g^{6k+6}\lambda^{6k+4}, \\ \textbf{c}&=\sum_{k=0}^{n} c^{6k+3}\lambda^{6k+1}, &\textbf{s}&=\sum_{k=0}^{n-1} s^{6k+7}\lambda^{6k+5}, \\ \textbf{f}&=\sum_{k=0}^{n} f^{6k+4}\lambda^{6k+2}, &\textbf{t}&= \sum_{k=0}^{n-1} t^{6k+7}\lambda^{6k+5}. \end{array} \end{equation} Here the unknown coefficients are $\{ \,b^{6n+3}, c^{6n+3}, f^{6n+4}, a^{6n+5} \,\}.$ Recall the determinant formula $$ \det(\textbf{X}_{6n+5}) =\textnormal{i}( 4\textbf{gsp} -4 \textbf{fga}-4\textbf{b$^2$c} -4\textbf{f$^2$t}+4\textbf{g$^2$c} +4\textbf{s$^2$t}+2\textbf{a$^3$}-2\textbf{ap$^2$}+8\textbf{act}-4\textbf{bsa}+4\textbf{bfp}) . $$ Expanding as a series in $\lambda$, let us denote $$\det(\textbf{X}_{6n+5}):=\sum_{j=0}^{3n+1} \textbf{x}_{6n+5}^{6j+3} \lambda^{6j+3}. $$ Consider now the derivative $\partial_{\ol{\xi}}(\det(\textbf{X}_{6n+5}))$. The structure equation shows that this term stems from the absence of $g^{6n+6}$-term in $\textbf{X}_{6n+5}$. Hence, only the terms that contain $a^{6n+5}$ contribute to $\partial_{\ol{\xi}}(\det(\textbf{X}_{6n+5}))$. From the determinant formula above, one finds again by checking the $\lambda$-degree that $$\partial_{\ol{\xi}} \textbf{x}_{6n+5}^{6j+3}=0, \quad \textnormal{for}\; j\leq n. $$ By Cor.\ref{cor:lemma5.4} and weighted homogeneity, this implies $$ \textbf{x}_{6n+5}^{6j+3}=0, \quad \textnormal{for}\; j\leq n. $$ Choose the term $\textbf{x}_{6n+5}^{6n+3}$ of the highest $\lambda$-degree among these. Then \begin{align} \label{eq:x6n+53} \textbf{x}_{6n+5}^{6n+3} &= -4\textnormal{i}(2b^3 c^3 b^{6n+3}+(b^3)^2 c^{6n+3}) +\textbf{y}_{6n+5}^{6n+3}, \\ &= 81 \textnormal{i} \gamma^2 h_3^{-\frac{1}{3}}(\gamma b^{6n+3}-h_3 c^{6n+3}) +\textbf{y}_{6n+5}^{6n+3},\notag \end{align} where $\textbf{y}_{6n+5}^{6n+3} \in\mathcal O(6n+2)\cap\widetilde{\mathcal R}$. On the other hand, we have \begin{equation}\label{eq:delxp6n+2} \partial_{\xi} p^{6n+2}= \textnormal{i}\gamma b^{6n+3}+2\textnormal{i} h_3 c^{6n+3}. \end{equation} Combining \eqref{eq:x6n+53}, \eqref{eq:delxp6n+2}, one gets \begin{align}\label{eq:bc6n+3} b^{6n+3}&= - \frac{\textnormal{i}}{243\gamma^3}\left(81\gamma^2 \partial_{\xi} p^{6n+2} -2 h_3^{\frac{1}{3}} \textbf{y}_{6n+5}^{6n+3} \right), \\ c^{6n+3}&= - \frac{\textnormal{i}}{243\gamma^2} h_3^{-1} \left(81\gamma^2 \partial_{\xi} p^{6n+2} + h_3^{\frac{1}{3}} \textbf{y}_{6n+5}^{6n+3} \right). \notag \end{align} \vspace{2mm}\noindent \textbf{[Formulae for $s^{6n+1}, t^{6n+1}$]}.\, Given the truncated formal Killing field $\textbf{X}_{6n+5}$ as above, let $$ \det(\mu \textnormal{I}_3-\textbf{X}_{6n+5})=\mu^3+\sigma_2(\textbf{X}_{6n+5})\mu-\det(\textbf{X}_{6n+5}) $$ be the characteristic polynomial. Recall the formula $$\sigma_2(\textbf{X}_{6n+5}) = 3\textbf{a}^2+\textbf{p}^2-4\textbf{c}\textbf{t}-4\textbf{b}\textbf{s}-4\textbf{f}\textbf{g}. $$ Expanding $\sigma_2(\textbf{X}_{6n+5})$ as a series in $\lambda$, let us denote $$\sigma_2(\textbf{X}_{6n+5}):=\sum_{j=1}^{2n+1} \textbf{x}_{6n+5}^{6j} \lambda^{6j}. $$ Consider the derivative $\partial_{\ol{\xi}}( \sigma_2(\textbf{X}_{6n+5}) )$. Similarly as in the previous case, a computation shows that $$\partial_{\ol{\xi}} \textbf{x}_{6n+5}^{6j}=0, \quad \textnormal{for}\; j\leq n, $$ and hence $$ \textbf{x}_{6n+5}^{6j}=0, \quad \textnormal{for}\; j\leq n. $$ Choose the term $\textbf{x}_{6n+5}^{6n}$ of the highest $\lambda$-degree among these. Then \begin{align} \label{eq:x6n} \textbf{x}_{6n+5}^{6n}&= -4 b^3 s^{6n+1} -4 c^3 t^{6n+1} +\textbf{y}_{6n+5}^{6n}, \\ &=-9\gamma h_3^{-\frac{2}{3}}(-2h_3 s^{6n+1}+\gamma t^{6n+1})+\textbf{y}_{6n+5}^{6n}.\notag \end{align} Here $\textbf{y}_{6n+5}^{6n} \in\mathcal O(6n)\cap\widetilde{\mathcal R}$. On the other hand, one gets \begin{equation}\label{eq:delxg6n} \partial_{\xi} g^{6n}=-\textnormal{i} h_3 s^{6n+1}-\textnormal{i}\gamma t^{6n+1}. \end{equation} Combining \eqref{eq:x6n}, \eqref{eq:delxg6n}, we get \begin{align}\label{eq:st6n+1} s^{6n+1}&=- \frac{1}{27\gamma}h_3^{-1} \left(-9\textnormal{i}\gamma \partial_{\xi} g^{6n} + h_3^{\frac{2}{3}} \textbf{y}_{6n+5}^{6n} \right), \\ t^{6n+1}&=- \frac{\textnormal{i} }{27\gamma^2} \left(-18 \gamma \partial_{\xi} g^{6n} + \textnormal{i} h_3^{\frac{2}{3}} \textbf{y}_{6n+5}^{6n} \right). \notag \end{align} \fi \begin{thm}\label{thm:FKformulaea5} Given the ansatz $a^5=z_5-\frac{5}{3}z_4^2$ and the initial data described in \S\ref{sec:522}; \begin{enumerate}[\qquad a)] \item there exists a $\liealgebra{g}^{\mathbb C}[[\lambda]]$-valued canonical formal Killing field $\textbf{X}(a^5)$ which extends these data. The coefficients of its components are determined by the structure equation \eqref{eq:formalKilling_n'}, and the constraint, $$\det(\mu \textnormal{I}_3+\textbf{X}(a^5))=\mu^3-\left(\frac{729}{4}\textnormal{i}\gamma^4 \right)\lambda^3.$$ \item each coefficient of $\textbf{X}(a^5)$ is an element in the polynomial ring $\mathbb C[z_4, z_5, \, ... \, ]$ up to scaling by appropriate powers of $h_3^{\frac{1}{3}}$. \end{enumerate} \end{thm} \begin{cor}\label{cor:FKformulaea5} Given the formal Killing field $\textbf{X}(a^5)$, the sequence of coefficients $$\{\, a^{6n+5}, \overline{a}^{6n+5} \,\}_{n=0}^{\infty}$$ are distinct higher-order Jacobi fields, and the sequence of coefficients $$\{\, p^{6n+8}, \overline{p}^{6n+8} \,\}_{n=0}^{\infty}$$ are distinct higher-order pseudo-Jacobi fields. \end{cor} \section{Higher-order conservation laws}\label{sec:highercvlaws1} Recall that the classical conservation laws are defined as the elements in the 1-st characteristic cohomology of the quotient complex $$(\Omega^*(X)/\mathcal I, \underline{{\rm d}}).$$ Generalizing this, the conservation laws of the minimal Lagrangian system are defined as the elements in the 1-st characteristic cohomology of the quotient complex of the infinitely prolonged differential system $(\hat{X}^{(\infty)},\hat{\rm I}^{(\infty)})$, $$(\Omega^*(\hat{X}^{(\infty)})/\hat{\rm I}^{(\infty)}, \underline{{\rm d}}),$$ where $\underline{{\rm d}}={\rm d}\mod\hat{\rm I}^{(\infty)}.$ In this section, we give a description of the infinite sequence of higher-order conservation laws generated by the canonical formal Killing fields $\textbf{X}(p^4), \textbf{X}(a^5).$ \subsection{Definition} Let $(\Omega^*(\hat{X}^{(\infty)}), {\rm d})$ be the de-Rham complex of $\mathbb C$-valued differential forms on $\hat{X}^{(\infty)}$. Let $$(\underline{\Omega}^*=\Omega^*(\hat{X}^{(\infty)})/\hat{\rm I}^{(\infty)}, \underline{{\rm d}})$$ be the quotient space equipped with the induced differential $\underline{{\rm d}}={\rm d}\mod\hat{\rm I}^{(\infty)}$. The prolongation sequence of Pfaffian systems $\Ih{k}$ satisfy the inductive closure conditions $${\rm d}\Ih{k}\equiv 0\mod \Ih{k+1}, \; k\geq 1.$$ It follows that $\hat{\rm I}^{(\infty)}=\cup_{k=0}^{\infty}\Ih{k}$ is formally Frobenius, and $(\underline{\Omega}^*, \underline{{\rm d}})$ becomes a complex. Let $H^{q}(\underline{\Omega}^*,\, \underline{{\rm d}})$ be the cohomology at $\underline{\Omega}^q.$ The set $$\{ \;H^{q}(\underline{\Omega}^*,\, \underline{{\rm d}}) \;\}_{q=0}^2$$ is called the \emph{characteristic cohomology} of the differential system $(\hat{X}^{(\infty)},\hat{\rm I}^{(\infty)})$. \begin{defn} Let $(\hat{X}^{(\infty)},\hat{\rm I}^{(\infty)})$ be the triple cover of the infinite prolongation of the differential system for minimal Lagrangian surfaces. A \textbf{conservation law} is an element of the 1-st characteristic cohomology $H^1(\underline{\Omega}^*,\, \underline{{\rm d}})$ of $(\hat{X}^{(\infty)},\hat{\rm I}^{(\infty)})$. The $\mathbb C$-vector space of conservation laws is denoted by \[ \mathcal C^{(\infty)}:=H^1(\underline{\Omega}^*,\, \underline{{\rm d}}).\] Let $\mathcal C^{(\infty)}_{loc}$ denote the space of local conservation laws of $\hat{\rm I}^{(\infty)}$ restricted to a small contractible open subset of $\hat{X}^{(\infty)}$. \end{defn} For simplicity, we shall suppress the global issues and identify $\mathcal C^{(\infty)}\simeq\mathcal C^{(\infty)}_{loc}.$ Note by definition that the classical conservation laws $\mathcal C^{(0)}\subset\mathcal C^{(\infty)}.$ \subsubsection{Spectral sequence} Consider the filtration by the subspaces $$ F^p\Omega^q =\textnormal{Image}\{\underbrace{\hat{\rm I}^{(\infty)}{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\hat{\rm I}^{(\infty)}\, ... }_{p}{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}:\Omega^*(\hat{X}^{(\infty)})\to\Omega^{q}(\hat{X}^{(\infty)})\}.$$ From the associated graded $F^p\Omega^*/F^{p+1}\Omega^*,$ a standard construction yields the spectral sequence $$(E^{p,q}_r, {\rm d}_r), \quad {\rm d}_r \;\textnormal{has bidegree}\;(r, 1-r), \quad r\geq 0.$$ From the fundamental theorem \cite[p562, Theorem 2 and Eq.(4)]{Bryant1995}, at least locally the following sub-complex is exact, \begin{equation}\label{eq:exactsequence} 0\to E^{0,1}_1\hookrightarrow E^{1,1}_1\to E^{2,1}_1. \end{equation} Here, by definition, the first piece is given by \begin{align* E^{0,1}_1 &=\{ \varphi\in\Omega^1(\hat{X}^{(\infty)}) \vert {\rm d}\varphi\equiv 0\mod\hat{\rm I}^{(\infty)}\}/ \{ {\rm d}\Omega^0(\hat{X}^{(\infty)}) +\Omega^1(\hat{\rm I}^{(\infty)})\} \\ &=H^1(\Omega^*(\hat{X}^{(\infty)})/\hat{\rm I}^{(\infty)}, \underline{{\rm d}})\notag\\ &=\mathcal C^{(\infty)},\notag \end{align*} and it is the space of conservation laws. \subsubsection{$E^{1,1}_1$}\label{sec:cvsymbol} The second piece $E^{1,1}_1$ is called the space of cosymmetries. In the present case, the differential system is formally self-adjoint and this is the space of Jacobi fields, $$E^{1,1}_1=\mathfrak J^{(\infty)}.$$ The differential ${\rm d}_1: E^{0,1}_1\hookrightarrow E^{1,1}_1$ can be considered as the symbol map for conservation laws. The space $E^{1,1}_1$ admits the following analytic description. Let $\Phi$ be a 2-form which represents a class in $E^{1,1}_1$. By definition, one may write $$\Phi\equiv A \Psi -\theta_0 {\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge} \sigma \mod F^2\Omega^2, $$ for a scalar coefficient $A$ and a 1-form $\sigma$, where $$\Psi=\textnormal{Im}(\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi)=-\frac{\textnormal{i}}{2}(\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi-\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}). $$ Recall \begin{align}\label{1dPsi} {\rm d}\Psi&= 3\textnormal{i}\gamma^2\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}(\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}+\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\theta}_1) \\ &\equiv 0\mod\theta_0. \notag \end{align} We wish to show that the coefficient $A$ is a Jacobi field. Differentiating $\Phi$, one gets $$0\equiv {\rm d} A{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\Psi-{\rm d}\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\sigma \mod \theta_0, F^2\Omega^3. $$ Since ${\rm d}\theta_0=-\frac{1}{2}(\theta_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi+\ol{\theta}_1{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi})$, this implies that $$\sigma\equiv - \textnormal{i}\left((\partial_{\xi} A)\xi-(\partial_{\ol{\xi}} A)\ol{\xi}\right)\mod\hat{\rm I}^{(\infty)}. $$ With the given $\sigma$, the coefficient of $\theta_0{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\xi{\mathchoice{\,{\scriptstyle\wedge}\,}{{\scriptstyle\wedge}\ol{\xi}$-term in ${\rm d}\Phi$ then shows that $$\mathcal E(A)=0$$ and $A$ is a Jacobi field. We thus have the isomorphism $E^{1,1}_1\simeq \mathfrak J^{(\infty)}.$ \subsection{Conservation laws from formal Killing fields}\label{sec:cvlawfromFK} A question arises as to if the infinite sequence of Jacobi fields for the minimal Lagrangian system indeed correspond to the sequence of conservation laws, i.e., if the symbol map $${\rm d}_1: \mathcal C^{(\infty)}_{loc}=E^{0,1}_1\hookrightarrow E^{1,1}_1$$ is surjective and a higher-order version of Noether's theorem holds for the minimal Lagrangian system. We show that, from the two formal Killing fields constructed in the previous section, we are able to assemble an infinite sequence of higher-order conservation laws. It is likely that they are nontrivial and, considering their spectral weights, the higher-order Noether's theorem holds for the minimal Lagrangian system. \vspace{2mm}\noindent [\textbf{Formal Killing field $\textbf{X}(p^4)$}] Recall the structure equation \eqref{eq:formalKilling_n}. Set \begin{equation}\label{eq:varphin} \varphi_n:=b^{6n+5}\xi+s^{6n+3}\ol{\xi}. \end{equation} The structure equation shows that \[ {\rm d}\varphi_n\equiv 0 \mod\hat{\rm I}^{(\infty)}, \] and $\varphi_n$ represents a conservation law. \vspace{2mm}\noindent [\textbf{Formal Killing field $\textbf{X}(a^5)$}] Recall the structure equation \eqref{eq:formalKilling_n'}. Set \begin{equation}\label{eq:varphin'} \varphi'_n:=b^{6n+3}\xi+s^{6n+1}\ol{\xi}. \end{equation} The structure equation shows that \[ {\rm d}\varphi'_n\equiv 0 \mod\hat{\rm I}^{(\infty)}, \] and $\varphi'_n$ represents a conservation law. \begin{thm}\label{thm:FKcvlaw} Let $\textbf{X}(p^4), \textbf{X}(a^5)$ be the formal Killing fields generated from the initial data $p^4=z_4, a^5=z_5-\frac{5}{3}z_4^2$ respectively in \S\ref{sec:formalKilling}. Then the associated sequence of 1-forms $$\varphi_n, \; \varphi'_n, \quad n=0, 1, 2, \, ... \, ,$$ represent the higher-order conservation laws. \end{thm} It remains to verify that these conservation laws are indeed nontrivial. But, Thm.\ref{thm:FKcvlaw} points to the relevant questions such as periods, residues, and the related application of the higher-order conservation laws to the global problems for minimal Lagrangian surfaces. \begin{rem} Suppose the conservation laws $[\varphi_n], [\varphi_n']$ are nontrivial. Then, by an analysis of the differential ${\rm d} _1$ as in \S\ref{sec:cvsymbol}, the spectral weight count shows that, \[ {\rm d}_1([ \varphi_n]) =a^{6n+7},\quad {\rm d}_1([ \varphi_n']) =a^{6n+5}, \] up to constant scale, where ${\rm d}_1$ is the symbol map ${\rm d}_1:\mathcal C^{(\infty)}_{loc}\to E^{1,1}_1\simeq \mathfrak J^{(\infty)}$. \end{rem} \iffalse \section{Remarks}\label{sec:remarks} There is one relevant problem unanswered: \emph{nontriviality of $[\varphi_n], [\varphi'_n]$}. Combined with the argument above and Thm.\ref{thm:classifyJacobi}, the nontriviality would imply the Noether's theorem for the minimal Lagrangian system, i.e., the injective symbol map $E^{0,1}_1\hookrightarrow E^{1,1}_1$ in \eqref{eq:exactsequence} is also surjective and there exists a canonical isomorphism between the symmetries (Jacobi fields) and conservation laws. \vspace{2mm} We first give an idea for a proof via minimal Lagrangian surfaces of finite type in $\mathbb C\mathbb P^2$. This may be applied to show the nontriviality of the conservation laws $[\varphi_n], [\varphi'_n]$ for even $n$ for the case $\gamma^2>0$. Consider $\varphi_n$. Note first by Lem.\ref{lem:delbpoly0} that if $\varphi_n$ is (locally) trivial and there exists a function $f$ such that $${\rm d} f\equiv \varphi_n\mod\hat{\rm I}^{(\infty)},$$ then $f\in\mathcal R=\mathbb C[z_4,z_5, \, ... \, ]$ and $f$ is globally defined on $\hat{X}^{(\infty)}_*,$ away from the umbilic divisor $\{ \, h_3=0\, \}$. This allows one to work globally. From \cite{Carberry2004}, there exist many minimal Lagrangian tori equipped with polynomial Killing fields. This translates to our setting the existence of admissible Bonnet data defined on tori such that $$p^{6n+4}=0 $$ for even $n$. All the higher-order formal Killing field coefficients vanish and $\textbf{X}(p^4)$ reduces to a polynomial Killing field. Note in this case $b^{6n+5}=0$ and $\varphi_n=s^{6n+3}\ol{\xi}.$ Note also from the structure equation that $\partial_{\xi} s^{6n+3}=0$. Since the underlying torus is compact, this implies $$\varphi_n=\ell_n \overline{\omega} $$ is a constant multiple of $\overline{\omega}$, the complex conjugate of the cube root of Hopf differential. Generically $\ell_n\ne0$. Since $\overline{\omega}$ has periods, $\varphi_n$ restricted to the torus is not exact. We conclude that $[\varphi_n]$ is a nontrivial conservation law. \vspace{2mm} It is more plausible to show the nontrviality by finding the 1-dimensional integral cycles in $\hat{X}^{(\infty)}_*$ which has nonzero periods for $\varphi_n, \varphi'_n.$ \fi \providecommand{\MR}[1]{} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
\section{Introduction} The existence of fake Lax pairs is well known but the phenomenon is not widely understood. The term Lax pair can refer to several different types of systems, the aspect these all have in common is that a pair of linear equations (an overdetermined system) is associated with a nonlinear integrable system through a compatibility condition. The most important property of a real Lax pair is that it proves the integrability of the associated nonlinear system. Indeed, real Lax pairs are very useful for finding information about the solutions to nonlinear integrable systems. Fake Lax pairs, on the other hand, say nothing about the integrability of the associated nonlinear system. Fake Lax pairs often appear very similar to their real counterparts and experts in the area of integrable systems continue to inadvertently publish fake Lax pairs that they believe are real (see sections \ref{sec:hone} and \ref{sec:gramani} for examples). As such, there is serious need for a straightforward method to distinguish between real and fake Lax pairs. In this letter, we provide two very simple methods to identify fake Lax pairs. Lax pairs can be discrete or continuous, matrix or scalar, used for inverse scattering or isomonodromy, and fake Lax pairs reside in all of these categories. The methods to identify fake Lax pairs presented in this article is applicable to any kind of Lax pair that contains a non-removable spectral variable. The question of inserting a non-removable spectral variable into a Lax pair that does not have one is explored in \cite{m10}. The methods outlined in this article highlight properties that are sufficient for a Lax pair to be fake, but are not necessary. It is possible that a Lax pair could pass these tests and still be fake, however, every fake Lax pair known to the authors fails both tests. Although there are many references to fake Lax pairs in the literature, the most famous being \cite{cn91}, there are fewer articles that set out to explain what fake Lax pairs are and how to identify them. So far the methods given to identify fake Lax pairs \cite{lsz90, lll10, m04, s01, s02} have been limited in their applicability and often difficult to apply. In contrast, the tests given here are easily comprehensible and widely applicable. This article is organised as follows. In section \ref{sec:id} we briefly describe two methods to identify fake Lax pairs. The methods are best understood using examples so in section \ref{sec:examples} we give several of those taken from the literature. The article terminates with a conclusion in section \ref{sec:conclusion}. \section{Two methods to identify fake Lax pairs}\label{sec:id} In this section we describe the methods used to identify fake Lax pairs (FLPs). The use of these methods is illustrated by way of examples in section \ref{sec:examples} below. \subsection{Method by removal of the dependent variable}\label{sec:gauge} In all FLPs found we have been able to use gauge transformations to remove all dependent variables in the associated nonlinear system from the Lax equations. \begin{definition}\label{defFLP} A Lax pair is called {\bf g-fake} if, on solutions of the equation appearing in its compatibility condition, one can remove all dependent variables in the associated nonlinear system from the Lax pair by applying gauge transformations. \end{definition} Definition \eqref{defFLP} applies to Lax pairs of any form. To illustrate how it works we consider an $N\times N$ discrete Lax pair of the form \begin{equation}\label{LP1} \ol{\phi}=L\phi, \qquad \widehat{\phi} = M\phi, \end{equation} where $\ol{\phi}$ and $\widehat{\phi}$ denote iterations of $\phi$ in two discrete independent variables. Suppose the entries of $L$ and $M$ depend on the dependent variable $u$, say, but are otherwise autonomous, and the compatibility condition $\widehat{L}M=\ol{M}L$ of this Lax pair is the (nonlinear) equation $Q(u,\ol{u},\widehat{u},...)=0$. If we make the invertible gauge transformation \begin{equation} \phi= G\psi \end{equation} where $G$ is an $N\times N$ matrix, then this Lax Pair becomes \begin{equation}\label{LP2} \ol{G}\,\ol{\psi}=LG\psi, \qquad \widehat{G}\widehat{\psi}=MG\psi. \end{equation} By Definition \ref{defFLP} this Lax pair is fake if we can choose a particular $G$ such that by using $Q(u,\ol{u},\widehat{u},...)=0$ in \eqref{LP2}, the equations \begin{equation}\label{LP3} \ol{G}^{-1}LG=C, \hspace{1cm} \widehat{G}^{-1}MG=D \end{equation} holds for some constant matrices $C$ and $D$, where each element of $G$ is a closed-form expression involving $u$ and its various shifts, that is \begin{equation}\label{LPsol} G_{ij}=G_{ij}(u,\ol{u},\widehat{u},...). \end{equation} The right hand sides of \eqn{LP3} are constant in this case because the elements of $L$ and $M$ are autonomous (except for their dependence on $u$). In general $C$ and $D$ may depend on the independent variables of the system, but must be independent of $u$ for the Lax pair to be fake. The Lax pair is therefore transformed to $\ol{\psi}=C\psi$, $\widehat{\psi}=D\psi$. When $C=I$ the first equation in \eqref{LP3} has the formal solution \begin{equation}\label{LP4} \ol{G}=L\ul{L}\,\ul{\ul L}...L_o \end{equation} for some constant matrix $L_o$. Fake Lax pairs of type \eqref{LP1} have the special property that whenever we use $Q(u,\ol{u},\widehat{u},...)=0$, the products $L\ul{L}$, $L\ul{L}\,\ul{\ul L}$ all simplify as algebraic expressions of $u$, $\ol{u}$, etc. such that \eqref{LP4} gives a closed-form expression for $G$. In practice, to find the simplest gauge transformation (usually when $C\neq I$) for fake Lax pairs of type \eqref{LP1}, one can use these products of $L$ to determine the location of all $u$'s within $G$, and choose the remaining constants in the simplest way such that $\det(G)\neq0$. An explicit example of this is shown in Example \eqref{example1}. \newline\par For Lax pairs of a different form to \eqref{LP1}, one can use similar techniques to find the required gauge transformation, as shown in the examples below. \begin{remark} {\bf The Inverse Scattering Transform (IST) cannot be used on Fake Lax pairs}. The IST is a method of finding solutions to integrable nonlinear partial differential and difference equations, using their associated Lax pairs. In the 1+1 continuous case, given an initial condition $u(x,0)$, one first solves $\phi_x=L\phi$ (say), obtaining $\phi(x,0)$ and the scattering data as Neumann series in $u(x,0)$. The second Lax equation $\phi_t=M\phi$ then gives the simple time evolution of the scattering data, from which $\phi(x,t)$ (and thus $u(x,t)$) is obtained from a singular integral equation or Gelfand-Levitan-Marchenko (GLM) integral equation. For g-fake Lax pairs however one may gauge out ($\phi=G(u,u_x,...)\psi$) all dependent variables from the Lax pair and obtain $\psi(x,t)$ explicitly, for any initial condition. The time-dependent eigenfunction is then given by $\phi(x,t)=G(u(x,t),u_x(x,t),...)\psi(x,t)$, where the arguments of $G$ are still unknown, and are the very objects that we are trying to determine. Any attempt to use a singular integral or GLM equation for $\phi$ yields only identities such as $u(x,t)=u(x,t)$. In other words the eigenfunctions do not contain enough nontrivial information about the solution to be used for the IST. \end{remark} \begin{remark}\label{rem:omega} In the $2\times 2$ case, for autonomous Lax pairs of type \eqref{LP1}, one can perform a simple test to see if a diagonal gauge transformation \begin{equation} \label{simpleG} G=\left(\begin{array}{cc} g_1 & 0 \\ 0 & g_2 \end{array}\right) \end{equation} exists that renders the Lax pair fake. Let $\left(\begin{array}{cc} A & B \\ C & D \end{array}\right)$ denote either $L$ or $M$, then from $\ol{G}^{-1}LG=const.$ and $\widehat{G}^{-1}MG=const.$ one finds that if $A,B,C,D$ are all nonzero, such a gauge transformation exists only if \begin{equation}\label{dettest} \frac{AD}{BC}=const. \end{equation} If \eqref{dettest} holds for both $L$ and $M$, then the gauge transformation $G$ exists and the Lax pair is a fake. If \eqref{dettest} does not hold for both $L$ and $M$, then we must do more work to determine if a more complicated $G$ can be found. For the $2\times2$ continuous case, if the autonomous Lax pair $\phi_x=L\phi$, $\phi_t=M\phi$ satisfies \begin{equation} BC=const., \quad \frac{\partial B}{B}=A-D+const. \end{equation} for both $L$ and $M$, where $\partial=\partial_x$ for $L$ and $\partial_t$ for $M$, then a gauge transformation of the form \eqref{simpleG} may be found to remove all dependent variables, rendering the Lax pair a fake. \end{remark} \subsection{Method to directly identify excess freedom}\label{sec:freedom} Given a Lax pair, the first step in applying the excess freedom test is to construct the generalised Lax pair of the same form. We can write the generalised Lax pair by maintaining the same dependence on the spectral parameter, and observing the coefficients of the linearly independent terms in the spectral parameter. Any of these coefficients that contain the dependent variable from the associated nonlinear system are replaced with arbitrary functions of the independent variables. The second step in the test is to find the form that these arbitrary functions must take, as determined by the equations that arise from the compatibility condition. \begin{definition}\label{defFLP2} A Lax pair is called {\bf u-fake} if the set of equations yielded by the compatibility condition, operating on the generalised Lax pair of the same form, is underdetermined. \end{definition} If the system of equations resulting from the compatibility condition is underdetermined, we can use the excess freedom to associated any equation, integrable or not, with the Lax pair. Such a Lax pair must be fake since the alternative contradicts the property that the associated system is integrable, this notion is discussed in \cite{c01}. The excess freedom test is best illustrated with examples, see section \ref{sec:examples}. \section{Examples}\label{sec:examples} Many examples of fake Lax pairs arise in the literature, some were published as real but were later found to be fake, others were deliberately written as examples of fake Lax pairs. One expects that there are still others that are erroneously thought to be real and have gone unnoticed. In all cases of fake Lax pairs known to the authors, both of the tests presented in this article correctly identify them. Furthermore, it is impossible for either method, correctly applied, to wrongly identify a real Lax pair as fake. \subsection{Example 1}\label{example1} The following example is a fully discrete fake Lax pair that was originally found in \cite{h11}. \subsubsection{Removal of dependent variable} Here we give details of the gauge analysis (section \ref{sec:gauge}) that is used to remove the dependent variable. The Lax pair is \begin{equation}\label{ex1eq1} \ol{\phi}=L\phi=\left(\begin{array}{cc} f & \ol{u}/\nu \\ 1/(\nu u) & 0 \end{array}\right)\phi, \qquad \widehat{\phi}=M\phi=\left(\begin{array}{cc} \hat{u}/u & \nu \hat{u} \\ \nu/u & 1-\nu^2 \end{array}\right)\phi, \end{equation} where $\nu$ is the spectral parameter and $u$ is the dependent variable of the nonlinear system. The compatibility condition of \eqref{ex1eq1} is \begin{equation}\label{ex1eq2} f=\ol{u}/u, \end{equation} where $f=f(u,\ol{u},...)$ can be {\it any} function of $u$ and its shifts. Since \eqref{ex1eq2} may be any equation whatsoever, it clearly should not have a real Lax pair. To show that \eqref{ex1eq1} is indeed fake, we use \eqref{ex1eq2} in \eqref{ex1eq1} to calculate \[ L\ul{L}=\left(\begin{array}{cc} \ol{u}/\ul{u}(1+1/\nu^2) &\ol{u}/\nu \\ 1/(\nu \ul{u}) & 1/\nu^2 \end{array}\right), \qquad L\ul{L}\,\ul{\ul L}=\left(\begin{array}{cc} \ol{u}/\ul{\ul u}(1+2/\nu^2) & \ol{u}/\nu(1+1/\nu^2) \\ 1/(\nu \ul{\ul u})(1+1/\nu^2) & 1/\nu^2 \end{array}\right), \] which, as hinted by \eqref{LP4} because we first look for a local gauge that only depends on $u$ and not its shifts, suggests that $G$ will be of the form \begin{equation}\label{ex1eq3} G=\left(\begin{array}{cc} ua & ub \\ c & d \end{array}\right), \end{equation} for some constants $a$, $b$, $c$ and $d$. This is indeed the case, and the only restriction on these constants is that $\det(G)\neq0$, so the simplest choice is \begin{equation}\label{ex1eq3} G=\left(\begin{array}{cc} u & 0 \\ 0 & 1 \end{array}\right). \end{equation} Notice that $G$ is diagonal, which by Remark \ref{rem:omega} we could have obtained from the fact that for $M$ we have $\frac{AD}{BC}=const.$ The gauge \eqref{ex1eq3} transforms \eqref{ex1eq1} to \begin{equation} \ol{\phi}=\left(\begin{array}{cc} 1 & 1/\nu \\ 1/\nu & 0 \end{array}\right)\phi \qquad \widehat{\phi}=\left(\begin{array}{cc} 1 & \nu \\ \nu & 1-\nu^2 \end{array}\right)\phi. \end{equation} \subsubsection{Excess freedom} To write the generalised form of \eqn{ex1eq1} we maintain the dependence on the spectral parameter and replace any quantities that depend on $u$ with arbitrary functions of the independent variables. Thus we obtain \begin{equation} L'= \left(\begin{array}{cc} a & b/\nu \\ c/\nu & 0 \end{array}\right), \qquad M'=\left(\begin{array}{cc} \alpha & \nu \beta \\ \nu\gamma & 1-\nu^2 \end{array}\right). \end{equation} The compatibility condition, $\widehat{L'}M'=\ol{M'}L'$ yields a set of equations for $a, b, c, \alpha, \beta$ and $\gamma$. If this set of equations is underdetermined, then the Lax pair is fake, which indeed it is in this case. The analysis was carried out in \cite{h11}, Appendix A, case 4, where $\delta$ must be set to unity and the dependence on the spectral variable is different, but equivalent because it yields the same set of equations from the compatibility condition. \subsection{Example 2} \label{sec:hone} In this section we consider a Lax pair that was published as being real in \cite{h07}. The Lax pair is actually fake, but it was published as being associated with a nonlinear QRT mapping. The Lax pair is \begin{equation}\label{h07LP} L=\left(\begin{array}{cc} u\bar{u} & -k_1 \\ \nu-u-\bar{u} & k_1(1/u+1/\bar{u})+k_2/(u\bar{u})\end{array}\right),\qquad M=\left(\begin{array}{cc} 0 & -k_1 \\ \nu-u-\bar{u} & k_1/\bar{u} \end{array}\right), \end{equation} where $\nu$ is the spectral parameter, $k_1$ and $k_2$ are constants and $u$ is the dependent variable in the associated nonlinear system. The linear problem is \begin{equation}\label{eg2LP} L\phi=\nu\phi,\qquad \bar{\phi}=M\phi \end{equation} so the compatibility condition is $\ol{L}M=ML$, which associates the Lax pair with the following QRT mapping \begin{equation} \bar{\bar{u}}=\frac{k_1 \bar{u}+k_2}{u\bar{u}^2}.\label{h07map} \end{equation} \subsubsection{Removal of dependent variable} An obvious property of \eqn{h07LP} is that \begin{equation}\label{eg2property} L=M+D, \end{equation} where $D$ is the diagonal matrix \begin{equation} D=\left(\begin{array}{cc} u\ol{u} & 0 \\ 0 & \ol{u}\,\ol{\ol u} \end{array}\right). \end{equation} From \eqref{eg2LP} this implies \begin{equation}\label{eg2neweqn} \bar{\phi}=(\nu I-D)\phi, \end{equation} so we can set \begin{equation}\label{eg2gauge} G=(\nu-u{\ol u})(\nu-{\ul u}u)(\nu-\ul{\ul u}\,{\ul u})\cdots(\nu-u_o)\left(\begin{array}{cc} 1 & 0 \\ 0 & \nu-u{\ol u} \end{array}\right) \end{equation} which transforms \eqref{eg2neweqn} to $\ol{\psi}=\psi$. The Lax pair \eqref{h07LP} is therefore fake, as will be any Lax pair of the form \eqref{eg2LP} with the property $L=M+D$. \subsubsection{Excess freedom} The first thing to do is construct the generalised Lax pair with the same form in terms of the spectral parameter, $\nu$, but with arbitrary terms replacing the coefficients of the various powers of $\nu$ in each entry, including $\nu^0$, that depend on $u$. We arrive at \begin{equation} L'=\left(\begin{array}{cc}a&-k_1\\\nu+c&d\end{array}\right),\qquad M'=\left(\begin{array}{cc}0&-k_1\\\nu+\gamma&\delta\end{array}\right), \end{equation} where $\nu$ is the spectral parameter, $k_1$ is a constant and all other terms are arbitrary functions of the lattice variable $n$. The required form of the arbitrary terms is found by substituting the Lax pair into the compatibility condition, which is $\ol{L'}M'=M'L'$ in this case. Below we write out all the equations coming from the compatibility condition, separated into different powers of the spectral parameter, which is independent of other variables. This yields \begin{align} \gamma&=c,& \bar{a}+\delta&=d,& \bar{d}\gamma&=a\gamma+c\delta,\label{andy1}\\ \bar{d}&=a+\delta,&\bar{d}\delta-k_1 \bar{c}&=d\delta-k_1\gamma.\label{andy2} \end{align} Now we solve this set of equations as follows. Allow (\ref{andy1}a) to define $\gamma$, which causes (\ref{andy1}c) to coincide with (\ref{andy2}a). Thus, the remaining compatibility conditions are: \begin{align} \bar{a}+\delta&=d,\label{andya}\\ \bar{d}&=a+\delta,\label{andyb}\\ b(\bar{c}-c)+\delta(\bar{d}-d)&=0.\label{andyc} \end{align} The difference between equations \eqn{andya} and \eqn{andyb} implies that \begin{equation} d=k_3-a,\label{andyd} \end{equation} where $k_3$ is a constant. Substituting this back into either \eqn{andya} or \eqn{andyb} gives us $\delta=k_3-\bar{a}-a$. Now the only remaining compatibility condition to satisfy is \eqn{andyc}, substituting the above results brings this equation to \begin{equation} k_1(\bar{c}-c)=\bar{a}^2-k_3\bar{a}-(a^2-k_3a), \label{cd1} \end{equation} which is satisfied by \begin{equation} c=\frac{a}{k_1}(a-k_3)+k_4, \label{cd} \end{equation} where, $k_4$ is a constant.. We have now satisfied all of the elements of the compatibility condition. Collecting the results, we find that the Lax pair takes the following form, or one that is gauge equivalent: \begin{equation} L'=\left(\begin{array}{cc}a&-k_1\\\nu+(a-k_3)a/k_1+k_4&k_3-a\end{array}\right),\qquad M'=\left(\begin{array}{cc}0&-k_1\\\nu+(a-k_3)a/k_1+k_4&k_3-\bar{a}-a\end{array}\right). \label{lp1} \end{equation} If this was a real Lax pair, substituting it into the compatibility condition would yield an equation for $a$, which would be the associated nonlinear map. However, one can check that, in this case, the compatibility condition is identically satisfied and no conditions remain. This leaves $a$ completely free. This freedom proves that the Lax pair is fake and we can use it to write any arbitrary equation into the Lax pair. As an example, we can retrieve an equivalent Lax pair to \eqn{h07LP} if, instead of allowing \eqn{andyd} to define $d$, we choose $d=k_1/u+k_1/\bar{u}+k_2/(u\bar{u})$, $a=u\bar{u}$ and set the $c$, $\gamma$ and $\delta$ according to the calculations given above with $k_4=k_2$. We can see that the resulting Lax pair is equivalent by noting that \eqn{h07map} can be summed to obtain the first order difference equation \[ u\bar{u}+\frac{k_1}{u}+\frac{k_1}{\bar{u}}+\frac{k_2}{u\bar{u}}=k_3, \] where $k_3$ is a constant. Using \eqn{h07map} to replace $k_1/u+k_2/(u\bar{u})$ in this expression shows that $u$ must also satisfy \[ u\bar{u}+\bar{u}\bar{\bar{u}}+\frac{k_1}{u\bar{u}}=k_3, \] It follows that the two Lax pairs, both fake, are equivalent. \subsection{Example 3} \label{sec:gramani} In this section we analyse a fake Lax pair that was published as a real one in \cite{gr04} (equations (3.23) in that paper). The Lax pair is \begin{equation}\label{eg3LP} \phi(q\nu)=L\phi(\nu), \qquad \bar{\phi}(\nu)=M\phi(\nu) \end{equation} where \begin{equation}\label{eg3LPmat} L=\left(\begin{array}{cccc} 0 & 0 & k/u & 0 \\ 0 & 0 & \ul{u} & q\ul{u} \\ \nu u & 0 & 1 & q \\ 0 & \nu \ul{k}/\ul{u} & 0 & 0 \end{array}\right) \qquad M=\left(\begin{array}{cccc} 0 & u/(k(u+1)) & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 1/u & q/u \\ \nu & 0 & 0 & 0 \end{array}\right) \end{equation} and $k$ is a function of $n$ such that $\bar{\bar{k}}=qk$. The compatibility $\ol{L}(\nu)M(\nu)=M(q\nu)L(\nu)$ yields $q$-P$_I$ \begin{equation}\label{qP1} \ul{u}\bar{u}=k\bar{k}\left(\frac{1+u}{u^2}\right). \end{equation} \subsubsection{Removal of dependent variable} We look for a diagonal gauge transformation. Define the function $f_n$ by \begin{equation} f_n=\prod_{i=i_o}^{n-1}1/u_i, \end{equation} and let the gauge transformation $G$ be \begin{equation}\label{eg3gauge} G=\left(\begin{array}{cccc} \bar{f}/q & 0 & 0 & 0 \\ 0 & \ul{f} & 0 & 0 \\ 0 & 0 & f & 0 \\ 0 & 0 & 0 & f/q \end{array}\right). \end{equation} Using \eqref{qP1} we then find that \begin{equation} G^{-1}LG=\left(\begin{array}{cccc} 0 & 0 & qk & 0 \\ 0 & 0 & 1 & 1 \\ \nu/q & 0 & 1 & 1 \\ 0 & \nu \bar{k} & 0 & 0 \end{array}\right), \qquad \ol{G}^{-1}MG=\left(\begin{array}{cccc} 0 & q\bar{k} & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 1 & 1 \\ \nu & 0 & 0 & 0 \end{array}\right), \end{equation} and so the Lax pair \eqref{eg3LPmat} is transformed to \begin{equation} \psi(q\nu)=\left(\begin{array}{cccc} 0 & 0 & qk & 0 \\ 0 & 0 & 1 & 1 \\ \nu/q & 0 & 1 & 1 \\ 0 & \nu \bar{k} & 0 & 0 \end{array}\right)\psi(\nu), \qquad \bar{\psi}(\nu)=\left(\begin{array}{cccc} 0 & q\bar{k} & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 1 & 1 \\ \nu & 0 & 0 & 0 \end{array}\right)\psi(\nu), \end{equation} showing that it is indeed fake. \subsubsection{Excess freedom} To show that the excess freedom exists in the compatibility condition of this Lax pair, we first write a generalised form of the Lax pair with the same dependence on the spectral parameter. Noting \eqn{eg3LPmat} we immediately write the generalised form \begin{equation} L'=\left(\begin{array}{cccc} 0&0&a&0\\0&0&b&c\\\nu u&0&1&q\\0&\nu d&0&0\end{array}\right),\qquad M'=\left(\begin{array}{cccc} 0&\alpha&0&0\\0&0&1&0\\0&0&\beta&\gamma\\ \nu&0&0&0\end{array}\right), \end{equation} where $\nu$ is the spectral parameter and all other variables inside the Lax matrices are arbitrary functions of the discrete lattice variable $n$. Notice that we have left the dependent variable $u$ in the (3,1) entry of $L'$, but at this stage we do not know what equation $u$ must satisfy, so essentially it is an arbitrary function of the independent variable, $n$, and this is the generalised Lax pair. Substituting these matrices into the compatibility condition $L'(q\nu)M'=\ol{M'}L'$ yields the following set of equations \begin{align} \bar{a}\beta&=b\alpha,&\bar{a}\gamma&=c\alpha,&\bar{b}\beta&=1,\label{gr1}\\ \bar{b}\gamma&=q,&u\beta&=q,&\bar{u}\alpha&=d\gamma,\label{gr2}\\ \gamma&=q\beta,&\bar{d}&=qa.\label{gr3} \end{align} Thus, the compatibility condition delivers a set of eight equations for the eight unknown functions, and all may appear to be well, but some equations coincide leaving excess freedom in this system. Naming the equations in \eqn{gr1} as (\ref{gr1}a), (\ref{gr1}b) and (\ref{gr1}c), respectively (and naming \eqn{gr2} and \eqn{gr3} the same way), we do the following. Allow (\ref{gr2}b), (\ref{gr1}c), (\ref{gr3}a) and (\ref{gr3}b) to define $\beta$, $b$, $\gamma$ and $d$, respectively. Notice that these definitions cause (\ref{gr2}a) to coincide with (\ref{gr1}c). Thus, the remaining equations become \begin{align} q^2\bar{a}/u&=\alpha\ul{u},&q^2\bar{a}/u&=c\alpha,&\bar{u}\alpha&=q^3\ul{a}/u.\label{gr4} \end{align} Comparing (\ref{gr4}a) and (\ref{gr4}b) shows that $c=\ul{u}$. Now use (\ref{gr4}c) to define $a$, which causes the final remaining equation, (\ref{gr4}a) to become \begin{equation} \bar{\bar{\bar{u}}}\bar{\bar{u}}\bar{u}\bar{\bar{\alpha}}=q\bar{u}u\ul{u}\alpha.\label{grlinear} \end{equation} \begin{remark} At this point in the analysis, one should make a change of variables such as $\alpha=x/(\bar{u}u\ul{u})$ so that \eqn{grlinear} becomes $\bar{\bar{x}}=qx$. The Lax pair is real when associated with this trivial equation for $x$. However, we can construct fake Lax pairs by continuing without this change of variables. \end{remark} Naturally, \eqn{grlinear} is solved by \[\alpha=\frac{k}{\bar{u}u\ul{u}},\] where $\bar{\bar{k}}=qk_n$ and there is one excess degree of freedom. This freedom can be used to write a fake Lax pair for any equation, writing $\alpha=u/k(1+u)$ returns the fake Lax pair for $q$-P$_I$ from \cite{gr04}. For the reader's convenience, the full fake Lax pair is written below: \begin{equation} L'=\left(\begin{array}{cccc} 0&0&\bar{\bar{u}}\bar{u}\bar{\bar{\alpha}}/q^3&0\\0&0&\ul{u}/q&\ul{u}\\\nu u&0&1&q\\0&\nu \bar{u}u\alpha/q^2&0&0\end{array}\right),\qquad M'=\left(\begin{array}{cccc} 0&\alpha&0&0\\0&0&1&0\\0&0&q/u&q^2/u\\ \nu&0&0&0\end{array}\right), \end{equation} where $\alpha$ is arbitrary. \subsection{Example 4} As a final example we consider the fake Lax pair \begin{equation}\label{eg4LP} \phi_x=\left(\begin{array}{cc} 0 & 1 \\ \lambda f^2 & \mu f +f_x/f \end{array}\right)\phi, \qquad \phi_t=\left(\begin{array}{cc} \nu & (g+\rho)/f \\ \lambda f(g+\rho) & \nu+\mu(g+\rho)+g_x/f \end{array}\right)\phi \end{equation} which was given in \cite{cn91} and identified as fake in that paper. Note that the Lax pair was given in scalar form in \cite{cn91}, equation (6), but transforming it into matrix form is straightforward, as mentioned in \cite{s01}. The compatibility condition of \eqref{eg4LP} is \begin{equation}\label{eg4compat} f_t=g_x \end{equation} for any functions $f$ and $g$, and $\lambda, \mu, \nu, \rho$ are all independent parameters. This fake Lax pair can be used to represent a large class of nonlinear systems by arbitrarily writing $f$ and $g$ in terms of $u$, which becomes the dependent variable in the nonlinear system. \subsubsection{Removal of dependent variable} Consider first the diagonal gauge transformation \begin{equation} G=\left(\begin{array}{cc} 1 & 0 \\ 0 & f \end{array}\right), \end{equation} which, using \eqref{eg4compat}, transforms \eqref{eg4LP} to \begin{equation}\label{eg4LPa} \psi_x=f\left(\begin{array}{cc} 0 & 1 \\ \lambda & \mu \end{array}\right)\psi, \qquad \psi_t=\nu\psi+(g+\rho)\left(\begin{array}{cc} 0 & 1 \\ \lambda & \mu \end{array}\right)\psi. \end{equation} If we let $A=\left(\begin{array}{cc} 0 & 1 \\ \lambda & \mu \end{array}\right)$ then we can make the second gauge transformation \begin{equation} \psi=\exp\left(A\int(g+\rho)dt+\nu t\right)\psi^{o}, \end{equation} which using \eqref{eg4compat} transforms \eqref{eg4LPa} to $\psi^o_x=\psi^o_t=0$. Thus the combined gauge transformation $\phi=G\exp\left(A\int(g+\rho)dt+\nu t\right)\psi^{o}$ removes all dependent variables from the Lax pair \eqref{eg4LP}. \subsubsection{Excess freedom} In fact, this Lax pair fails the excess freedom test by construction, since the compatibility condition is associated with an underdetermined system in \eqn{eg4compat}. We go ahead with the analysis anyhow, to see how to analyse any case where, for example, $g$ is given in terms of $f$ (or $f$ and $g$ given in terms of some $u$), so that the Lax pair's fakeness is not obvious. A generalised form of the Lax pair in \eqn{eg4LP} is given by \[ L'=\left(\begin{array}{cc}0&1 \\ \lambda c & \mu d + b\end{array}\right),\qquad M'=\left(\begin{array}{cc}\nu&\beta \\ \lambda \gamma & \nu + \mu \delta + \alpha\end{array}\right). \] Here the extraneous parameter $\rho$ has been excluded because it is inconsequential and the introduced variables $b$, $c$, $d$, $\alpha$, $\beta$, $\gamma$ and $\delta$ all depend on both the continuous independent variables, $x$ and $t$. The remaining quantities, $\lambda$, $\mu$ and $\nu$ can all be considered spectral parameters. The compatibility condition is $L'_t+L'M'=M'_x+M'L'$ from which we obtain the following set of equations \begin{align} \gamma&=c\beta,&\delta&=d\beta,&d\gamma&=c\delta,\label{c1}\\ d_t&=\delta_x,&b_t&=\alpha_x,&\alpha&=\beta_x+b\beta,\label{c2}\\ c_t+b\gamma&=\gamma_x+c\alpha.\label{c3} \end{align} We solve this set of equations as follows: let (\ref{c1}a) and (\ref{c1}b) define $\gamma$ and $\beta$ respectively, this causes (\ref{c1}c) to become an identity. Let (\ref{c2}c) define $\alpha$ and substitute all these definitions into (\ref{c2}b) and (\ref{c3}), which become \begin{align} b_t&=\left[\left(\frac{\delta}{d}\right)_x+\frac{b\delta}{d}\right]_x,\label{eg4:b}\\ c_t&=c_x\frac{\delta}{d}+2c\left(\frac{\delta}{d}\right)_x,\label{eg4:c} \end{align} respectively. Equation \eqn{eg4:b} is solved by introducing $u=u(x,t)$ such that \[b=u_x,\qquad \left(\frac{\delta}{d}\right)_x+\frac{b\delta}{d}=u_t.\] Using (\ref{c2}a) shows that the latter of these is satisfied when $u=\log{d}$. Once again using (\ref{c2}a), we find that \eqn{eg4:c} is satisfied when $\log{c}=2\log{d}$. Now the system of equations obtained from the compatibility condition is satisfied. There are various ways to write the resulting Lax pair, allowing $d=f$, $\delta=g$ and expressing the other variables in terms of these retrieves \eqn{eg4LP} (with $\rho=0$). The system of equations is undetermined, which confirms that the Lax pair is fake. We could take any of the compatibility conditions to be the associated nonlinear system, in \cite{cn91} they choose (\ref{c2}a) to play that role. \section{Conclusion}\label{sec:conclusion} In this letter, we have provided two very simple methods to identify fake Lax pairs. The methods have been used to confirm that several Lax pairs taken from the literature are fake, some of which were intentionally given as fake, while others were thought to be real. Both methods are widely applicable, with the only restriction being that the Lax par should contain a non-removable spectral parameter. We leave as open questions any discussion about the usability and reality of Lax pairs that do not contain a spectral parameter and do not allow for one to be inserted. However, it is the authors' opinion that such Lax pairs are also fake, since they do not allow for any spectral analysis. The extent to which a fake Lax pair can be used to gain information about its associated system is also open to debate. For example in \cite{k80} a Lax pair for Burger's equation is given. The Lax pair is fake according to the tests given here and it cannot be used for inverse scattering. However, the Lax pair is related to a Cole-Hopf transformation which linearises Burger's equation and is thus useful. We do not discuss it further here, but some sections of \cite{c01} are devoted to this issue.
\section{Introduction} \label{intro} In Refs.~\cite{Biernat:2009my,Biernat:2010tp,Biernat:2011mp,GomezRocha:2012zd,Gomez-Rocha:2013bga} the point form of relativistic quantum mechanics has been advocated as an appropriate framework for calculating the electroweak structure of bound few-body systems, in particular of mesons and baryons within the scope of constituent-quark models. The route to derive electroweak form factors, pursued in these papers and followed also here, is to describe the physical process in which particular form factors are measured in a Poincar\'e invariant way, calculate the invariant 1$\gamma$- or 1$W$-exchange amplitude, extract the hadronic current from this amplitude, analyze its covariant structure and identify the form factors. Poincar\'e invariance is thereby guaranteed by employing the Bakamjian-Thomas construction~\cite{Bakamjian:1953kh}. The essence of the Bakamjian-Thomas construction in point form is that the 4-momentum operator factorizes into an interaction-dependent mass operator and a free 4-velocity operator \begin{equation}\label{eq:massop} \hat{P}^{\mu}=\hat{\mathcal M}\, \hat V^{\mu}_{\mathrm{free}}= \left(\hat{\mathcal M}_{\mathrm{free}}+ \hat{\mathcal M}_{\mathrm{int}} \right) \hat V^{\mu}_{\mathrm{free}}\, . \end{equation} The dynamics of the system is thus completely encoded in the mass operator. From Eq.~(\ref{eq:massop}) it is quite obvious that all four components of the 4-momentum operator are interaction dependent. The generators of Lorentz transformations, on the other hand, are interaction free. These are the properties which characterize the point form of relativistic dynamics. They make it comparably simple to boost and rotate wave functions and add angular momenta. Since $\gamma$- and $W$-exchange are treated dynamically, one has to allow for particle emission and absorption. This is accomplished by using a coupled-channel framework with a matrix mass operator $\hat{\mathcal M}$ that acts on the direct sum of the pertinent multiparticle Hilbert spaces. The diagonal entries $\hat{M}_i$ of $\mathcal{M}$ are the sum of the relativistic kinetic energies of the particles in channel $i$. In addition, the $\hat M_i$ may contain instantaneous interactions between the particles, like the confinement potential between quark and antiquark. Off-diagonal entries of $\mathcal{M}$ are vertex operators $\hat{K}_{i\rightarrow j}$ and $\hat{K}_{j\rightarrow i}=\hat{K}_{i\rightarrow j}^\dag$ which describe the absorption and emission of particles and hence the transition from one channel to the other. A most convenient basis to represent all these operators is formed by a complete set of velocity states~\cite{Klink:1998zz}. A velocity state is a multiparticle momentum state in the rest frame that is boosted to an overall 4-velocity $V$ ($V_\mu V^\mu=1$) by means of a rotationless boost $B_c(V)$: \begin{equation}\label{eq:vstates} \vert V; {\bf k}_1, \mu_1; {\bf k}_2, \mu_2; \dots ; {\bf k}_n, \mu_n\rangle := \hat{U}_{B_c(V)} \vert {\bf k}_1, \mu_1; {\bf k}_2, \mu_2; \ldots ; {\bf k}_n, \mu_n\rangle\quad \mathrm{with} \quad \sum_{i=1}^{n}{{\bf k}_i}=0\, . \end{equation} The $\mu_i$s are the spin projections of the individual particles. Using velocity states, matrix elements of vertex operators can be simply related to field theoretical interaction Lagrangean densities \cite{Klink:2000pp} \begin{equation} \label{vertexop} \langle V^{\prime}; { {\bf k}^{\prime}_{j}, \mu_{j}^{\prime} } \vert \hat{K} \vert V; {\bf k}_i, \mu_i \rangle \propto V^0\, \delta^3({\bf V} - {\bf V}^{\prime})\, \langle {\bf k}^{\prime}_j, \mu_j^{\prime} \vert \hat{\mathcal{L}}_{\mathrm{int}}(0)\vert { {\bf k}_i, \mu_i } \rangle \, . \end{equation} At this point it is worthwhile to remark that conservation of the overall 3-velocity at interaction vertices is a specific feature of the Bakamjian-Thomas construction and does not hold, in general, for point-form quantum-field theories. It is this overall velocity-conserving delta function that leads to wrong cluster properties, an unwanted feature of the Bakamjian-Thomas construction which is observed in any form of relativistic dynamics and is not just specific to the point form~\cite{Keister:1991sb}. The physical consequences of wrong cluster properties in our case are that the gauge-boson-hadron vertices, which we analyze to obtain the transition form factors, may not only depend on the momenta attached to the vertex, but also on the lepton momenta. Formally such wrong cluster properties could be cured by means of, so called, \lq\lq packing operators\rq\rq, but practically these are hard to construct. The strategy to obtain sensible results for the form factors adopted in Refs.~\cite{Biernat:2009my,Biernat:2010tp,Biernat:2011mp,GomezRocha:2012zd,Gomez-Rocha:2013bga} and also followed here is thus rather to look for kinematical situations in which those spurious dependencies are minimized or vanish. \vspace{-0.3cm} \section{Weak $B\rightarrow D$ transition form factors} \label{sec:ffs} \subsection{Space-like momentum transfers} \label{sec:ffspace} \begin{figure}[t!] \includegraphics[width=0.42\textwidth]{F1BDspace.eps}\hfill \includegraphics[width=0.42\textwidth]{F0BDspace.eps} \caption{The weak form factors $F1$ and $F0$ for the $B\rightarrow D$ transition in the space-like momentum-transfer region as functions of $Q^2=-(p_B-p_D')^2$. The solid and dashed lines refer to calculations in the infinite-momentum frame and the Breit frame, respectively. The shaded area indicates the frame dependence caused by the violation of cluster separability.} \label{bdspacelike} \end{figure} Our first goal is to derive weak $B\rightarrow D$ transition form factors for space-like momentum transfers as they can, in principle, be measured in $\nu_e\, B^-\rightarrow e^- D^0$ scattering. In order to account for dynamical exchange of $W$-bosons we set up a 4-channel problem that includes all states which occur during such a scattering process if considered within a valence-quark picture (i.e. $|\nu_e, b, \bar u \rangle$, $|e, W^+, b, \bar u \rangle$, $|e, c, \bar u \rangle$, $|\nu_e, W^-, c, \bar u \rangle$). An instantaneos confinement potential between quark and antiquark is included in the diagonal entries of the matrix mass operator. Using perturbation theory for the weak coupling we calculate the invariant 1$W$-exchange amplitude for $\nu_e\, B^-\rightarrow e^- D^0$ scattering. It is the sum of two time-ordered contributions which, as expected, is proportional to the contraction of a lepton with a meson current times the covariant $W$-boson propagator. This allows us to identify the weak meson current in a unique way. It is an overlap integral of $B$- and $D$-meson wave functions multiplied by the weak quark current, a Wigner-rotation factor and a kinematical factor (see, e.g., Refs.~\cite{GomezRocha:2012zd,Gomez-Rocha:2013bga}). After replacing the CM momenta $k_B$ and $k_D'$ by the respective physical particle momenta $p_B=B_c(V) k_B$ and $p_D'=B_c(V) k_D'$ we end up with a meson current that transforms like a 4-vector and has the covariant decomposition~\cite{GomezRocha:2012zd,Gomez-Rocha:2013bga} (with $q=p_B-p_D'$): \begin{equation} J^{\mu}({\bf p}_D',{\bf p}_B)=\left[(p_B+p_D')-\frac{m_B^2-m_D^2}{q^2}\, q\right]^\mu\, F_1(Q^2,s)+ \frac{m_B^2-m_D^2}{q^2}\, q^{\mu}\, F_0(Q^2,s)\, . \end{equation} The physical consequences of wrong cluster properties, inherent in the Bakamjian-Thomas construction, become obvious in this decomposition. The form factors cannot be chosen such that they are only functions of the squared 4-momentum transfer at the $W$-meson vertex, they also depend on Mandelstam $s$, i.e. the invariant mass squared of the whole neutrino-$B$-meson (or equivalently electron-$D$-meson) system. This does not spoil Poincar\'e invariance of the scattering amplitude, it just means that the $WBD$-vertex is also influenced by the presence of the scattering lepton. The Mandelstam-$s$ dependence may also be interpreted as a frame dependence of the $W^\ast B\rightarrow D$ subprocess. We consider two extreme cases, namely the minimum value of $s$ necessary to reach a particular $q^2<0$ and $s\rightarrow \infty$. The first choice corresponds to the Breit frame (BF), the second to the infinite-momentum frame (IF). Using simple harmonic-oscillator wave functions with the oscillator parameters and masses taken from a front-form calculation of weak $B\rightarrow D$ transition form factors in the time-like region~\cite{Cheng:1996if} we obtain the results that are plotted in Fig.~\ref{bdspacelike}. Whereas the differences between IF and BF are small for $F_1$, they can be sizeable for $F_0$. From a physical point of view the IF seems to be preferable: unlike the BF, so called \lq\lq Z-graphs\rq\rq\ are suppressed in the IF and it is thus not necessary to take them into account, the Mandelstam-$s$ dependence vanishes for $s\gg m_B^2$, and our IF results agree with front-form calculations in the $q^+=0$ frame. \vspace{-0.2cm} \subsection{Time-like momentum transfers} \label{sec:fftime} \begin{figure}[t!] \centering \includegraphics[width=0.42\textwidth]{F1BDtime.eps}\hfill \includegraphics[width=0.42\textwidth]{F0BDtime.eps} \caption{The weak form factors $F1$ and $F0$ for the $B\rightarrow D$ transition for space- and time-like momentum transfers. The solid line refers to the analytic continuation from space- to time-like momentum transfers by making the replacement $Q\rightarrow iQ$ in the infinite-momentum-frame result. The dashed line is the outcome of a decay calculation in the $B$ rest frame~\cite{GomezRocha:2012zd,Gomez-Rocha:2013bga}.} \label{bdtime} \end{figure} In the time-like momentum transfer region these form factors can be measured in semileptonic weak decay processes. Theoretically it is straightforward to adapt our relativistic multichannel approach such that one can deal with decay processes like $B\rightarrow D e \bar\nu_e$. Working in the velocity-state representation the decaying $B$-meson has to be at rest. For this kinematical situation it is, however, known from front-form calculations~\cite{Choi:1999bg} that non-valence contributions leading to Z-graphs may become important. As already mentioned, a similar observation can be made for space-like momentum transfers, if the form factors are calculated in the BF and not in the IF (see also Ref.~\cite{Simula:2002vm}). In order to avoid problems with Z-graphs it is thus suggestive to take the form factor expressions obtained in the IF for space-like momentum transfers and continue them analytically to time-like momentum transfers by the replacement $Q\rightarrow iQ$. This is done in Fig.~\ref{bdtime} (solid lines). In this figure the results of a direct decay calculation~\cite{GomezRocha:2012zd,Gomez-Rocha:2013bga} (Z-graphs absent) are shown for comparison (dashed lines). The differences may be considered as an estimate of the size of Z-graph contributions. From the considerations just made it is clear that the solid line should be closer to experiment than the dashed one. This is indeed the case. What one knows experimentally is the slope of $F_1$ at zero recoil as measured in $B\rightarrow D e \bar\nu_e$ decays~\cite{Asner:2010qj}. It agrees approximately with the value which we get from our analytic continuation, whereas the decay calculation provides a much smaller value. \vspace{-0.2cm} \section{Outlook} \label{sec:outlook} Having estimated the size of Z-graph contributions indirectly, our next task is now the explicit inclusion of Z-graphs in the decay calculation. One possible time ordering of a Z-graph contribution to the $B\rightarrow D \, e\, \bar\nu_e$ decay is shown in Fig.~\ref{zgraph}. It obviously requires the creation of a $c\bar{c}$-pair, which we plan to describe by means of a $^{3}P_0$ model~\cite{Segovia:2012cd}. Within our relativistic coupled-channel approach such a pair creation is easily accommodated by an additional channel and a vertex interaction obtained from the field-theoretical $^{3}P_0$ interaction Lagrangean~\cite{Klink:2000pp}. Assuming instantaneous confinement between the respective quark-antiquark pairs (indicated by blobs in Fig.~\ref{zgraph}) on ends up with a simple, vector-meson-dominance-like physical picture. The $Z$-graph contribution to $B\rightarrow D$ decay can be understood as a $B\rightarrow D$ transition induced by the emission of a $B_c^\ast$ and, less important, excited ($b\bar{c}$) vector mesons, which subsequently decay into $e\bar{\nu}_e$ by means of a $W$. This simplifies the calculation of Z-graphs considerably, since the main part can be done on the hadronic level. Only the $B_c^\ast B D$-vertex (i.e. coupling strength and form factors) and the $B_c^\ast$ decay has to be determined on the quark level. Proceeding in this way we hope to achieve a more quantitative estimate of Z-graph contributions in weak meson decay form factors. \begin{figure}[t!] \centering \includegraphics[width=0.45\textwidth]{zgraph.eps} \caption{Z-graph contribution to the 1$W$-exchange amplitude of the semileptonic $B\rightarrow D e \bar\nu_e$ decay} \label{zgraph} \end{figure} \vspace{-0.2cm} \begin{acknowledgements} M. G\'omez-Rocha acknowledges the support of the \lq\lq Fonds zur F\"orderung der wissenschaftlichen Forschung in \"Osterreich (FWF DK W1203-N16 and in part under project P25121-N27).\end{acknowledgements} \vspace{-0.6cm}
\section{\label{sec:1}Introduction} Today everyone knows that Quantum Chromodynamics (QCD) is a promising theory of strong interactions. In the low energy regime of this theory, there are still open unsolved questions like quark confinement, which has been a challenging problem over thirty years. A lot of phenomenological models have been introduced to determine the behavior of QCD in this regime. In these models, it is believed that the QCD vacuum is responsible for confinement. Among wide variety of phenomenological models, center vortex theory based on magnetic degrees of freedom gives an acceptable explanation of quark confinement \cite{thooft, *vin, *cornwall, *feynman, *nielsen, *ambjorn1, *ambjorn2, *ambjorn3, *olesen}. Numerical lattice calculations show the evidence for the existence of the topological objects called center vortices in the vacuum \cite{deldebbio,*deldebbio2,*tomboulis,*tomboulis2,*reinhardt}. Center vortices are quantized magnetic flux tubes (or surfaces) which are closed according to the Bianchi identity \cite{dirac}. They are created by a singular gauge transformation and have finite energy per unit length (or surface). Based on the center vortex theory, the area law fall-off for the Wilson loop which leads to quark confinement is due to the quantum fluctuations of the vortices interacting with the Wilson loop. This interaction affects the Wilson loop in the representation $r$ by an element of the $SU(N)$ gauge group center: \begin{equation} W_r(C)\longrightarrow z^{k_r} W_r(C), \label{W} \end{equation} where $k_r$ is the N-ality of the representation $r$. On the other hand, one might say that when the vortex does not link with the Wilson loop, the loop remains unaffected. Although the confinement at large distances is produced by these types of vortices, they cannot give the linear potential at intermediate distance. To get the intermediate potentials, Faber {\it et al.} generalized this model to thick center vortex model to eliminate this defect \cite{faber}. One might claim that the thick center vortex model explains the following features of confinement very well:\\ \begin{enumerate} \item The coulombic potential at short distances that has been proved in two different methods: Deldar {\it et al.} showed by increasing the role of vortex fluxes piercing Wilson loops with contributions close to the trivial center element and by fluctuating the vortex core size, the coulombic potential is produced \cite{deldar2009}. In a different work, the fluctuation of non-quantized, closed magnetic flux lines has been considered by Faber {\it et al.} to calculate the short range potential \cite{denis}. \item Casimir Scaling of string tensions extended from the onset of confinement to the onset of screening which has been proved precisely by lattice calculations \cite{deldarlat,bali}. This theory claims that string tensions are roughly proportional to the eigenvalues of the quadratic Casimir operators. The results obtained from the thick center vortex model show that the string tension of different representations is qualitatively in agreement with Casimir scaling \cite{faber,greensite, deldar-jhep,deldar2005,deldar2007,deldar2009,deldar2010}, even for exceptional G(2) gauge group \cite{deldar-g2}. \item N-ality dependence of asymptotic string tensions at large distances which frames that the asymptotic string tension of each representation is the same as that of the lowest dimensional representation with the same N-ality. Calculations in thick center vortex model confirm this feature of confinement \cite{faber,greensite, deldar-jhep,deldar2005,deldar2007,deldar2009,deldar2010}. \end{enumerate} When the vortex is thickened, the vortex core may overlap the perimeter of the Wilson loop and a part of the vortex flux enters the loop. Therefore, the role of the center element is replaced by the group factor which interpolates between a center element -when the vortex core is completely inside the Wilson loop- and $1$ -when the Wilson loop and the vortex do not interact. Therefore, studying the behavior of the group factor at different quark separation distances will give useful information about the potential between quark and anti-quark. In the next section, the thick center vortex model is studied and it is explained how the group factor appears in the model. Moreover, the maximum flux of the $SU(4)$ vortices is calculated and it is discussed which of the Cartan generators should be used in the calculation. The behavior of the group factor at different quark separation is discussed in section three and the results are given in section four. \section{\label{sec:2}The group factor in the thick center vortex model} In the original vortex model, center elements of the gauge group are responsible for confinement. For thin vortices two scenarios may occur:\\ 1) Vortices and the Wilson loop do not link and therefore the Wilson loop is unaffected: \begin{equation} W(C)=Tr \big[U...U\big]\longrightarrow Tr \big[U...I...U\big]. \label{W1} \end{equation} 2) Vortices may pierce the minimal area of the Wilson loop. The effect of this interaction is inserting a center element $z$ between link operators: \begin{equation} W(C)=Tr \big[U...U\big]\longrightarrow Tr \big[U...z...U\big]. \label{W2} \end{equation} Therefore, if {\it $f$} is the probability of piercing a plaquete by the $z$ vortex of the $SU(2)$ gauge group, the Wilson loop might be written as: \begin{eqnarray} <W(C)>=\prod_{x\in A}{[(1-f)+f(-I)]}<W_0(C)> \nonumber\\ =\exp \big[-\sigma(C)A\big]<W_0(C)>, \label{<W>} \end{eqnarray} where $<W_0(C)>$ is the Wilson loop expectation value when no vortices pierce the minimal area of the loop and the string tension is \begin{equation} \sigma=-\ln(1-2f). \label{sigma} \end{equation} Then the thin vortices give the correct N-ality dependence of the potentials at large distances while the intermediate linear potential is lost. Thus, to get the intermediate potentials, the vortices are thickened. Thickening the vortices leads to a new scenario. The vortex may overlap the perimeter of the Wilson loop and one has to consider a distribution for the flux carried by the vortex. In this case, a part of the vortex flux might enter the Wilson loop and the center element is replaced by a group element $G$ which is a unitary matrix called the $SU(N)$ group factor. This factor parametrizes the influence of the vortex on the Wilson loop: \begin{equation} W(C)=Tr \big[UU...U\big]\longrightarrow Tr\big[UU...G...U\big], \label{W3} \end{equation} where \begin{equation} G(x,S)=S \exp \big[i \vec{\alpha}^n_c.\vec{H}\big]S^\dagger, \label{G} \end{equation} where the $H_i$'s, $\big\lbrace i=1,...,N-1 \big \rbrace $ are the generators spanning the Cartan sub-algebra and $S$ is an $SU(N)$ group element in representation $r$. $\alpha^{(n)}_c$ represents the flux distribution of the $z_n$ vortex. In general, in an $SU(N)$ gauge group, there are $N-1$ types of center vortices corresponding to the number of the center elements of the gauge group. Based on the assumptions of the model, the random group orientations associated with $S_i$ are considered uncorrelated. Therefore, one has to average $G$ over all orientations in the group manifold: \begin{eqnarray} \bar{G}(\vec{\alpha})\equiv {\cal G}_r \big[\vec{\alpha}^{(n)} \big]I=\int dS S \exp\big[i\vec{\alpha}.\vec{H}\big]S^\dagger \nonumber\\ =\frac{1}{d_r}Tr \exp\big[i\vec{\alpha}^{(n)}.\vec{H}\big], \label{gr} \end{eqnarray} where $d_r$ is the dimension of representation $r$. Therefore, the potential energy between static sources induced by the vortices is \begin{equation} V(R) = -\sum_{x \in A}\ln\left\{ 1 - \sum^{N-1}_{n=1} f_{n} (1 - {\mathrm {Re}} {\cal G}_{r} [\vec{\alpha}^n_{C}(x)])\right\}. \label{V} \end{equation} In Eq.~(\ref{V}), it is observed that the factor ${\cal G}_r \big[\vec{\alpha} \big]$ owns an important role in producing the potentials. According to Eq.~(\ref{gr}), this factor depends on the flux $\alpha_c(x)$. Furthermore, the vortex profile $\alpha_c(x)$ depends on what fraction of the vortex flux enters the loop $C$. Therefore, it depends on both the shape of the loop $C$, and the position $x$ of the center of the vortex core, relative to the perimeter of the loop, as the following:\\ \begin{enumerate} \item When the vortex core is entirely outside the planar area enclosed by the Wilson loop, it cannot affect the loop: \begin{equation} \exp \big[i \vec{\alpha}^{(n)}.\vec{H}\big]=I\Longrightarrow {\vec{\alpha}}^{(n)}=0. \label{alpha_min} \end{equation} In this case, the lower limit for $\alpha$ is achieved.\\ \item When the vortex core is completely inside the Wilson loop area, then the influence of the vortex on the Wilson loop is given by a center element: \begin{equation} \exp\big[i \vec{\alpha}^{(n)}.\vec{H}\big]=z_n I \quad \Longrightarrow {\vec{\alpha}}^{(n)}={\vec{\alpha}}^{(n)}_{\mathrm max}, \label{alpha-max} \end{equation} where $I$ is the unit matrix. ${\alpha}^{(n)}_{max}$ is the upper limit of the vortex profile and should be calculated for every vortex type in the $SU(N)$ gauge group. \\ \item As $R\rightarrow 0$ then $\alpha \rightarrow 0$. \end{enumerate} It should be noticed that all choices for $\alpha (x)$ must lead to a well-defined potential which means both Casimir proportionality and N-ality dependence must be preserved. The vortex profiles checked in this paper are listed below: \begin{enumerate} \item The flux introduced by Faber {\it et al.} with two free parameters $a$ and $b$ is \cite{faber}: \begin{equation} \alpha(x)=\frac{\alpha_{\mathrm {max}}}{2}(1-\tanh(ay(x)+\frac{b}{R})), \label{tanh} \end{equation} where $R$ is the space-extent of the Wilson loop with finite time-extent and $x$ represents the position of the vortex core. $a$ is proportional to the inverse of the vortex thickness and the parameter $b$ introduces a dependency on the space extent $R$ of the Wilson loop into the vortex profile. $y(x)$ is \begin{equation} y(x)=\begin{cases} x-R \quad\mathrm{for}\quad|R-x|\le|x|\\ -x ~~~\quad\mathrm{for}\quad|R-x|> |x|. \end{cases} \end{equation} \item A similar profile with only one parameter $\acute{a}$ which corresponds to the inverse of the vortex thickness, was introduced by Faber {\it et al.} \cite{denis}: \begin{equation} \alpha(x)=\frac{\alpha_{\mathrm {max}}}{2}(\tanh({\acute{a}}(x+\frac{R}{2}))-\tanh({\acute{a}}(x-\frac{R}{2})). \label{newtanh} \end{equation} The parameter $b$ has been removed because lattice calculations do not show the influence of the Wilson loop on the vortices. It should be noticed that $\acute{a}$ is not the same as $a$ in Eq.~(\ref{tanh}). \item The following profile was introduced by Deldar in Ref.~\cite{deldar-jhep}: \begin{equation} \alpha (x)=\beta (x)-\beta (x-R), \label{newflux} \end{equation} where $\beta (x)$ -the amount of the vortex flux contained in different regions- is given by: \begin{equation} \beta(x)= \begin{cases} \frac{\alpha_{\mathrm {max}}}{2} & \quad x\geq a\\ \frac{\alpha_{\mathrm {max}}}{2}\big(1-\exp[b(1-\frac{1}{(\frac{x}{a}-1)^2})]\big) & \quad 0\le x \le a\\ -\frac{\alpha_{\mathrm {max}}}{2}\big(1-\exp[b(1-\frac{1}{(\frac{x}{a}+1)^2})]\big) & \quad -a\le x \le 0\\ -\frac{\alpha_{\mathrm {max}}}{2} & \quad x\leq -a. \end{cases} \label{beta} \end{equation} By varying the free parameters $a$ and $b$, the shape of the profile changes. For instance, $a=150$ and $b=10$ lead to a profile similar to Eq.~(\ref{tanh}) while changing the parameter $b$ to $0.01$ turns the profile to a delta function which violates Casimir scaling. \end{enumerate} In Sec.~\ref{sub:1}, $\alpha_{\mathrm{max}}$ for the two vortices of the $SU(4)$ gauge group -using the upper limit criterion- is calculated as shown in Eq.~(\ref{alpha-max}). \subsection{\label{sub:1}Calculating the maximum flux} As mentioned in Sec.~\ref{sec:2}, there are $N-1$ types of vortices in an $SU(N)$ gauge group, but not all of them are independent. In fact, the vortices of types $n$ and $N-n$ are conjugate and the fluxes carried by them are in the opposite directions. So, in the $SU(4)$ gauge group, two types of center vortices fill the QCD vacuum. Now from Eq.~(\ref{alpha-max}), the maximum flux of the two vortices of this gauge group can be calculated. For the $z_{1}=\exp(\frac{\pi i}{2})$ vortex of the $SU(4)$, one may write: \begin{equation} \exp\big[i \vec{\alpha}^{(1)}.\vec{H}\big] =\exp (\frac{\pi i}{2})I. \label{grmax-su4} \end{equation} The cartan generators of the $SU(4)$ are as the following: \begin{eqnarray} H_1&=&\frac{1}{2} \mathrm{diag}(1,-1,0,0), \\ H_2&=&\frac{1}{2\sqrt{3}}\mathrm{diag}(1,1,-2,0),\\ H_3&=&\frac{1}{2\sqrt{6}}\mathrm{diag}(1,1,1,-3). \label{H-su4} \end{eqnarray} If the three of the above matrices are used, the calculations go as follows: \begin{equation} \exp \big[i{\alpha}_1^{(1)} H_1 + i{\alpha}^{(1)}_2 H_2 + i{\alpha}^{(1)}_3 H_3\big]=\exp(\frac{\pi i}{2})I, \label{alpha-su4} \end{equation} where upper index $(1)$ represents the vortex type and lower indices imply the projection of $\vec{\alpha}_{\mathrm{max}}$ on the directions of the cartan generators. It should be noticed that the $\mathrm{max}$ indices for the projection parts have been omitted. Therefore, \begin{equation} {\alpha}_1^{(1)}=-2\pi \quad , \quad {\alpha}_2^{(1)}=-\frac{2\pi}{\sqrt{3}} \quad , \quad {\alpha}_3^{(1)}=-\frac{2\pi}{\sqrt{6}}. \label{all-su4-max} \end{equation} Another possible way is to use only $H_3$ matrix: \begin{equation} \exp \big[i{\alpha}^{(1)}_{\mathrm{max}} H_3 \big]=\exp(\frac{\pi i}{2})I\Longrightarrow {\alpha}^{(1)}_{\mathrm{max}}=\pi\sqrt{6}. \label{z1-max} \end{equation} Both of the above normalization methods lead to the same result as the factor ${\cal G}_r[\vec{\alpha}]$ includes all the center elements of the gauge group in both cases (see Fig.~\ref{fig:trace}). However, if one uses only $H_1$ or $H_2$ in the calculations, it is impossible to get all the center elements. \begin{figure} \includegraphics[scale=0.7]{Tr.eps} \caption{The imaginary part of ${\cal G}_r[\alpha]$ versus the real part in the $SU(4)$ gauge group for $z_1=\exp(\frac{\pi i}{2})$, $z_2=\exp(\pi i)$, $z_3=\exp(\frac{3\pi i}{2})$ vortices and the vacuum trivial structure. In the calculation of $\mathrm{Re}{\cal G}_r[\alpha]$ and $\mathrm{Im}{\cal G}_r[\alpha]$, the vortex flux given in Eq.~(\ref{tanh}) with $a=0.05$ and $b=4$ and the Cartan generator $H_3$ have been used. It is observed that in case of using only $H_3$, the group factor can produce the four center elements of this gauge group- including the trivial element.} \label{fig:trace} \end{figure} The same method could be used to calculate the maximum flux of the $z_2$ vortex. I recall that using only $H_3$ in the calculation gives the same result as using the three diagonal generators of the $SU(4)$. Thus, the second normalization method given in Eq.~(\ref{z1-max}) is used: \begin{equation} \exp \big[i{\alpha}^{(2)}_{\mathrm{max}} H_3 \big]=\exp(\pi i)I\Longrightarrow {\alpha}^{(2)}_{\mathrm{max}}=2 \pi \sqrt{6}. \label{z2-max} \end{equation} Now Cartan generators should be calculated for the higher representations which is done in the next section. \subsection{\label{sub:2}Higher representations of the $SU(4)$ gauge group} To calculate ${\cal G}_r[\vec \alpha]$ for higher representations, the corresponding $H_3$ matrix should be obtained from \cite{georgi} \begin{equation} {\big(H_3^{D_1\otimes D_2}\big)}_{ix,iy}=\big (H_3^{D_1}\big) \delta_{xy}+\delta_{ij} \big (H_3^{D_2}\big), \label{higher} \end{equation} where $H_3^a$s are the $SU(4)$ generators for $D_1\otimes D_2$, $D_1$ and $D_2$ representations, respectively. If ${X_r^i;i=1,2,...,d_r}$ is considered as the vector basis for the representation $d_r$, the elements of $H_3^a$ could be found by: \begin{equation} H_3^rX_r^i=\sum_{j=1}^{d_r} C_{ij} X_r^j. \label{H-high} \end{equation} The aim of this paper is to investigate the behavior of the group factor for higher representations. Now we find a few representations which were discussed in the previous works \cite{deldar2005, deldar2007, deldar2009}. Then by finding the basis vectors of them, the $H_3$ generator of the higher representations is obtained. \begin{equation} 4 \otimes \bar4=\bm {15} \oplus 1, \label{adjoint} \end{equation} \begin{equation} 4 \otimes 4=\bm {10} \oplus \bm {6}, \label{diquark} \end{equation} \begin{equation} 4 \otimes 4 \otimes 4=\bm {20_s} \oplus 20_m \oplus 20_m \oplus \bar4, \label{20} \end{equation} \begin{equation} 4 \otimes 4 \otimes 4 \otimes 4=3 \times {45} \oplus \bm {35_s} \oplus 2 \times 20_{\Box} \oplus 3 \times 15 \oplus 1. \label{35} \end{equation} It should be mentioned that the behavior of the representations written in bold is investigated in this paper. If $v^i$ and $v_j$ are considered as the basis vectors for representations $4$ and $\bar 4$, respectively, the basis vectors for the higher representations -mentioned above- could be calculated from tensor calculus. More details for the basis vectors of each representation are given in the appendix. From Eq.~(\ref{H-high}) and the basis vectors, the $H_3$ generator for the above representations is calculated.\\ Representation $6$: \begin{equation} H_3^6=\frac{1}{\sqrt6}diag\big(1,1,-1,1,-1,-1\big), \label{H6} \end{equation} Representation $10$: \begin{equation} H_3^{10}=\frac{1}{\sqrt6}diag\big(1,1,1,-1,1,1,-1,1,-1,-3\big), \label{H10} \end{equation} Representation $15$: \begin{equation} H_3^{15}=\frac{2}{\sqrt6}diag\big(0,0,0,1,0,0,0,1,0,0,0,1,-1,-1,-1\big), \label{H15} \end{equation} Representation $20_s$: \begin{widetext} \begin{equation} H_3^{20_s}=\frac{1}{2\sqrt6}diag\big(3,3,3,-1,3,3,-1,3,-1,3,3,-1,3,-1,-5,3,-1,-5,-5,-9\big), \label{H20} \end{equation} \end{widetext} Representation $35_s$: \begin{widetext} \begin{equation} H_3^{35_s}=\frac{2}{\sqrt6}diag\big(1,1,1,0,1,1,0,1,0,-1,1,1,0,1,0,-1,1,0,-1,-2,1,1,0,1,0,-1,1,0,-1,-2,1,0,-1,-2,-3\big). \label{H35} \end{equation} \end{widetext} In Sec.~\ref{results}, the plots of $Re{\cal G}_r(\vec{\alpha})$ are investigated for different quark separations and the place of the minimum points observed in the plots is discussed. \section{\label{results}Results and discussions} As it was mentioned previously, ${\cal G}_r(\vec{\alpha})$ plays an important role in the potential between quarks. This factor changes between $1$ and $\exp (\frac{2 \pi i n k}{N})$ corresponding to the type of vortex $n$ and N-ality=$k$ of the representation. Depending on the representations, the way ${\cal G}_r(\vec{\alpha})$ changes between these two limits differs which means in some representations such as $15$, $20_s$ and $35_s$, the group factor passes the phase factor. Fig.~\ref{fig:both} shows the real part of the $SU(4)$ group factor versus the position of the $z_1$ vortex core for the fundamental and diquark representations with $4$-ality equal to $1$ and $2$, respectively. The Wilson loop legs are located at $x=-50$ and $x=50$ so $R=100$. The flux used to produce this figure is the same as in Eq.~(\ref{newtanh}) and the parameter $\acute{a}$ is chosen equal to $0.04$, hence the vortex thickness -which is proportional to $\frac{1}{a}$- is equal to 25. As we expect, $Re{\cal G}_r({\alpha})$ varies between 1 -when there is no overlap between the vortex and the Wilson loop- and the real part of the phase factor $\exp (\frac{ \pi i k}{2})$ -when the vortex is entirely enclosed by the Wilson loop. Therefore, $Re{\cal G}_r({\alpha})$ behaves as expected for these two representations. But the results are different for higher representations. \begin{figure} \includegraphics[scale=0.7]{both.eps} \caption{\label{fig:both} $Re{\cal G}_r({\alpha})$ versus the position of the vortex core $x$ for the $z_1$ vortex of the $SU(4)$ and for diquark and fundamental representations of this gauge group using the flux given in Eq.~(\ref{newtanh}) with $\acute {a}=0.04$ and $R=100$. The vortex thickness is equal to $25$ so when $x=0$, the vortex is completely inside the Wilson loop. As it is expected, the group factor changes between $1$ and $0$ for the fundamental representation and between $1$ and $-1$ for the diquark one. Therefore, for these representations $Re{\cal G}_r({\alpha})$ behaves normally.} \end{figure} Fig.~\ref{fig:gr20} is the same as Fig.~\ref{fig:both} but for representation $20_s$ with $4$-ality=$1$ and for two values of the space extent $R$. The group factor is expected to alter between $1$ and $0$ for $R=100$. But as it is seen in this figure, the two minimum values are observed at $x=-45$ and $x=45$ which have a value equal to $-0.33$. So $Re{\cal G}_r({\alpha})$ has passed $0$ which is the real part of the phase factor for representation $20_s$. Now the influence of the vortex on the Wilson loop is investigated at different positions to find out how the group factor changes. The center vortex theory states that the Wilson loop is unaffected when there is no interaction between the vortex and the Wilson loop hence $Re{\cal G}_r({\alpha})=1$. The effect of the vortex on the Wilson loop starts when it overlaps the perimeter of the loop. If the position of the center of the vortex core is placed at $x=-45$, about $60\%$ of the maximum flux enters the Wilson loop. In this case: \begin{equation} Re{\cal G}_r(\vec\alpha)=Re\frac{1}{d_r} Tr \exp\big[\frac{3i}{5}\alpha_{\mathrm max}\times H_3^{(20)} \big]=-0.33, \label{G-half} \end{equation} \begin{figure} \includegraphics[scale=0.7]{gr20.eps} \caption{\label{fig:gr20} $Re{\cal G}_r({\alpha})$ versus the position of the $z_1$ vortex core $x$ for representation $20_s$ and for $R=40$ and $R=100$. The flux used to produce this figure is given in Eq.~(\ref{newtanh}) with $\acute {a}=0.04$. When $R=100$, two minimum values are observed at $x=-45$ and $x=45$ where about $60\%$ of the maximum vortex flux enters the Wilson loop. The value of $Re{\cal G}_r({\alpha})$ at the minimum points is equal to $-0.33$. The vortex thickness is equal to $25$ so when the position of the vortex core is located at $x=0$, the vortex is completely inside the Wilson loop. Thus, $Re{\cal G}_r({\alpha})=1$ and one may conclude that the asymptotic string tension is produced. When the Wilson loop spatial extent is decreased to $40$ and the location of the vortex core is at the middle of the Wilson loop, $Re{\cal G}_r(\vec{\alpha})$ is near $-0.33$. So $(Re{\cal G}_r(\alpha))_{\mathrm {min}}$ at large distances, is equal to the center of the group factor at intermediate distances.} \end{figure} and the minimum point of the group factor is created. The vortex is completely inside the Wilson loop when the position of the center of the vortex is placed at $x=0$. So according to Eq.~(\ref{W}), the Wilson loop of representation $20_s$ is multiplied by a phase factor $\exp (\frac{\pi i k}{2})$ where $k=1$. Therefore, we expect at this point that $Re{\cal G}_r(\alpha)$ is equal to $0$. It means that the Wilson loop and the vortex are totally linked to each other and the asymptotic string tension is achieved. As the vortex starts leaving the Wilson loop, a part of the vortex flux quits the loop; So that when the position of the vortex core is placed at $x=45$, $60\%$ of the maximum flux carried by the vortex still remains in the Wilson loop. So again the value of $Re{\cal G}_r(\alpha)$ becomes equal to $-0.33$. Now the distance between quark and anti-quark is decreased to produce the intermediate potential. In Fig.~\ref{fig:gr20}, it is seen when $R=40$ and the vortex core is located at $x=0$, which is the middle point of the Wilson loop, the value of the group factor is equal to $-0.33$. At this stage, the intermediate string tension is produced. Strictly speaking, the value of the group factor when the position of the vortex core is located at the middle of the Wilson loop determines the potential behavior and it is called the "center of the group factor" in this paper. Therefore, one may conclude that the minimum values seen for $R=100$ in Fig.~\ref{fig:gr20} are the points which the center of the group factor reaches at intermediate distances. The most interesting aspect about the group factor is that the value of $Re{\cal G}_r(\alpha)$ at the minimum points is independent of the flux used in the calculation. This issue is understood by a comparison between Fig.~\ref{fig:gr20} and the three curves in Fig.~\ref{fig:new20} where the flux in Eq.~(\ref{newflux}) has been used with $a=150$ and three values for the parameter $b$. The space extent $R$ of the Wilson loop is equal to $250$ so that the center of the group factor reaches $0$ for the three curves. It is observed that value of $(Re{\cal G}_r(\alpha))_{\mathrm {min}}$ remains the same even when the shape and the type of the vortex flux change. The similar result is produced if the flux in Eq.~(\ref{tanh}) is investigated. \begin{figure} \includegraphics[scale=0.7]{new20.eps} \caption{\label{fig:new20}The same as Fig.~\ref{fig:gr20} but for the flux given in Eq.~(\ref{newflux}) with $a=150$, $b=10$, $b=1$, $b=0.1$ and $R=250$. The two minimum points are observed in all three curves and $(Re{\cal G}_r({\alpha}))_{\mathrm{min}}=-0.33$. The minimum points are located at the positions where $60\%$ of the maximum flux of the $z_1$ vortex enters the Wilson loop.} \end{figure} Another fact about the group factor is that the portion of the vortex flux which enters the Wilson loop at the minimum points, is different for representations with different N-ality. This matter makes us investigate the behavior of $Re{\cal G}_r(\alpha)$ for the adjoint representation with $4$-ality=$0$. The real part of the group factor of the adjoint representation versus $x$ for $R=100$ is observed in Fig.~\ref{fig:gr15} using the same flux as of Fig.~\ref{fig:gr20}. When the $z_1$ vortex core is located at $x=-50$ and $x=50$, half of the maximum flux is in the Wilson loop and at these points $Re{\cal G}_r({\alpha})=0.2$. \begin{figure} \includegraphics[scale=0.7]{gr15.eps} \caption{\label{fig:gr15} The same as Fig.~\ref{fig:gr20} but for the adjoint representation. Two minimum values are observed at $x=-50$ and $x=50$. The value of $Re{\cal G}_r(\vec{\alpha})$ at the minimum points is equal to $0.2$. These points are the positions where half of the vortex maximum flux enters the Wilson loop. When the position of the vortex core is located at $x=50$, $Re{\cal G}_r({\alpha})=1$ and one may conclude that the asymptotic string tension is produced.} \end{figure} The behavior of $Re{\cal G}_r(\alpha)$ for the $z_2$ vortex can be analyzed in the similar way. Fig.~\ref{fig:gr152} is the same as Fig.~\ref{fig:gr15} but for the $z_2$ vortex. In this figure, $R$ has changed to $130$ so that the center of the group factor reaches $1$ for the $z_2$ vortex which contains two $z_1$ vortices. When the position of the $z_2$ vortex core is placed at $x=-79$ and $x=78$, one fourth of the $z_2$ vortex flux is in the Wilson loop. This amount of flux is equivalent to half of the maximum flux carried by one $z_1$ vortex. Thus one can expect -from the above discussion for the $z_1$ vortex- that the amount of $Re{\cal G}_r(\alpha)$ is equal to $0.2$. At $x=-65$ and $x=64$, half of the $z_2$ vortex flux or equivalently the total flux of one $z_1$ vortex is in the Wilson loop. Therefore, $Re{\cal G}_r(\alpha)$ reaches $1$. When $x=-52$ and $x=51$, three fourth of the $z_2$ vortex enter the Wilson loop which is equal to half of the flux carried by one $z_1$ vortex and $Re{\cal G}_r(\alpha)$ equals to $0.2$. The $z_2$ vortex is completely inside the loop when $x=0$ and the Wilson loop changes by a phase factor $(\exp \pi i k)$ with $k=0$ hence $Re{\cal G}_r(\alpha)$ becomes equal to $1$. The behavior of the group factor of representation $35_s$ is like as of the adjoint one. For this representation, ${(Re{\cal G}_r(\alpha))}_{\mathrm min}=-0.25$. Thus, one might claim that the number of the quarks and anti-quarks which make the representation changes the portion of the remaining flux inside the Wilson loop at the minimum points. For more investigation, we look at Eqs.~(\ref{adjoint}), (\ref{20}) and (\ref{35}). It is observed that representation $20_s$ is built from three quarks while four quarks form representations $35_s$. Also, one quark and one anti-quark make the adjoint representation. On the other hand, for zero $4$-ality representations built from even numbers of quarks and anti-quarks, the minimum points in the plots are the places where half of the maximum vortex flux is in the Wilson loop while for higher representations which are produced from an odd number of quarks, the minimum points are located at the positions where $60\%$ of the maximum flux is in the loop. One may find the reason of this behavior in the difference between the slope of the potentials at intermediate and large distances. \begin{figure} \includegraphics[scale=0.7]{gr152.eps} \caption{The same as Fig.~\ref{fig:gr15} but for the $z_2$ vortex. For $x=-79$ and $x=78$, one fourth of the maximum flux of the $z_2$ vortex is in the Wilson loop which is equal to half of the maximum flux of the $z_1$ vortex. So, as in Fig.~\ref{fig:gr15}, $Re{\cal G}_r(\vec{\alpha})=0.2$ and analogue for $x=-52$ and $x=51$ where the three fourth of the maximum vortex flux are in the loop. The Wilson loop and the vortex are linked completely at $x=0$ and the value of $Re{\cal G}_r(\vec{\alpha})$ is equal to $1$ at this point.} \label{fig:gr152} \end{figure} In our previous works \cite{deldar2005, deldar2007}, it was shown that at intermediate distances, the ratio of the string tension of each representation to that of the fundamental one is proportional to Casimir scaling. At large distances, the representations with the same $4$-ality get the same slope. The first row of Tab.~\ref{tab} shows the representations along with the corresponding Casimir scaling in the second row. The third row represents the $4$-ality of each representation. Thus, for representations $15$ and $35_s$ screening is observed. The potential of representation $20_s$ becomes parallel to the one of the fundamental representation and also diquark representations $(6,10)$ get a parallel slope. Now, the string tensions of the representations which belong to the same $4$-ality class can be compared. It was mentioned before that $Re{\cal G}_r(\alpha)$ behaves normally for the fundamental representation. It varies between $1$ and $0$ for the $z_1$ vortex and between $1$ and $-1$ for the $z_2$ one. As the same description works for each vortex type, the $z_1$ vortex is focused. Moreover, although the contribution of the $z_2$ vortex to the potential is not negligible, it is less than the contribution of the $z_1$ vortex \cite{diakonov}. Tab.~\ref{tab} shows that the string tension of representation $20_s$ at intermediate distances is larger than the fundamental one. Therefore, the center of $Re{\cal G}_r(\vec\alpha)$ passes $0$ and reaches $-0.33$ to produce an intermediate string tension larger than that of the fundamental one. Then it goes back to $0$ and the same asymptotic string tension as the fundamental one is produced. The center of the group factor has the same value equal to $0$ for both representations $15$ and $35_s$. However, at intermediate distances, it is equal to $0.2$ for the adjoint representation. This factor reaches $-0.25$ for representation $35_s$ so that the intermediate string tension of this representation becomes larger than the adjoint one. For the diquark representations, $Re{\cal G}_r(\alpha)$ changes normally between $1$ and $-1$ for the $z_1$ vortex. From Tab.~\ref{tab} it is seen that the intermediate string tensions of the diquark representations are in the same range. So the group factor is expected to behave the same way for these two representations. \begin{table} \caption{\label{tab}Casimir scaling and $4$-ality of some SU(4) representations. the behavior of the group factor for representations with the same $4$-ality can be compared. } \begin{ruledtabular} \begin{tabular}{ccccccc} \textrm{representation}& \textrm{4}& \textrm{6}& \textrm{10}& \textrm{$15_s$(adjoint)}& \textrm{$20_s$}& \textrm{$35_s$}\\ \colrule $\frac{C_r}{C_F}$ & 1 & 1.33 & 2.13 & 2.4 & 4.2 & 6.4 \\ $4$-ality & 1 & 2 & 2 & 0 & 1 & 0\\ \end{tabular} \end{ruledtabular} \end{table} \section{\label{con}Conclusion} The prediction of the potential between $SU(N)$ static color sources by the thick center vortices model has been very precise in various quark distances. In other words, both the intermediate and asymptotic string tensions are in agreement with the proposals given in the corresponding regions, e.g. N-ality dependence at large distances and the Sine and Casimir scaling at intermediate distances. In fact, The behavior of the string tensions depends on how the $SU(N)$ group factor interpolates between $1$ and the phase factor $\exp(\frac{2\pi i nk}{N}$) where $k$ is the N-ality of the representation and $n$ represented the vortex type. It is seen that the real part of the $SU(N)$ group factor changes abnormally for some representations. In the $SU(4)$ gauge group, $Re{\cal G}_r(\alpha)$ for representations $15$, $20_s$ and $35_s$, ... passes the corresponding phase factor while this factor interpolates normally between expected limits for fundamental and diquark representations. The abnormal behavior of the group factor is observed for representations $8$,$10$, $15_m$, $15_s$, ... in the $SU(3)$ gauge group. In this paper, I have shown that the minimum points seen in the plots of the group factor versus $x$ at large distances are located at the points where the vortex and the Wilson loop are partially linked. It means that when the position of the vortex core is placed at the minimum points, a part of the vortex flux enters the Wilson loop. The sub-Wilson loops -which make the representation $r$- affect this portion in a way that $50\%$ of the maximum flux enters the Wilson loop of the representations with N-ality=$0$. This portion turns to $60\%$ for representations which belong to N-ality=$1$ class except the fundamental representation. My calculations show that at very short distances, the center of the group factor is nearly equal to $1$ which means the vortex and the Wilson loop are slightly linked to each other. As the distance increases, this amount becomes near to $(Re{\cal G}_r(\vec\alpha))_{\mathrm{min}}$ at intermediate distances. When the quarks are separated more, the group factor center changes in a way that it gets equal to the phase factor and at this stage the asymptotic string tensions are achieved. Therefore, one might conclude that the minimum points are the positions which the group factor reaches at intermediate distances. The comparison between the asymptotic and the intermediate string tensions of the representations with the same N-ality shows that the group factor must pass the phase factor so that an intermediate string tension larger than the asymptotic one is achieved. This interpretation works for diquark representations ($6$ and $10$) which have nearly the same asymptotic and intermediate string tensions. \begin{acknowledgments} I would like to thank S. Deldar for her helpful discussions. \end{acknowledgments}
\section{Introduction} Image analysis is a scientific discipline providing theoretical foundations and methods for solving problems appearing in a range of areas as diverse as chemistry, physics, biology, geography, medicine, astronomy, robotics and industrial manufacturing. Besides traditional approaches based on continuous models, which require numeric computation, and which always involve the problem of rounding and approximation, ``combinatorial'' approaches to image analysis (also named ``digital'' or ``discrete'' approaches) have been developed during the last 60 years. These latter approaches are based on studying combinatorial properties of the digital data sets under consideration and generally providing useful algorithms for image analysis which are more efficient and accurate than those based on continuous models. In order to cope with the variety of image processing and computer vision challenges, several techniques have been introduced and developed, quite often with great success. Among the different techniques that are currently in use, there are, for example, soft computing techniques. Soft computing is an emerging field that consists of complementary elements of fuzzy logic, neural computing, evolutionary computation, machine learning and statistical reasoning, and often offers solutions where conventional approaches fail. Segmentation is the fundamental process which partitions a data space into meaningful salient regions. Image segmentation essentially affects the overall performance of any automated image analysis system. Thus, its quality is of the utmost importance. Image regions, homogeneous with regard to some usually statistical criterion or color measure, which result from a segmentation algorithm are analyzed in subsequent interpretation steps. Statistical criterion based image segmentation has been an area of intense research activity during the past forty years and many algorithms were published in consequence of all this effort, starting from simple thresholding methods up to the most sophisticated random field type methods. Unsupervised methods which do not assume any prior scene knowledge which can be learned to help segmentation processes are obviously more challenging than the supervised ones. The mean shift (MSH) is a non-parametric procedure that has demonstrated to be an extremely versatile tool for feature analysis. It can provide reliable solutions for many computer vision tasks \cite{Comaniciu02}. The mean shift method was proposed in 1975 by Fukunaga and Hostetler \cite{Fukunaga75}. It was largely forgotten until Cheng's paper rekindled interest in it \cite{Cheng95}. Unsupervised segmentation by means of the mean shift method carries out as a first step a smoothing filter before segmentation is performed \cite{Comaniciu02}. Mean shift iterative algorithm was proposed in 2006 and this has been performed in many works by using the entropy as a stopping criterion \cite{Rodriguez11,Rodriguez11a,Rodriguez12,Rodriguez08,Dominguez11}. The term of entropy is not a new concept in the field of information theory. Entropy has been used in image restoration, edge detection and recently as an objective evaluation method for image segmentation \cite{Zhang03}. The novelty of the proposed algorithm is the use of the entropy as a stopping criterion. The choice of entropy as a measure of goodness deserves several observations, which will be detailed in next section. This paper proposes a new stopping criterion for the MSH iterative algorithm, where stopping threshold via entropy is used, now, in another way. Many segmentation experiments, by utilizing standard images, were carried out using this new stopping criterion. This paper compares the stability of MSH iterative algorithm using the new stopping criterion with regard to the old stopping criterion used in \cite{Rodriguez11,Rodriguez11a,Rodriguez12,Rodriguez08,Dominguez11}. Good segmentation was reached and the algorithm had better stability. An analysis on the convergence, through a theorem, with the new stopping criterion was carried out. The remainder of the paper is as follows. In Section \ref{THEORICAL ASPECTS}, the more significant theoretical aspects of the mean shift and entropy are given. Section \ref{THE MEAN SHIFT ITERATIVE ALGORITHM} describes our MSH iterative algorithm with the old and new stopping criterion. In this section the theorem that ensures the convergence is proposed. The experimental results, comparisons and discussion are presented in Section \ref{EXPERIMENTAL RESULTS. ANALYSIS AND DISCUSSION}. Finally, in Section \ref{CONCLUSIONS} the conclusions are given. \section{THEORICAL ASPECTS} \label{THEORICAL ASPECTS} \subsection{Mean Shift} The basic concept of the mean shift algorithm is as follows: Let $x_{i}$ be an arbitrary set of $n$ points in the $d$ dimensional space. The kernel density estimation $f(x)$ is obtained by means of the kernel function $K(x)$ and window radius $h$. Function $f(x)$ is defined as \begin{eqnarray} f(x) =\frac{1}{nh^d}\sum\limits_{i=1}^{n}K\left( \frac{x-x_{i}}{h}\right) . \end{eqnarray} Here, the {\itshape Epanechnikov} function is chosen as the kernel function. The {\itshape Epanechnikov} function is defined as, \begin{align} K_E(x)=\left\{ \begin{array}{ll} \frac{1}{2} c^{-1}_{d} (d+2)\left( 1-\left\|x\right\|^2\right), & \mbox{if}\ \left\|x\right\|<1\\ & \\ 0, & \mbox{otherwise}.\\ \end{array} \right. \end{align} The differential function $f(x)$ is formulated as \begin{eqnarray} \widehat{\nabla} f(x) = \nabla \widehat{f(x)} =\frac{1}{nh^d}\sum\limits_{i=1}^{n}\widehat K\left( \frac{x-x_{i}}{h}\right) , \end{eqnarray} \begin{align} \label{label4} \widehat{\nabla}f_E(x)&= \frac{1}{n(h^d c_d)}\frac{d+2}{h^2}\sum\limits_{x_i \in S_h(x)} (x_i-x)\nonumber\\ &=\frac{n_x}{n(h^d c_d)}\frac{d+2}{h^2}\frac{1}{n_x}\sum\limits_{x_i \in S_h(x)} (x_i-x), \end{align} where region $S_{h}(x)$ is a hyper sphere of radius $h$ having volume $h^{d} c_{d}$, centred at $x$, and containing $n$ data points; that is, the uniform kernel. In addition, in this case $d=3$, for the $x$ vector of three dimensions, two for the spatial domain and one for the range domain (gray levels). The last factor in expression (\ref{label4}) is called the sample mean shift, \begin{align} \label{label5} M_{h,U}(x)&=\frac{1}{n_x}\sum\limits_{x_i \in S_h(x)}(x_i-x)\nonumber\ \\ &=\left(\frac{1}{n_x}\sum\limits_{x_i \in S_h(x)}x_i\right)-x. \end{align} The quantity $\frac{n_x}{n(h^{d} c_{d})}$ is the kernel density estimate $\widehat{f_U}(x)$ (where $U$ means the uniform kernel) computed with the hyper sphere $S_{h}(x)$, and thus we can write the expression (\ref{label4}) as: \begin{eqnarray}\label{label6} \widehat{\nabla}f_{E}(x)=\widehat{f_{U}}(x)\frac{d+2}{h^2} M_{h,U}(x) \end{eqnarray} which yields, \begin{eqnarray}\label{label7} M_{h,U}(x)=\frac{h^2}{d+2}\frac{\widehat{\nabla}f_E(x)}{\widehat{f_{U}}(x)}. \end{eqnarray} Equation (\ref{label7}) shows that an estimate of the normalized gradient can be obtained by computing the sample mean shift in a uniform kernel centered on $x$. In addition, the mean shift has the gradient direction of the density estimate at $x$ when this estimate is obtained with the {\itshape Epanechnikov kernel}. Since the mean shift vector always points towards the direction of the maximum density increase, it can define a path leading to a local density maximum; that is, to the density mode. In \cite{Comaniciu00}, it was proved that the obtained {\itshape mean shift procedure} by the following steps, guarantees the convergence: \begin{itemize} \item computing the mean shift vector $M_h(x)$ \item translating the window $S_h(x)$ by $M_h(x)$ \end{itemize} Therefore, if the individual mean shift procedure is guaranteed to converge, a recursively procedure of the mean shift also converges. Other related works with this issue can be seen in \cite{Suyash06,Grenier06}. \subsection{Entropy} From the point of view of digital image processing the entropy of an image is defined as: \begin{eqnarray} \label{entropy} E=-\sum\limits_{x=0}^{2^B-1}p_{x} \log_2{p_{x}}, \end{eqnarray} where $B$ is the total quantity of bits of the digitized image and by agreement $\log_2(0)=0$ ; $p(x)$ is the probability of occurrence of a gray-level value. Within a totally uniform region, entropy reaches the minimum value. Theoretically speaking, the probability of occurrence of the gray-level value, within a uniform region is always one. In practice, when one works with real images the entropy value does not reach, in general, the zero value. This is due to the existent noise in the image. Therefore, if we consider entropy as a measure of the disorder within a system, it could be used as a good stopping criterion for an iterative process, by using the mean shift iterative algorithm. More goodness on entropy can be seen in \cite{Rodriguez11a,Zhang03}. \section{THE MEAN SHIFT ITERATIVE ALGORITHM} \label{THE MEAN SHIFT ITERATIVE ALGORITHM} The Mean Shift Iterative Algorithm (MSHIA) with the old stopping criterion is composed of the following steps. Let $ent1$ be the initial value of the entropy of the first iteration. Let $ent2$ be the second value of the entropy after the first iteration. Let $errabs$ be the absolute value of the difference of entropy between the first and the second iteration. Let $edsEnt$ be the threshold to stop the iterations; that is, to stop when the relative rate of change of the entropy from one iteration to the next, falls below this threshold. Then, the segmentation algorithm comprises the following steps: \begin{algorithm}[h] \caption{Algorithm with the Old Stopping Criterion}\label{AlgoritmoOld} Initialize $ent2$, $errabs$ and $edsEnt$\; While $errabs > edsEnt$, then\; Filter the image according to the mean shift algorithm; store in $Z[k]$ the filtered image\; Calculate the entropy from the filtered image according to expression (\ref{entropy}); store in ent1\; Calculate the absolute difference with the entropy value obtained in the previous step; $errabs = \vert ent1 - ent2 \vert$\; Update the value of the parameters an image; $ent2 = ent1$; $Z[k +1] = Z[k]$\; \end{algorithm} \begin{algorithm}[h] \caption{The Algorithm with the New Stopping Criterion}\label{AlgoritmoNew} Initialize $edsEnt$ and $B1$, let $B1$ be equal to original image\; While $errabs > edsEnt$, then\; Filter the original image according to the mean shift algorithm; store in $B2$ the filtered image\; Calculate the absolute difference with the image obtained in the previous step; $C = \vert B2 - B1\vert$\; Calculate the entropy to the previous image; store in $errabs$\; Update the images: $B1 = B2$\; \end{algorithm} One can observe that with the new stopping criterion the entropy is used in another way. This paper will prove that with the new stopping criterion the MSHIA offers greater stability. Moreover, the following theorem resulted very interesting. \begin{theorem} \label{theorem1} When the entropy of the absolute difference between the iteration image and the following is taken as a stopping criterion in the mean shift iterative algorithm, whatever the chosen threshold this is enough condition to achieve the convergence. In addition, at the limit the entropy is zero. \end{theorem} This theorem ensures the convergence of the MSHIA with the new stopping criterion and determines what happens with entropy at the limit. The proof of this theorem can be found in the appendix. \section{EXPERIMENTAL RESULTS. ANALYSIS AND DISCUSSION} \label{EXPERIMENTAL RESULTS. ANALYSIS AND DISCUSSION} All segmentation experiments were carried out by using a uniform kernel. In order to be effective the comparison between the old stopping criterion and the new stopping criterion, the same value of $h_r$ and $h_s$ in the MSHIA were used. The principal goal of this section is to evaluate the new stop criterion in the MSHIA and to prove the stability of the algorithm with regard to the old stopping criterion. For this reason, comparisons with other segmentation approaches will not be carried out. In \cite{Rodriguez11a} were compared the obtained results with the MSHIA through the old stopping criterion with other segmentation methods. \begin{figure}[ht] \centering \subfigure[Astro]{\label{Figura1a}\includegraphics[width=2.733cm, height=2.75cm]{images/images/AstroOriginal.JPG} }\hspace{-0.1cm} \subfigure[Old criterion]{\label{Figura1b}\includegraphics[width=2.75cm, height=2.75cm]{images/images/AstroOld.JPG}} \subfigure[New criterion]{\label{Figura1c}\includegraphics[width=2.75cm, height=2.75cm]{images/images/AstroNew.JPG}} \subfigure[Barbara]{\label{Figura1d}\includegraphics[width=2.75cm, height=2.75cm]{images/images/BarbaraOriginal.JPG}} \subfigure[Old criterion]{\label{Figura1e}\includegraphics[width=2.75cm, height=2.75cm]{images/images/BarbaraOld.png}} \subfigure[New criterion]{\label{Figura1f}\includegraphics[width=2.75cm, height=2.75cm]{images/images/BarbaraNew.png}} \subfigure[Cameraman]{\label{Figura1g}\includegraphics[width=2.75cm, height=2.75cm]{images/images/CameramanOriginal.JPG}} \subfigure[Old criterion]{\label{Figura1h}\includegraphics[width=2.75cm, height=2.75cm]{images/images/CameramanOld.JPG}} \subfigure[New criterion]{\label{Figura1i}\includegraphics[width=2.75cm, height=2.75cm]{images/images/CameramanNew.JPG}} \subfigure[Lena]{\label{Figura1j}\includegraphics[width=2.75cm, height=2.75cm]{images/images/LenaOriginal.JPG}} \subfigure[Old criterion]{\label{Figura1k}\includegraphics[width=2.75cm, height=2.75cm]{images/images/LenaOld.JPG}} \subfigure[New criterion]{\label{Figura1l}\includegraphics[width=2.75cm, height=2.75cm]{images/images/LenaNew.JPG}} \caption{Used images for segmentation. In the first column are shown the original images; in the second, the segmentation using the old stopping criterion and the third column shows the segmented images using the new stopping criterion.} \label{fig:segmentation} \end{figure} Figure \ref{fig:segmentation} shows the segmentation of the used images. Observe that, in all cases, the iterative mean shift algorithm had better results when using the new stopping criterion. When one compares the segmented images with the old and new criterion it can be observed that with the new stopping criterion (see, for example, central images in Figure \ref{fig:segmentation}), more defined homogeneous zones were obtained. It can be seen that with the old stopping criterion the segmentation gave regions where different gray levels were originated. However, these regions really should have only one gray level. In other zones happened the opposite (see arrow in Figure \ref{Figura1f}). For example, Figure \ref{Figura1e} shows (visually) that the segmentation, with the old stopping criterion, is more diluted (see arrow). However, in Figure \ref{Figura1f} the segmentation with the new stopping criterion has the zones more delimited. One can find this same result when observing the other segmented images. This is one of the principal advantages of the new stopping criterion with regard to the old criterion. These good results are obtained because the defined new stopping criterion through the natural distance between images offers greater stability to the mean shift iterative algorithm. Figure \ref{fig:Profile} shows the profiles of the obtained segmented images by using the two stopping criteria. \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile2/AstroPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/AstroPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile2/AstroPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/AstroPerGrafNew.JPG}} \end{figure} \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile2/BarbaraPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile/BarbaraPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile2/BarbaraPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile/BarbaraPerGrafNew.JPG}} \end{figure} \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile2/CameramanPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile/CameramanPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile2/CameramanPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile/CameramanPerGrafNew.JPG}} \end{figure} \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.2cm, height=4.4cm]{images/profile2/LenaPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/LenaPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.2cm, height=4.4cm]{images/profile2/LenaPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/LenaPerGrafNew.JPG}} \caption{Intensity profiles through the segmented images. Profiles are indicated by lines. First two columns are the segmented images with the old stopping criterion and its profile. The last columns are the segmented images with the new stopping criterion and its profile.} \label{fig:Profile} \end{figure} The plates that appear in Figures \ref{fig:Profile} of the second and the last columns are indicative of equal intensity levels. In both graphics the abrupt falls of intensities to others represent different regions in the segmented images. Note, for example, that in Figures where the segmented image appears with the new stopping criterion there are, in the same region of the segmentation, least variation of the pixel intensities with regard to the segmented images with the old stopping criterion. This illustrates that, in these cases the segmentation was better when the new stopping criterion was used. Figure \ref{Iteration graphs} shows three examples of the performance of the two stopping criterion in the experimental results. In the ``x'' axis appears the iterations of the MSHIA and in the ``y'' axis the obtained values by the stopping criterion in each iteration of the algorithm are shown. \begin{figure}[ht] \centering \subfigure[Barbara]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/BarbaraOld.JPG}} \subfigure[Cameraman]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/CameramanOld.JPG}} \subfigure[Lena]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/LenaOld.JPG}} \subfigure[Barbara]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/BarbaraNew.JPG}} \subfigure[Cameraman]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/CameramanNew.JPG}} \subfigure[Lena]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/LenaNew.JPG}} \caption{Stopping criterion. In the first row appears the performance of the old stopping criterion, while in the second row the performance of the new stopping criterion is shown.} \label{Iteration graphs} \end{figure} The graphics of iterations of the new stopping criterion (second row) show a smooth behavior; that is, the stopping criterion has a better stable performance through the MSHIA. This is the goal of the presentation of these graphics, to observe the oscillations in the ``y'' axis (stability). The new stopping criterion not only has good theoretical properties (see theorem), but also, in practice, has very good behavior. When one analyzes the performance of the old stopping criterion in the experimental results (first row), one can see that the performance in the MSHIA is unstable. In general, this can originate unsuitable segmented images. \section{CONCLUSIONS} \label{CONCLUSIONS} In this work, a new stopping criterion for the MSHIA was proposed. The new stopping criterion is based on obtaining the entropy of the absolute difference between two images. In such sense, a theorem for the convergence was proposed. It was proven, through many experiments by using standard images, that the new stopping criterion had very good performance in the MSHIA. In addition, this was more stable that the old criterion. \section*{Acknowledgment} H. Sossa thanks SIP-IPN and CONACYT for the economical supports under grants 20131182 and 155014, respectively. We all thank the reviewers for their comments on the improvement of this paper. \section{Introduction} Image analysis is a scientific discipline providing theoretical foundations and methods for solving problems appearing in a range of areas as diverse as chemistry, physics, biology, geography, medicine, astronomy, robotics and industrial manufacturing. Besides traditional approaches based on continuous models, which require numeric computation, and which always involve the problem of rounding and approximation, ``combinatorial'' approaches to image analysis (also named ``digital'' or ``discrete'' approaches) have been developed during the last 60 years. These latter approaches are based on studying combinatorial properties of the digital data sets under consideration and generally providing useful algorithms for image analysis which are more efficient and accurate than those based on continuous models. In order to cope with the variety of image processing and computer vision challenges, several techniques have been introduced and developed, quite often with great success. Among the different techniques that are currently in use, there are, for example, soft computing techniques. Soft computing is an emerging field that consists of complementary elements of fuzzy logic, neural computing, evolutionary computation, machine learning and statistical reasoning, and often offers solutions where conventional approaches fail. Segmentation is the fundamental process which partitions a data space into meaningful salient regions. Image segmentation essentially affects the overall performance of any automated image analysis system. Thus, its quality is of the utmost importance. Image regions, homogeneous with regard to some usually statistical criterion or color measure, which result from a segmentation algorithm are analyzed in subsequent interpretation steps. Statistical criterion based image segmentation has been an area of intense research activity during the past forty years and many algorithms were published in consequence of all this effort, starting from simple thresholding methods up to the most sophisticated random field type methods. Unsupervised methods which do not assume any prior scene knowledge which can be learned to help segmentation processes are obviously more challenging than the supervised ones. The mean shift (MSH) is a non-parametric procedure that has demonstrated to be an extremely versatile tool for feature analysis. It can provide reliable solutions for many computer vision tasks \cite{Comaniciu02}. The mean shift method was proposed in 1975 by Fukunaga and Hostetler \cite{Fukunaga75}. It was largely forgotten until Cheng's paper rekindled interest in it \cite{Cheng95}. Unsupervised segmentation by means of the mean shift method carries out as a first step a smoothing filter before segmentation is performed \cite{Comaniciu02}. Mean shift iterative algorithm was proposed in 2006 and this has been performed in many works by using the entropy as a stopping criterion \cite{Rodriguez11,Rodriguez11a,Rodriguez12,Rodriguez08,Dominguez11}. The term of entropy is not a new concept in the field of information theory. Entropy has been used in image restoration, edge detection and recently as an objective evaluation method for image segmentation \cite{Zhang03}. The novelty of the proposed algorithm is the use of the entropy as a stopping criterion. The choice of entropy as a measure of goodness deserves several observations, which will be detailed in next section. This paper proposes a new stopping criterion for the MSH iterative algorithm, where stopping threshold via entropy is used, now, in another way. Many segmentation experiments, by utilizing standard images, were carried out using this new stopping criterion. This paper compares the stability of MSH iterative algorithm using the new stopping criterion with regard to the old stopping criterion used in \cite{Rodriguez11,Rodriguez11a,Rodriguez12,Rodriguez08,Dominguez11}. Good segmentation was reached and the algorithm had better stability. An analysis on the convergence, through a theorem, with the new stopping criterion was carried out. The remainder of the paper is as follows. In Section \ref{THEORICAL ASPECTS}, the more significant theoretical aspects of the mean shift and entropy are given. Section \ref{THE MEAN SHIFT ITERATIVE ALGORITHM} describes our MSH iterative algorithm with the old and new stopping criterion. In this section the theorem that ensures the convergence is proposed. The experimental results, comparisons and discussion are presented in Section \ref{EXPERIMENTAL RESULTS. ANALYSIS AND DISCUSSION}. Finally, in Section \ref{CONCLUSIONS} the conclusions are given. \section{THEORICAL ASPECTS} \label{THEORICAL ASPECTS} \subsection{Mean Shift} The basic concept of the mean shift algorithm is as follows: Let $x_{i}$ be an arbitrary set of $n$ points in the $d$ dimensional space. The kernel density estimation $f(x)$ is obtained by means of the kernel function $K(x)$ and window radius $h$. Function $f(x)$ is defined as \begin{eqnarray} f(x) =\frac{1}{nh^d}\sum\limits_{i=1}^{n}K\left( \frac{x-x_{i}}{h}\right) . \end{eqnarray} Here, the {\itshape Epanechnikov} function is chosen as the kernel function. The {\itshape Epanechnikov} function is defined as, \begin{align} K_E(x)=\left\{ \begin{array}{ll} \frac{1}{2} c^{-1}_{d} (d+2)\left( 1-\left\|x\right\|^2\right), & \mbox{if}\ \left\|x\right\|<1\\ & \\ 0, & \mbox{otherwise}.\\ \end{array} \right. \end{align} The differential function $f(x)$ is formulated as \begin{eqnarray} \widehat{\nabla} f(x) = \nabla \widehat{f(x)} =\frac{1}{nh^d}\sum\limits_{i=1}^{n}\widehat K\left( \frac{x-x_{i}}{h}\right) , \end{eqnarray} \begin{align} \label{label4} \widehat{\nabla}f_E(x)&= \frac{1}{n(h^d c_d)}\frac{d+2}{h^2}\sum\limits_{x_i \in S_h(x)} (x_i-x)\nonumber\\ &=\frac{n_x}{n(h^d c_d)}\frac{d+2}{h^2}\frac{1}{n_x}\sum\limits_{x_i \in S_h(x)} (x_i-x), \end{align} where region $S_{h}(x)$ is a hyper sphere of radius $h$ having volume $h^{d} c_{d}$, centred at $x$, and containing $n$ data points; that is, the uniform kernel. In addition, in this case $d=3$, for the $x$ vector of three dimensions, two for the spatial domain and one for the range domain (gray levels). The last factor in expression (\ref{label4}) is called the sample mean shift, \begin{align} \label{label5} M_{h,U}(x)&=\frac{1}{n_x}\sum\limits_{x_i \in S_h(x)}(x_i-x)\nonumber\ \\ &=\left(\frac{1}{n_x}\sum\limits_{x_i \in S_h(x)}x_i\right)-x. \end{align} The quantity $\frac{n_x}{n(h^{d} c_{d})}$ is the kernel density estimate $\widehat{f_U}(x)$ (where $U$ means the uniform kernel) computed with the hyper sphere $S_{h}(x)$, and thus we can write the expression (\ref{label4}) as: \begin{eqnarray}\label{label6} \widehat{\nabla}f_{E}(x)=\widehat{f_{U}}(x)\frac{d+2}{h^2} M_{h,U}(x) \end{eqnarray} which yields, \begin{eqnarray}\label{label7} M_{h,U}(x)=\frac{h^2}{d+2}\frac{\widehat{\nabla}f_E(x)}{\widehat{f_{U}}(x)}. \end{eqnarray} Equation (\ref{label7}) shows that an estimate of the normalized gradient can be obtained by computing the sample mean shift in a uniform kernel centered on $x$. In addition, the mean shift has the gradient direction of the density estimate at $x$ when this estimate is obtained with the {\itshape Epanechnikov kernel}. Since the mean shift vector always points towards the direction of the maximum density increase, it can define a path leading to a local density maximum; that is, to the density mode. In \cite{Comaniciu00}, it was proved that the obtained {\itshape mean shift procedure} by the following steps, guarantees the convergence: \begin{itemize} \item computing the mean shift vector $M_h(x)$ \item translating the window $S_h(x)$ by $M_h(x)$ \end{itemize} Therefore, if the individual mean shift procedure is guaranteed to converge, a recursively procedure of the mean shift also converges. Other related works with this issue can be seen in \cite{Suyash06,Grenier06}. \subsection{Entropy} From the point of view of digital image processing the entropy of an image is defined as: \begin{eqnarray} \label{entropy} E=-\sum\limits_{x=0}^{2^B-1}p_{x} \log_2{p_{x}}, \end{eqnarray} where $B$ is the total quantity of bits of the digitized image and by agreement $\log_2(0)=0$ ; $p(x)$ is the probability of occurrence of a gray-level value. Within a totally uniform region, entropy reaches the minimum value. Theoretically speaking, the probability of occurrence of the gray-level value, within a uniform region is always one. In practice, when one works with real images the entropy value does not reach, in general, the zero value. This is due to the existent noise in the image. Therefore, if we consider entropy as a measure of the disorder within a system, it could be used as a good stopping criterion for an iterative process, by using the mean shift iterative algorithm. More goodness on entropy can be seen in \cite{Rodriguez11a,Zhang03}. \section{THE MEAN SHIFT ITERATIVE ALGORITHM} \label{THE MEAN SHIFT ITERATIVE ALGORITHM} The Mean Shift Iterative Algorithm (MSHIA) with the old stopping criterion is composed of the following steps. Let $ent1$ be the initial value of the entropy of the first iteration. Let $ent2$ be the second value of the entropy after the first iteration. Let $errabs$ be the absolute value of the difference of entropy between the first and the second iteration. Let $edsEnt$ be the threshold to stop the iterations; that is, to stop when the relative rate of change of the entropy from one iteration to the next, falls below this threshold. Then, the segmentation algorithm comprises the following steps: \begin{algorithm}[h] \caption{Algorithm with the Old Stopping Criterion}\label{AlgoritmoOld} Initialize $ent2$, $errabs$ and $edsEnt$\; While $errabs > edsEnt$, then\; Filter the image according to the mean shift algorithm; store in $Z[k]$ the filtered image\; Calculate the entropy from the filtered image according to expression (\ref{entropy}); store in ent1\; Calculate the absolute difference with the entropy value obtained in the previous step; $errabs = \vert ent1 - ent2 \vert$\; Update the value of the parameters an image; $ent2 = ent1$; $Z[k +1] = Z[k]$\; \end{algorithm} \begin{algorithm}[h] \caption{The Algorithm with the New Stopping Criterion}\label{AlgoritmoNew} Initialize $edsEnt$ and $B1$, let $B1$ be equal to original image\; While $errabs > edsEnt$, then\; Filter the original image according to the mean shift algorithm; store in $B2$ the filtered image\; Calculate the absolute difference with the image obtained in the previous step; $C = \vert B2 - B1\vert$\; Calculate the entropy to the previous image; store in $errabs$\; Update the images: $B1 = B2$\; \end{algorithm} One can observe that with the new stopping criterion the entropy is used in another way. This paper will prove that with the new stopping criterion the MSHIA offers greater stability. Moreover, the following theorem resulted very interesting. \begin{theorem} \label{theorem1} When the entropy of the absolute difference between the iteration image and the following is taken as a stopping criterion in the mean shift iterative algorithm, whatever the chosen threshold this is enough condition to achieve the convergence. In addition, at the limit the entropy is zero. \end{theorem} This theorem ensures the convergence of the MSHIA with the new stopping criterion and determines what happens with entropy at the limit. The proof of this theorem can be found in the appendix. \section{EXPERIMENTAL RESULTS. ANALYSIS AND DISCUSSION} \label{EXPERIMENTAL RESULTS. ANALYSIS AND DISCUSSION} All segmentation experiments were carried out by using a uniform kernel. In order to be effective the comparison between the old stopping criterion and the new stopping criterion, the same value of $h_r$ and $h_s$ in the MSHIA were used. The principal goal of this section is to evaluate the new stop criterion in the MSHIA and to prove the stability of the algorithm with regard to the old stopping criterion. For this reason, comparisons with other segmentation approaches will not be carried out. In \cite{Rodriguez11a} were compared the obtained results with the MSHIA through the old stopping criterion with other segmentation methods. \begin{figure}[ht] \centering \subfigure[Astro]{\label{Figura1a}\includegraphics[width=2.733cm, height=2.75cm]{images/images/AstroOriginal.JPG} }\hspace{-0.1cm} \subfigure[Old criterion]{\label{Figura1b}\includegraphics[width=2.75cm, height=2.75cm]{images/images/AstroOld.JPG}} \subfigure[New criterion]{\label{Figura1c}\includegraphics[width=2.75cm, height=2.75cm]{images/images/AstroNew.JPG}} \subfigure[Barbara]{\label{Figura1d}\includegraphics[width=2.75cm, height=2.75cm]{images/images/BarbaraOriginal.JPG}} \subfigure[Old criterion]{\label{Figura1e}\includegraphics[width=2.75cm, height=2.75cm]{images/images/BarbaraOld.png}} \subfigure[New criterion]{\label{Figura1f}\includegraphics[width=2.75cm, height=2.75cm]{images/images/BarbaraNew.png}} \subfigure[Cameraman]{\label{Figura1g}\includegraphics[width=2.75cm, height=2.75cm]{images/images/CameramanOriginal.JPG}} \subfigure[Old criterion]{\label{Figura1h}\includegraphics[width=2.75cm, height=2.75cm]{images/images/CameramanOld.JPG}} \subfigure[New criterion]{\label{Figura1i}\includegraphics[width=2.75cm, height=2.75cm]{images/images/CameramanNew.JPG}} \subfigure[Lena]{\label{Figura1j}\includegraphics[width=2.75cm, height=2.75cm]{images/images/LenaOriginal.JPG}} \subfigure[Old criterion]{\label{Figura1k}\includegraphics[width=2.75cm, height=2.75cm]{images/images/LenaOld.JPG}} \subfigure[New criterion]{\label{Figura1l}\includegraphics[width=2.75cm, height=2.75cm]{images/images/LenaNew.JPG}} \caption{Used images for segmentation. In the first column are shown the original images; in the second, the segmentation using the old stopping criterion and the third column shows the segmented images using the new stopping criterion.} \label{fig:segmentation} \end{figure} Figure \ref{fig:segmentation} shows the segmentation of the used images. Observe that, in all cases, the iterative mean shift algorithm had better results when using the new stopping criterion. When one compares the segmented images with the old and new criterion it can be observed that with the new stopping criterion (see, for example, central images in Figure \ref{fig:segmentation}), more defined homogeneous zones were obtained. It can be seen that with the old stopping criterion the segmentation gave regions where different gray levels were originated. However, these regions really should have only one gray level. In other zones happened the opposite (see arrow in Figure \ref{Figura1f}). For example, Figure \ref{Figura1e} shows (visually) that the segmentation, with the old stopping criterion, is more diluted (see arrow). However, in Figure \ref{Figura1f} the segmentation with the new stopping criterion has the zones more delimited. One can find this same result when observing the other segmented images. This is one of the principal advantages of the new stopping criterion with regard to the old criterion. These good results are obtained because the defined new stopping criterion through the natural distance between images offers greater stability to the mean shift iterative algorithm. Figure \ref{fig:Profile} shows the profiles of the obtained segmented images by using the two stopping criteria. \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile2/AstroPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/AstroPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile2/AstroPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/AstroPerGrafNew.JPG}} \end{figure} \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile2/BarbaraPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile/BarbaraPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile2/BarbaraPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.2cm]{images/profile/BarbaraPerGrafNew.JPG}} \end{figure} \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile2/CameramanPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile/CameramanPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile2/CameramanPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.1cm, height=4.1cm]{images/profile/CameramanPerGrafNew.JPG}} \end{figure} \begin{figure}[ht] \centering \subfigure[Old criterion]{\includegraphics[width=4.2cm, height=4.4cm]{images/profile2/LenaPerImageOld.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/LenaPerGrafOld.JPG}} \subfigure[New criterion]{\includegraphics[width=4.2cm, height=4.4cm]{images/profile2/LenaPerImageNew.JPG}} \subfigure[Profile]{\includegraphics[width=4.2cm, height=4.3cm]{images/profile/LenaPerGrafNew.JPG}} \caption{Intensity profiles through the segmented images. Profiles are indicated by lines. First two columns are the segmented images with the old stopping criterion and its profile. The last columns are the segmented images with the new stopping criterion and its profile.} \label{fig:Profile} \end{figure} The plates that appear in Figures \ref{fig:Profile} of the second and the last columns are indicative of equal intensity levels. In both graphics the abrupt falls of intensities to others represent different regions in the segmented images. Note, for example, that in Figures where the segmented image appears with the new stopping criterion there are, in the same region of the segmentation, least variation of the pixel intensities with regard to the segmented images with the old stopping criterion. This illustrates that, in these cases the segmentation was better when the new stopping criterion was used. Figure \ref{Iteration graphs} shows three examples of the performance of the two stopping criterion in the experimental results. In the ``x'' axis appears the iterations of the MSHIA and in the ``y'' axis the obtained values by the stopping criterion in each iteration of the algorithm are shown. \begin{figure}[ht] \centering \subfigure[Barbara]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/BarbaraOld.JPG}} \subfigure[Cameraman]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/CameramanOld.JPG}} \subfigure[Lena]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/LenaOld.JPG}} \subfigure[Barbara]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/BarbaraNew.JPG}} \subfigure[Cameraman]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/CameramanNew.JPG}} \subfigure[Lena]{\includegraphics[width=2.72cm, height=2.72cm]{images/iterations/LenaNew.JPG}} \caption{Stopping criterion. In the first row appears the performance of the old stopping criterion, while in the second row the performance of the new stopping criterion is shown.} \label{Iteration graphs} \end{figure} The graphics of iterations of the new stopping criterion (second row) show a smooth behavior; that is, the stopping criterion has a better stable performance through the MSHIA. This is the goal of the presentation of these graphics, to observe the oscillations in the ``y'' axis (stability). The new stopping criterion not only has good theoretical properties (see theorem), but also, in practice, has very good behavior. When one analyzes the performance of the old stopping criterion in the experimental results (first row), one can see that the performance in the MSHIA is unstable. In general, this can originate unsuitable segmented images. \section{CONCLUSIONS} \label{CONCLUSIONS} In this work, a new stopping criterion for the MSHIA was proposed. The new stopping criterion is based on obtaining the entropy of the absolute difference between two images. In such sense, a theorem for the convergence was proposed. It was proven, through many experiments by using standard images, that the new stopping criterion had very good performance in the MSHIA. In addition, this was more stable that the old criterion. \section*{Acknowledgment} H. Sossa thanks SIP-IPN and CONACYT for the economical supports under grants 20131182 and 155014, respectively. We all thank the reviewers for their comments on the improvement of this paper.
\section{INTRODUCTION} \label{sec:intro} In reinforcement learning a tradeoff between exploration and exploitation of knowledge is considered. The multiarmed bandit problem is one formulation of the reinforcement learning and is a model of a gambler playing a slot machine with multiple arms. A dilemma for the gambler is that he cannot know whether the expectation of an arm is high or not without pulling it many times but he suffers a loss if he pulls suboptimal (i.e., not optimal) arms many times. This problem was formulated by \emcite{robbins} and its theoretical bound was derived by \emcite{lai} for single parametric models, which was extended to multiparameter models by \emcite{burnetas}. These theoretical bounds show that any suboptimal arm has to be pulled at least logarithmic number of rounds and its coefficient is determined by the distributions of suboptimal arms and the expectation of the optimal arm. Along with the asymptotic bound for this problem, achievability of the bound has also been considered in many models. \emcite{lai} proved the asymptotic optimality of a policy based on the notion of {\it upper confidence bound} (UCB) for Laplace distributions (which do not belong to exponential families) and some exponential families including normal distributions with {\it known} variances. The achievability of the bound was later extended to a subclass of one-parameter exponential families \cite{kl_ucb}. On the other hand in multiparameter or nonparametric models, \emcite{burnetas} and \emcite{honda_colt} proved the achievability for finite-support distributions and bounded-support distributions, respectively. However, the above two models are both compact and achievability of the bound is not known for non-compact multiparameter models, which include normal distributions with unknown means and variances. Since the normal distribution model is one of the most basic settings of stochastic bandits, many researches have been conducted for this model \cite{burnetas,ucb,bayes_ucb}. However, to the authors' knowledge, only the UCB-normal policy \cite{ucb} assures a (non-optimal) logarithmic regret for this model\footnote{% The theoretical analysis of UCB-normal contains conjectures verified only numerically and the logarithmic regret is not assured in the strict sense.% }. In this paper we discuss the asymptotic optimality of Thompson sampling (TS) \cite{thompson_original} for this normal distribution model with unknown means and variances. TS is a Bayesian policy which chooses an arm randomly according to the posterior probability with which the arm is the optimal. This policy was recently rediscovered and is researched extensively because of its excellent empirical performance for many models \cite{thompson_empirical}. The theoretical analysis of TS was first given for Bernoulli model \cite{thompson_log,thompson} and was later extended to general one-parameter exponential families \cite{thompson_exponential}. The asymptotic optimality of TS under uniform prior is proved for Bernoulli model in \emcite{thompson}, whereas it is proved for a more general model, one-parameter exponential family, under Jeffreys prior in \emcite{thompson_exponential}. Therefore, TSs with uniform prior and Jeffreys prior are asymptotically equivalent at least for the Bernoulli model. Nevertheless, we prove for the normal distribution model that TS with uniform prior achieves the asymptotic bound whereas TS with Jeffreys prior and reference prior cannot. Furthermore, TS with Jeffreys prior cannot even achieve a logarithmic regret and suffers a polynomial regret in expectation. This result implies that TS may be more sensitive to the choice of priors than expected and non-informative priors are sometimes risky (in other words, too optimistic) for multiparameter models. This paper is organized as follows. In Sect.\,\ref{sect_pre}, we formulate the bandit problem for the normal distribution model and introduce Thompson sampling. We give the main result on the optimality of TS in Sect.\,\ref{sect_main}. The remaining sections are devoted to the proof of the main result. In Sect.\,\ref{sect_prob}, we derive inequalities for probabilities which appear in the normal distribution model. We prove the optimality of TS with conservative priors in Sect.\,\ref{sect_regret} and prove the non-optimality of TS with optimistic priors in Sect.\,\ref{sect_dame}. \section{Preliminaries}\label{sect_pre} We consider the $K$-armed stochastic bandit problem in the normal distribution model. The gambler pulls an arm $i\in\{1,\cdots,K\}$ at each round and receives a reward independently and identically distributed by $\mathcal{N}(\mu_i,\sigma_i^2)$, where $\mathcal{N}(\mu,\sigma^2)$ denotes the normal distribution with mean $\mu$ and variance $\sigma^2$. The gambler does not know the parameter $(\mu_i,\sigma_i^2) \in \mathbb{R}\times (0,\infty)$. The maximum expectation is denoted by $\mu^*=\max_{i\in\{1,2,\cdots,K\}}\mu_i$. Let $J(t)$ be the arm pulled at the $t$-th round and $N_i(t)$ be the number of times that the arm $i$ is pulled before the $t$-th round. Then the regret of the gambler at the $T$-th round is given for $\Delta_i=\mu^*-\mu_i$ by \begin{align} \mathrm{Regret}(T)=\sum_{t=1}^T\Delta_{J(t)}=\sum_{i}\Delta_i N_i(T+1)\enspace .\nonumber \end{align} Let $X_{i,n}$ be the $n$-th reward from the arm $i$. We define \begin{align} \muhatt{i}{n}&= \frac1n\sum_{m=1}^n X_{i,m}\enspace ,\nonumber\\ \vhatt{i}{n}&=\sum_{m=1}^n (X_{i,m}-\muhatt{i}{n})^2= \sum_{m=1}^n X_{i,m}^2-n\muhatt{i}{n}^2\enspace ,\nonumber \end{align} that is, $\muhatt{i}{n}$ and $\vhatt{i}{n}$ denote the sample mean and the sum of squares from $n$ samples from the arm $i$, respectively. We denote the sample mean and the sum of squares before the $t$-th round by $\muhatn{i}{t}= \muhatt{i}{N_i(t)}$ and $\vhatn{i}{t}= \vhatt{i}{N_i(t)}$. It is well known that \begin{align} \muhatt{i}{n}&\sim \mathcal{N}(\mu_i, \sigma_i^2/n)\enspace , \qquad \frac{\vhatt{i}{n}}{\sigma_i^2}\sim\chi_{n-1}^2\enspace ,\label{normal_density} \end{align} where the chi-squared distribution $\chi_{n-1}^2$ with degree of freedom $n-1$ has the density \begin{align} \chi_{n-1}^2(s)= \frac{s^{\frac{n-3}{2}}\mathrm{e}^{-\frac{s}{2}}}{2^{\frac{n-1}{2}}\Gamma(\frac{n-1}{2})}\enspace .\nonumber \end{align} \subsection{Asymptotic Bound} It is shown in \emcite{burnetas} that under any policy satisfying a mild regularity condition the expected regret satisfies \begin{align} \lefteqn{ \!\!\! \liminf_{T\to\infty}\frac{\mathrm{E}[\mathrm{Regret}(T)]}{\log T} }\nonumber\\ &\ge \sum_{i: \Delta_i>0}\frac{\Delta_i}{ \inf_{(\mu,\sigma): \mu>\mu^*} D(\mathcal{N}(\mu_i,\sigma_i^2)\Vert\mathcal{N}(\mu,\sigma^2)) }\enspace ,\label{bound_burnetas} \end{align} where $D(\cdot\Vert \cdot)$ is the KL divergence. Since the KL divergence between normal distributions is \begin{align} \lefteqn{ D(\mathcal{N}(\mu_a,\sigma_b^2)\Vert \mathcal{N}(\mu_a,\sigma_b^2)) }\nonumber\\ &= \frac12\left( \log \frac{\sigma_b^2}{\sigma_a^2}+\frac{\sigma_a^2+(\mu_b-\mu_a)^2}{\sigma_b^2}-1\nonumber \right)\enspace ,\nonumber \end{align} the infimum in \eqref{bound_burnetas} is expressed for $\mu_i<\mu^*$ as \begin{align} \frac12\log\left(1+\frac{(\mu^*-\mu_i)^2}{\sigma_i^2}\right)\enspace .\nonumber \end{align} Therefore, by letting \begin{eqnarray} D_{\mathrm{inf}}(\Delta,\sigma^2) = \frac12\log\left(1+\frac{\Delta^2}{\sigma^2}\right)\enspace ,\nonumber \end{eqnarray} we can rewrite the theoretical bound in \eqref{bound_burnetas} as \begin{align} \liminf_{T\to\infty}\frac{\mathrm{E}[\mathrm{Regret}(T)]}{\log T} &\ge \sum_{i: \Delta_i>0}\frac{\Delta_i}{ D_{\mathrm{inf}}(\Delta_i,\sigma_i^2)}\enspace .\label{bound_normal} \end{align} \subsection{Bayesian Theory and Thompson Sampling}\label{sect_bayes} Thompson sampling is a policy based on the Bayesian viewpoint. We mainly consider the prior $\pi(\mu_i,\sigma_i^2)\sim (\sigma_i^2)^{-1-\alpha}$, or equivalently, $\pi(\mu_i,\sigma_i)\sim \sigma_i^{-1-2\alpha}$. Since the density of the inverse gamma distribution is \begin{align} \frac{\beta^{\alpha}}{\Gamma(\alpha)}x^{-1-\alpha}\mathrm{e}^{-\frac{\beta}{x}}\enspace ,\nonumber \end{align} the above prior for $\sigma_i^2$ corresponds to this distribution with parameters $(\alpha,\beta)=(\alpha, 0)$. The cases $\alpha=-1, -1/2,\allowbreak 0,1/2$ correspond to uniform for parameter $\sigma_i^2$, uniform for parameter $\sigma_i$, reference and Jeffreys priors, respectively (see, e.g.~\emcite{bayes_robert} for results on Bayesian theory given in this section). Under this prior, the posterior distribution is \begin{align} \pi(\mu_i|\samplet{i}{n}) &\sim \left(1+\frac{n(\mu_i-\muhatt{i}{n})^2}{\vhatt{i}{n}}\right)^{-\frac{n}{2}-\alpha}\enspace ,\nonumber \end{align} where $\samplet{i}{n}=(\muhatt{i}{n},\vhatt{i}{n})$. Since the density of $t$-distribution with degree of freedom $\nu$ is \begin{align} f_{\nu}(x)=\frac{\Gamma(\frac{\nu+1}{2})}{\sqrt{\nu\pi}\Gamma(\frac{\nu}{2})} \left(1+\frac{x^2}{\nu}\right)^{-\frac{\nu+1}{2}}\enspace ,\label{posterior1} \end{align} we see that \begin{align} \pi\left(\sqrt{\frac{n(n+2\alpha-1)}{\vhatt{i}{n}}}(\mu_i-\bar{x}) \Bigg| \samplet{i}{n}\right) = f_{n+2\alpha-1}\enspace .\label{posterior2} \end{align} Thompson sampling is the policy which chooses an arm randomly according to the probability with which the arm is the optimal when each $\mu_i$ is distributed independently by the posterior $\pi(\mu_i|\samplen{i}{t})$ for $\samplen{i}{t}=(\muhatn{i}{t}, \vhatn{i}{t})$. This policy is formulated as Algorithm \ref{alg_thompson}. Note that we require $\max\{2,2-\lceil 2\alpha \rceil\}$ initial pulls to avoid improper posteriors. We use $n_0=\max\{2,3-\lceil 2\alpha \rceil\}$ for simplicity of the analysis. \begin{algorithm}[t] \caption{Thompson Sampling}\label{alg_thompson} \begin{algorithmic} \sonomama{Parameter:} $\alpha\in\mathbb{R}$. \sonomama{Initialization:} \textnormal{Pull each arm $n_0=\max\{2,3-\lceil 2\alpha\rceil\}$ times}. \sonomama{Loop:} \mbox{}\\ \begin{enumerate} \item Sample $\tilde{\mu}_i(t)$ from the posterior $\pi_i(\mu_i|\samplen{i}{t})$ under prior $\pi(\mu_i,\sigma_i^2)\sim (\sigma_i^2)^{-1-\alpha}$ for each arm $i$. \item Pull an arm $i$ maximizing $\tilde{\mu}_i(t)$. \end{enumerate} \end{algorithmic} \end{algorithm}% \section{Regret of Thompson Sampling}\label{sect_main} In this section we give the main result of this paper. First we show that TS achieves the asymptotic bound for the prior $(\sigma_i^2)^{-1-\alpha}$ with $\alpha<0$. \begin{theorem}\label{thm_possible} Let $\epsilon>0$ be arbitrary and assume that there is a unique optimal arm. Under Thompson sampling with $\alpha<0$, the expected regret is bounded as \begin{align} \mathrm{E}[\mathrm{Regret}(T)] &\le \sum_{i:\Delta_i>0}\frac{\Delta_i\log T}{D_{\mathrm{inf}}(\Delta_i,\sigma_i^2)}+\mathrm{O}((\log T)^{4/5})\enspace .\nonumber \end{align} \end{theorem} See Lemma \ref{lem_possible} for the specific representation of the reminder term $\mathrm{O}((\log T)^{4/5})$. We see from this theorem that TS with $\alpha<0$ is asymptotically optimal in view of \eqref{bound_normal}. Next we show that TS with $\alpha\ge 0$ cannot achieve the asymptotic bound. To simplify the analysis we consider a two-armed setting more advantageous to the gambler in which the full information on the arm 2 is known beforehand, that is, the prior on the arm 2 is the unit point mass measure $\delta_{\{(\mu_2,\sigma_2^2)\}}$ instead of $\pi(\mu_2,\sigma_2^2)\sim (\sigma_2^2)^{-1-\alpha}$. \begin{theorem}\label{thm_impossible} Assume that there are $K=2$ arms such that $\mu_1>\mu_2$. Then, under Thompson sampling such that $\tilde{\mu}_1(t)\sim \pi(\mu_1 |\samplen{1}{t})$ with $\alpha\ge 0$ and $\tilde{\mu}_2(t)=\mu_2$, there exists a constant $\xi>0$ independent of $\sigma_2$ such that \begin{align} \liminf_{T\to\infty}\frac{\mathrm{E}[\mathrm{Regret}(T)]}{\log T}\ge \xi\enspace .\label{cannot1} \end{align} In particular, if $\alpha>0$ then there exist $\xi'>0$ and $\eta>0$ such that \begin{align} \liminf_{N\to\infty}\frac{\mathrm{E}[\mathrm{Regret}(T)]}{T^{\eta}}\ge \xi'\enspace .\label{cannot2} \end{align} \end{theorem} Eq.\,\eqref{cannot2} means that TS with $\alpha>0$ suffers a polynomial regret in expectation. Also note that the asymptotic bound in \eqref{bound_normal} approaches zero for sufficiently small $\sigma_2$ in the above two-armed setting since $D_{\mathrm{inf}}(\Delta_i, \sigma_i^2)\to \infty$ as $\sigma_i\to 0$. Nevertheless, the LHS of \eqref{cannot1} does not go to zero as $\sigma_2\to 0$ because $\xi>0$ is independent of $\sigma_2$. Therefore TS with $\alpha= 0$ also does not achieve the asymptotic bound at least for sufficiently small $\sigma_2$. Recall that Jeffreys and reference priors correspond to $\alpha=1/2$ and $\alpha=0$, respectively. Therefore this theorem means that TS with these non-informative priors does not achieve the asymptotic bound. \begin{remark}{\rm Probability that the sample mean satisfies $\bar{x}_i<\mu$ for any $\mu<\mu_i$ becomes large when $\sigma_i^2$ is large. Therefore the posterior probability that the true expectation $\mu_i$ is larger than $\mu>\bar{x}_i$ becomes large when the prior has heavy weight at large $\sigma_i^2$, that is, $\alpha$ is small. As a result, as $\alpha$ decreases, TS becomes a ``conservative'' policy which chooses a seemingly suboptimal arm frequently. Theorems \ref{thm_possible} and \ref{thm_impossible} mean that the prior should be conservative to some extent and non-informative priors are too optimistic. }\end{remark} \begin{remark}{\rm Although TS with non-informative priors does not achieve the asymptotic bound in the sense of expectation, this fact does not necessarily mean that these priors are ``bad'' ones. As we can see from a close inspection of the proof of Theorem \ref{thm_impossible}, the expected regret of TS with these priors becomes large because an enormously large regret arises with fairly small probability. Therefore this policy performs well except for the case arising with this small probability, and the authors think that TS with these priors also becomes a good policy in the probably approximately correct (PAC) framework. In any case, we should be aware that these non-informative priors are ``risky'' in the sense of expectation. }\end{remark} \section{Inequalities for normal distributions and t-distributions}\label{sect_prob} In this section we derive fundamental inequalities for distributions appearing in Thompson sampling for the normal distribution model. We prove them in Appendix. First we give a simple inequality to evaluate the ratio of gamma functions which appears in the densities of normal, chi-squared and $t$-distributions. \begin{lemma}\label{lem_gamma} For $z\ge 1/2$ \begin{align} \mathrm{e}^{-2/3}\le \frac{\Gamma(z+\frac12)}{\Gamma(z)}\le \mathrm{e}^{1/6}\sqrt{z}\enspace .\nonumber \end{align} \end{lemma} Next we give large deviation probabilities (see, e.g., \emcite{LDP}) for empirical means and variances. \begin{lemma}\label{lem_ldp} For any $\mu>\mu_i$ \begin{align} \Pr[\muhatt{i}{n}\ge \mu]\le \mathrm{e}^{-\frac{n(\mu-\mu_i)^2}{2\sigma_i^2}}\label{ldp_mean} \end{align} and for any $\sigma^2>\sigma_i^2$ \begin{align} \Pr[\vhatt{i}{n}\ge n\sigma^2]\le \mathrm{e}^{-nh\left(\frac{\sigma^2}{\sigma_i^2}\right)}\label{ldp_var} \end{align} where $h(x)=(x-1-\log x)/2\ge 0$. \end{lemma} \begin{remark}{\rm It is well known that Mill's ratio \cite[Chap.\,5]{kendall} gives a tighter bound for the tail probability of normal distributions, and similar technique can also be applied to the tail weight of $\chi^2$ distributions. However, we use bounds in Lemma \ref{lem_ldp} based on the large deviation principle because they are simpler and convenient for our analysis. }\end{remark} Finally we evaluate the posterior distribution of the mean for Thompson sampling. Probability that the sample from the posterior is larger than or equal to $\mu$, which is formally defined as \begin{align} \pti{i}{n}{\mu} =\int_{\mu}^{\infty}\pi(x|\samplet{i}{n})\mathrm{d} x\enspace ,\nonumber \end{align} is bounded as follows. \begin{lemma}\label{lem_upper} If $\mu>\muhatt{i}{n}$ and $n\ge n_0$ then \begin{align} \pti{i}{n}{\mu} &\ge A_{n,\alpha}\left(1+\frac{n(\mu-\muhatt{i}{n})^2}{\vhatt{i}{n}}\right)^{-\frac{n-1}{2}-\alpha}\label{eq_lower} \end{align} and \begin{align} \pti{i}{n}{\mu} &\le \frac{\sqrt{\vhatt{i}{n}}}{\mu-\muhatt{i}{n}} \left(1+\frac{n(\mu-\muhatt{i}{n})^2}{\vhatt{i}{n}}\right)^{-\frac{n}{2}-\alpha+1}\enspace ,\label{eq_upper} \end{align} where \begin{align} A_{n,\alpha}&=\frac{1}{2\mathrm{e}^{1/6}\sqrt{\pi(\frac{n}{2}+\alpha)}}\enspace .\label{eq_an} \end{align} \end{lemma} \section{Analysis for Conservative Priors}\label{sect_regret} In this section we show that Thompson sampling achieves the asymptotic bound if $\alpha<0$. The main result of this section is given as follows. \begin{lemma}\label{lem_possible} Fix any $\alpha<0$ and assume that $(\mu_1,\sigma_1^2)=(0,1)$ and the arm 1 is the unique optimal arm. Then, for any $\epsilon<\min_{i:\Delta_i>0}\Delta_i/2$, \begin{align} \lefteqn{ \mathrm{E}[\mathrm{Regret}(T)] }\nonumber\\ &\le \sum_{i: \Delta_i>0}\Delta_i \Bigg( \frac{\log T}{D_{\mathrm{inf}}(\Delta_i-2\epsilon,\sigma_i^2+\epsilon)}+2-2\alpha \nonumber\\ &\qquad\qquad+\frac{\sqrt{\sigma_i^2+\epsilon}}{\Delta_i-2\epsilon} +\wa{\frac{\epsilon}{2\sigma_i^2}}+\wa{h(1+\frac{\epsilon}{\sigma_i^2})} \Bigg) \nonumber\\ &\quad+\Delta_{\max}\Bigg( \wa{\frac{\epsilon^2}{2}}+\wa{h(2)} +\frac{\mathrm{B}(1/2,-\alpha)}{(1-\mathrm{e}^{-\frac{\epsilon^2}{2}})^2}\nonumber\\ &\phantom{wwwwwwwwi}+ \frac{2\sqrt{2}}{\epsilon} \frac{\left(1+\epsilon^2/8\right)^{1-\alpha}}{1-\left(1+\epsilon^2/8\right)^{-1/2}} \Bigg) \nonumber\\ &= \sum_{i: \Delta_i>0} \frac{\Delta_i\log T}{D_{\mathrm{inf}}(\Delta_i-\epsilon,\sigma_i^2+\epsilon)} +\mathrm{O}(\epsilon^{-4})\enspace ,\nonumber \end{align} where $\Delta_{\max}=\max_i \Delta_i$ and $\mathrm{B}(\cdot,\cdot)$ is the beta function. \end{lemma} \begin{corollary}\label{cor_possible} Under the same assumption as Lemma \ref{lem_possible}, \begin{align} \mathrm{E}[\mathrm{Regret}(T)]\le\sum_{i: \Delta_i>0} \frac{\Delta_i\log T}{D_{\mathrm{inf}}(\Delta_i,\sigma_i^2)} +\mathrm{O}((\log T)^{4/5})\enspace .\nonumber \end{align} \end{corollary} This corollary is straightforward from Lemma \ref{lem_possible} with $\epsilon:=\mathrm{O}((\log T)^{-1/5})$. Note that $D_{\mathrm{inf}}(\mu^*-\mu_i, \sigma_2^2)$ is invariant under the location and scale transformation, that is, \begin{align} D_{\mathrm{inf}}(\mu^*-\mu_i, \sigma_i^2)=D_{\mathrm{inf}}\left(\frac{\mu^*-a}{b}-\frac{\mu_i-a}{b}, \frac{\sigma_i^2}{b^2} \right).\nonumber \end{align} Thus Theorem \ref{thm_possible} easily follows from Corollary \ref{cor_possible} by the transformation $((\mu_1,\sigma_1^2),(\mu_2,\sigma_2^2),\cdots,(\mu_K,\sigma_K^2))\mapsto ((0,1),\allowbreak ((\mu_2-\mu_1)/\sigma_1,\sigma_2^2/\sigma_1^2),\cdots, ((\mu_K-\mu_1)/{\sigma_1},\allowbreak \sigma_K^2/\sigma_1^2))$. \begin{proof}[Proof of Lemma \ref{lem_possible}] Define events \begin{align} \mathcal{A}(t)&=\{\tilde{\mu}^*(t)\ge -\epsilon\}\enspace ,\nonumber\\ \mathcal{B}_i(t)&=\{\muhatn{i}{t}\le \mu_i+\delta,\, \vhatn{i}{t}\le n(\sigma_i^2+\epsilon)\}\enspace ,\nonumber \end{align} where $\tilde{\mu}^*(t)=\max_{i}\tilde{\mu}_i(t)$. Then the regret at the round $T$ is bounded as \begin{align} \lefteqn{ \!\!\! \mathrm{Regret}(T) }\nonumber\\ &= \sum_{t=1}^T \Delta_{J(t)}\nonumber\\ &\le \Delta_{\max}\sum_{t=Kn_0+1}^T \idx{J(t)\neq 1,\; \mathcal{A}^c(t)}\nonumber\\ &\quad+\sum_{i=2}^K \Delta_i \left(n_0+\sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{A}(t)}\right)\nonumber\\ &\le \Delta_{\max}\sum_{t=Kn_0+1}^T \idx{J(t)\neq 1,\; \mathcal{A}^c(t)}\nonumber\\ &\quad+ \sum_{i=2}^K\Delta_i\Bigg( \sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{A}(t),\; \mathcal{B}_i(t)}\nonumber\\ &\qquad\qquad+\sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{B}_i^c(l)}+n_0\Bigg) \enspace ,\label{bunkai} \end{align} where $\idx{\cdot}$ is the indicator function and the superscript ``$c$'' denotes the complementary set. In the following Lemmas \ref{lem_difficult}--\ref{lem_easy} we bound the expectation of the above three terms and the proof is completed. \end{proof} \begin{lemma}\label{lem_difficult} If $\alpha<0$ then \begin{align} \lefteqn{ \mathrm{E}\left[\sum_{t=Kn_0+1}^T \idx{J(t)\neq 1 \cup \mathcal{A}^c(t)}\right] }\nonumber\\ &\le \frac{1}{1-\mathrm{e}^{-\frac{\epsilon^2}{8}}} +\frac{1}{1-\mathrm{e}^{-h(2)}} +\frac{2\sqrt{2}}{\epsilon} \frac{\left(1+\frac{\epsilon^2}{8}\right)^{1-\alpha}}{1-\left(1+\frac{\epsilon^2}{8}\right)^{-1/2}}\nonumber\\ &\quad+ \frac{\mathrm{B}(1/2,-\alpha)}{\left(1-\mathrm{e}^{-\frac{\epsilon^2}{2}}\right)^2}\nonumber\\ &=\mathrm{O}(\epsilon^{-4})\enspace .\nonumber \end{align} \end{lemma} \begin{lemma}\label{lem_main} For any $i\neq 1$, \begin{align} \lefteqn{ \mathrm{E}\left[\sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{A}(t),\; \mathcal{B}_i(t)}\right] }\nonumber\\ &\le \frac{\log N}{D_{\mathrm{inf}}(\Delta_i-2\epsilon,\sigma_i^2+\epsilon)} +2-2\alpha +\frac{\sqrt{\sigma_i^2+\epsilon}}{\Delta_i-2\epsilon}\nonumber\\ &=\frac{\log N}{D_{\mathrm{inf}}(\Delta_i-2\epsilon,\sigma_i^2+\epsilon)}+\mathrm{O}(1) \enspace .\nonumber \end{align} \end{lemma} \begin{lemma}\label{lem_easy} For any $i\neq 1$, \begin{align} \lefteqn{ \!\!\!\!\! \mathrm{E}\left[\sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{B}_i^c(t)}\right] }\nonumber\\ &\le \frac{1}{1-\mathrm{e}^{-\frac{\epsilon^2}{2\sigma_i^2}}} +\frac{1}{1-\mathrm{e}^{-h\left(1+\frac{\epsilon}{\sigma_i^2}\right)}} =\mathrm{O}(\epsilon^{-2})\enspace .\nonumber \end{align} \end{lemma} We prove Lemma \ref{lem_easy} in Appendix and prove Lemmas \ref{lem_difficult} and \ref{lem_main} in this section. Whereas the second term of \eqref{bunkai} becomes the main term of the regret, the evaluation of the first term is the most difficult point of the proof, which corresponds to Lemma \ref{lem_difficult}. In fact, it is reported in \emcite{burnetas} that they were not able to prove the asymptotic optimality of a policy for the normal distribution model because of difficulty of the evaluation corresponding to this term. Also note that this is the term which does not become a constant in the case $\alpha\ge 0$ and is considered in the proof of Theorem \ref{thm_impossible}. In this paper we evaluate this term by first bounding this term for a fixed statistic $\samplet{i}{n}=(\muhatt{i}{n},\vhatt{i}{n})$ and finally taking its expectation, whereas a probability on this statistic is first evaluated in \emcite{burnetas}. By leaving the evaluation on the distribution of $\samplet{i}{n}$ to the latter part, we can significantly simplify the integral by variable transformation. \begin{proof2}{of Lemma \ref{lem_difficult}} First we bound the summation as \begin{align} \lefteqn{ \sum_{t=Kn_0+1}^T \idx{J(t)\neq 1 ,\; \mathcal{A}^c(t)} }\nonumber\\ &= \sum_{n=n_0}^{T}\sum_{t=Kn_0+1}^{T} \idx{J(t)\neq 1,\; \mathcal{A}^c(t),\; N_1(t)=n}\nonumber\\ &= \sum_{n=n_0}^{T}\sum_{m=1}^{T}\nonumber\\ &\qquad\idx{m\le \sum_{t=Kn_0+1}^{T} \idx{J(t)\neq 1,\; \mathcal{A}^c(t),\; N_1(t)=n}}.\nonumber \end{align} Note that \begin{align} m\le \sum_{t=Kn_0+1}^{T} \idx{J(t)\neq 1,\; \mathcal{A}^c(t),\; N_1(t)=n}\nonumber \end{align} implies that $\tilde{\mu}_1(t)\le \tilde{\mu}^*(t)\le -\epsilon$ occurred for the first $m$ elements of $\{t: \mathcal{A}^c(t),\; N_1(t)=n\}$. Therefore, \begin{align} \lefteqn{ \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \Pr\left[ m\le \sum_{t=Kn_0+1}^{T} \idx{J(t)\neq 1,\; \mathcal{A}^c(t),\; N_1(t)=n} \right] }\nonumber\\ &\le (1-\pti{1}{n}{-\epsilon})^m\nonumber \end{align} and we have \begin{align} \lefteqn{ \mathrm{E}\left[\sum_{t=Kn_0+1}^T \idx{J(t)\neq 1 \cup \mathcal{A}^c(t)}\right] }\nonumber\\ &\le \mathrm{E}\left[ \sum_{n=n_0}^{T}\sum_{m=1}^{T} (1-\pti{1}{n}{-\epsilon})^m \right]\nonumber\\ &\le \sum_{n=n_0}^{T} \mathrm{E}\left[\frac{1-\pti{1}{n}{-\epsilon}}{\pti{1}{n}{-\epsilon}} \right]\enspace .\label{matizikan} \end{align} Since $\pti{i}{n}{-\epsilon}\ge 1/2$ for $-\epsilon\le \muhatt{i}{n}$ from the symmetry of $t$-distribution, this expectation is partitioned into \begin{align} \lefteqn{ \mathrm{E}\left[\frac{1-\pti{1}{n}{-\epsilon}}{\pti{1}{n}{-\epsilon}}\right] }\nonumber\\ &\le 2\mathrm{E}\left[\idx{-\epsilon\le\muhatt{1}{n}}(1-\pti{1}{n}{-\epsilon})\right]\nonumber\\ &\quad+ \mathrm{E}\left[\frac{\idx{\muhatt{1}{n}\le -\epsilon}}{\pti{1}{n}{-\epsilon}}\right] \nonumber\\ &\le \Pr\left[-\epsilon<\muhatt{1}{n}\le -\epsilon/2\right] \nonumber\\&\quad +\Pr\left[-\epsilon/2<\muhatt{1}{n},\; \vhatt{1}{n}\ge 2n\right]\nonumber\\ &\quad+2\mathrm{E}\big[\idx{-\epsilon/2<\muhatt{i}{n},\; \vhatt{1}{n}\le 2n} (1-\pti{i}{n}{-\epsilon})\big]\nonumber\\ &\quad +\mathrm{E}\left[\frac{\idx{\muhatt{1}{n}\le -\epsilon}}{\pti{1}{n}{-\epsilon}}\right] . \label{partition4} \end{align} From Lemma \ref{lem_ldp}, the first and second terms of \eqref{partition4} are bounded as \begin{align} \Pr\left[-\epsilon<\muhatt{i}{n}\le -\epsilon/2\right] &\le \mathrm{e}^{-\frac{n\epsilon^2}{8}}\enspace ,\nonumber\\ \Pr\left[-\epsilon/2<\muhatt{1}{n},\; \vhatt{1}{n}\ge 2n\right] &\le\mathrm{e}^{-nh(2)}\enspace ,\label{term3} \end{align} respectively. Next, recall that $\samplet{1}{n}=(\muhatt{i}{n}, \vhatt{i}{n})$. Then, from the symmetry of $t$-distribution \begin{align} 1-p_n(-\epsilon|\muhatt{i}{n}, \vhatt{i}{n}) &=1-p_n(-\muhatt{1}{n}-\epsilon|0,\vhatt{1}{n})\nonumber\\ &=p_n(\muhatt{1}{n}+\epsilon|0,\vhatt{1}{n})\nonumber\\ &=p_n(2\muhatt{1}{n}+\epsilon|\muhatt{1}{n},\vhatt{1}{n})\nonumber \end{align} and the third term of \eqref{partition4} is bounded from \eqref{eq_upper} as \begin{align} \lefteqn{ \mathrm{E}\big[\idx{-\epsilon/2<\muhatt{i}{n},\; \vhatt{1}{n}\le 2} (1-\pti{i}{n}{-\epsilon})\big] }\nonumber\\ &= \mathrm{E}\big[\idx{-\epsilon/2<\muhatt{i}{n},\; \vhatt{1}{n}\le 2} \pti{i}{n}{2\muhatt{i}{n}-\epsilon}\big]\nonumber\\ &\le \frac{2\sqrt{2}}{\epsilon} \left(1+\frac{\epsilon^2}{8}\right)^{-\frac{n}{2}-\alpha+1}\enspace .\label{term4} \end{align} Finally we evaluate the fourth term of \eqref{partition4}. From \eqref{normal_density} and \eqref{eq_lower}, we have \begin{align} \lefteqn{ \mathrm{E}\left[\frac{\idx{\muhatt{1}{n}\le -\epsilon}}{\pti{1}{n}{-\epsilon}}\right] }\nonumber\\ &\le \frac1{A_{n,\alpha}}\int_{-\infty}^{-\epsilon}\int_{0}^{\infty} \left(1+\frac{n(x+\epsilon)^2}{s}\right)^{\frac{n-1}{2}+\alpha}\nonumber\\ &\qquad\cdot \sqrt{\frac{n}{2\pi}}\mathrm{e}^{-\frac{nx^2}{2}} \frac{s^{\frac{n-3}{2}} \mathrm{e}^{-\frac{s}{2}}}{2^{\frac{n-1}{2}}\Gamma(\frac{n-1}{2})} \mathrm{d} s\mathrm{d} x\nonumber\\ &\le \frac{\mathrm{e}^{-n\epsilon^2/2}}{A_{n,\alpha}}\int_{-\infty}^{-\epsilon}\int_{0}^{\infty} \left(1+\frac{n(x+\epsilon)^2}{s}\right)^{\frac{n-1}{2}+\alpha}\nonumber\\ &\qquad\cdot \sqrt{\frac{n}{2\pi}}\mathrm{e}^{-\frac{n(x+\epsilon)^2}{2}} \frac{s^{\frac{n-3}{2}} \mathrm{e}^{-\frac{s}{2}}}{2^{\frac{n-1}{2}}\Gamma(\frac{n-1}{2})} \mathrm{d} s\mathrm{d} x\label{main_sekibun \end{align} by $x^2\ge (x+\epsilon)^2+\epsilon^2$ for $x\le -\epsilon\le 0$. By letting \begin{align} (x,s)=\left( -\epsilon-\sqrt{\frac{2zw}{n}},\, 2 z (1-w)\right)\enspace ,\nonumber \end{align} we have \begin{align} \mathrm{d} x \mathrm{d} s&= \det\left( \begin{array}{cc} -\sqrt{\frac{w}{2nz}}&2(1-w)\\ -\sqrt{\frac{z}{2nw}}&-2z \end{array} \right)\mathrm{d} z \mathrm{d} w\nonumber\\ &= \sqrt{\frac{2z}{nw}} \mathrm{d} z \mathrm{d} w\nonumber \end{align} and \eqref{main_sekibun} is rewritten as \begin{align} \lefteqn{ \mathrm{E}\left[\frac{\idx{\muhatt{i}{n}\le -\epsilon}}{\pti{i}{n}{-\epsilon}}\right] }\nonumber\\ &\le \frac{\mathrm{e}^{-n\epsilon^2/2}}{A_{n,\alpha}}\int_{0}^{1}\int_{0}^{\infty} \left(1+\frac{w}{1-w}\right)^{\frac{n-1}{2}+\alpha}\nonumber\\ &\qquad\cdot \sqrt{\frac{n}{2\pi}}\mathrm{e}^{-zw} \frac{(z(1-w))^{\frac{n-3}{2}} \mathrm{e}^{-z(1-w)}}{2\Gamma(\frac{n-1}{2})} \sqrt{\frac{2z}{nw}}\mathrm{d} z \mathrm{d} w\nonumber\\ &= \frac{\mathrm{e}^{-n\epsilon^2/2}}{2\sqrt{\pi}A_{n,\alpha}\Gamma(\frac{n-1}{2})}\int_{0}^{1} w^{-1/2}(1-w)^{-1-\alpha}\mathrm{d} w \nonumber\\ &\qquad\cdot \int_{0}^{\infty} \mathrm{e}^{-z} z^{\frac{n}{2}-1}\mathrm{d} z \nonumber\\ &= \frac{\mathrm{e}^{-n\epsilon^2/2}}{2\sqrt{\pi}A_{n,\alpha}\Gamma(\frac{n-1}{2})} \mathrm{B}(1/2, -\alpha) \Gamma(n/2)\nonumber\\ &\le \frac{\mathrm{e}^{1/3}\sqrt{(n+2\alpha)(n-1)}}{2} \mathrm{e}^{-n\epsilon^2/2}\mathrm{B}(1/2, -\alpha) \nonumber\\ &\phantom{wwwwwwwwwwwwwwww}(\mbox{by Lemma \ref{lem_gamma} and \eqref{eq_an}})\nonumber\\ &\le n\mathrm{e}^{-n\epsilon^2/2}\mathrm{B}(1/2, -\alpha)\enspace .\;\,\quad \mbox{(by $\mathrm{e}^{1/3}\le 2$ and $\alpha<0$)} \label{term1} \end{align} By combining \eqref{matizikan}, \eqref{partition4}, \eqref{term3}, \eqref{term4} with \eqref{term1}, we obtain \begin{align} \lefteqn{ \mathrm{E}\left[\sum_{t=Kn_0+1}^T \idx{J(t)\neq 1 ,\; \mathcal{A}^c(t)}\right] }\nonumber\\ &\le \sum_{n=n_0}^T\Bigg( \mathrm{e}^{-\frac{n\epsilon^2}{8}}+\mathrm{e}^{-nh(2)}+ \frac{2\sqrt{2}}{\epsilon} \left(1+\frac{\epsilon^2}{8}\right)^{-\frac{n}{2}-\alpha+1}\nonumber\\ &\phantom{wwwwwwwwwwwwwww}+ n\mathrm{e}^{-n\epsilon^2/2}\mathrm{B}(1/2, -\alpha)\Bigg)\nonumber\\ &\le \frac{1}{1-\mathrm{e}^{-\frac{\epsilon^2}{8}}} +\frac{1}{1-\mathrm{e}^{-h(2)}} +\frac{2\sqrt{2}}{\epsilon} \frac{\left(1+\frac{\epsilon^2}{8}\right)^{1-\alpha}}{1-\left(1+\frac{\epsilon^2}{8}\right)^{-1/2}}\nonumber\\ &\quad+ \frac{\mathrm{B}(1/2,-\alpha)}{\left(1-\mathrm{e}^{-\frac{\epsilon^2}{2}}\right)^2}\nonumber\\ &=\mathrm{O}(\epsilon^{-2})+\mathrm{O}(1)+\mathrm{O}(\epsilon^{-3})+\mathrm{O}(\epsilon^{-4})=\mathrm{O}(\epsilon^{-4}) \enspace .\!\!\relax\tag*{\qed}\nonumber \end{align} \end{proof2} \begin{proof}[Proof of Lemma \ref{lem_main}] Let $n_i>0$ be arbitrary. Then \begin{align} \lefteqn{ \sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{A}(t),\; \mathcal{B}_i(t)} }\nonumber\\ &\le \sum_{t=Kn_0+1}^T \idx{\tilde{\mu}_i(t)\ge -\epsilon,\; B_i(l)}\nonumber\\ &\le n_i+ \sum_{t=Kn_0+1}^{T} \idx{\tilde{\mu}_i(t)\ge -\epsilon,\; \mathcal{B}_i(t),\,N_i(t)\ge n_i}\enspace .\label{lognterm} \end{align} Under the condition $\{\mathcal{B}_i(t),\; N_i(t)=n\}$, the probability of the event $\tilde{\mu}_i(t)\ge -\epsilon=\mu^*-\epsilon$ is bounded from Lemma \ref{lem_upper} as \begin{align} \pti{i}{n}{-\epsilon} &\le \frac{\sqrt{\sigma_i^2+\epsilon}}{\Delta_i-2\epsilon} \left(1+\frac{(\Delta_i-2\epsilon)^2}{\sigma_i^2+\epsilon}\right)^{-\frac{n}{2}-\alpha+1}\nonumber\\ &= \frac{\sqrt{\sigma_i^2+\epsilon}}{\Delta_i-2\epsilon} \mathrm{e}^{-(n+2\alpha-2)D_{\mathrm{inf}}(\Delta_i-2\epsilon, \sigma_i^2+\epsilon)}.\nonumber \end{align} Therefore the expectation of \eqref{lognterm} is bounded as \begin{align} \lefteqn{ \mathrm{E}\left[ \sum_{t=Kn_0+1}^T\idx{J(t)=i,\; \mathcal{A}(t),\; \mathcal{B}_i(t)} \right] }\nonumber\\ &\le n_i+ \sum_{t=Kn_0+1}^{T} \Pr\left[\tilde{\mu}_i(t)\ge -\epsilon,\; \mathcal{B}_i(t),\,N_i(t)\ge n_i\right]\nonumber\\ &\le n_i+ T\frac{\sqrt{\sigma_i^2+\epsilon}}{\Delta_i-2\epsilon} \mathrm{e}^{-(n_i+2\alpha-2)D_{\mathrm{inf}}(\Delta_i-2\epsilon, \sigma_i^2+\epsilon)}\nonumber \end{align} and we complete the proof by letting $n_i=(\log T)/\allowbreak D_{\mathrm{inf}}(\Delta_i-2\epsilon,\sigma_i^2+\epsilon)+2-2\alpha$. \end{proof} \section{Analysis for Optimistic Priors}\label{sect_dame} In this section we prove Theorem \ref{thm_impossible}. As mentioned before, the evaluation in the proof corresponds to Lemma \ref{lem_difficult}, in which $\alpha<0$ is required so that $\mathrm{B}(1/2,-\alpha)$ becomes finite. We show in the following proof that this requirement is actually necessary to achieve the asymptotic bound. \begin{proof}[Proof of Theorem \ref{thm_impossible}] We assume $(\mu_1,\sigma_1^2)=(0,1)$ without loss of generality. Fix any $n\ge n_0$ and let $T_{1,n}\in\mathbb{N}\cup \{+\infty\}$ be the first round at which the number of samples from the arm 1 is $n$, that is, we define $T_{1,n}=\min\{t:N_1(t)=n\}$. Since $T_{1,n}=t$ implies that the arm $2$ is pulled $t-n-1$ times through the first $t-1$ rounds, we have \begin{align} \lefteqn{ \mathrm{E}[\mathrm{Regret}(T)] }\nonumber\\ &= \Delta_i\sum_{t=1}^{\infty}\Pr[T_{1,n}=t]\mathrm{E}\left[ \sum_{m=1}^{2T}\idx{J(m)=2} \Bigg|T_{1,n}=t\right]\nonumber\\ &\ge\Delta_i\sum_{t=1}^{T}\Pr[T_{1,n}=t] \mathrm{E}\left[\sum_{m=1}^{2T}\idx{J(m)=2}\Bigg|T_{1,n}=t\right]\nonumber\\ &\quad+\Delta_i\sum_{t=T+1}^{\infty}\Pr[T_{1,n}=t](T-n)\enspace .\label{bunkai_impo} \end{align} Now we consider the case $t\le T$. The conditional expectation in \eqref{bunkai_impo} is bounded as \begin{align} \lefteqn{ \mathrm{E}\left[\sum_{m=1}^{2T} \idx{J(m)=2}\Bigg|T_{1,n}=t\right] }\nonumber\\ &\ge\mathrm{E}\left[\sum_{m=t}^{T+t-1} \idx{J(m)=2,\; N_1(m)=n}\Bigg|T_{1,n}=t\right]\nonumber \end{align} Note that if $\{J(m)\neq 2,\; N_1(m)=n\}$ then $N_1(m')>n$ for any $m'>m$. Therefore, for any $m\ge T_{1,n}$, \begin{align} \{J(m)=2,\, N_1(m)=n\} &\Leftrightarrow\! \bigcup_{k=0}^{m-T_{1,n}}\{J(T_{1,n}+k)=2\}\nonumber\\ &\Leftarrow\! \bigcup_{k=0}^{m-T_{1,n}}\{\tilde{\mu}_1(T_{1,n}+k)<\mu_2\}\nonumber \end{align} since $\tilde{\mu}_2(t)=\mu_2$ always holds. By defining $\ptic{1}{n}{\mu_2}\allowbreak =1-\pti{1}{n}{\mu_2}$ we have \begin{align} \lefteqn{ \mathrm{E}\left[\sum_{m=1}^{T+t-1} \idx{J(m)=2}\Bigg|T_{1,n}=t\right] }\nonumber\\ &\ge\mathrm{E}\left[\sum_{m=t}^{T+t-1} \idx{\bigcup_{k=0}^{m-t} \{\tilde{\mu}_1(t+k)<\mu_2\}}\Bigg|T_{1,n}=t\right]\nonumber\\ &=\mathrm{E}\left[\sum_{m=t}^{T+t-1}(\ptic{1}{n}{\mu_2})^{m-t+1}\right]\nonumber\\ &\ge\mathrm{E}\left[\sum_{m=1}^{T}(\ptic{1}{n}{\mu_2})^{m}\right]\nonumber\\ &=\mathrm{E}\left[ \left(1-(\ptic{1}{n}{\mu_2})^{T}\right)\frac{\ptic{1}{n}{\mu_2}}{\pti{1}{n}{\mu_2}} \right]\nonumber\\ &\ge\frac12\mathrm{E}\left[ \idx{(\ptic{1}{n}{\mu_2})^T\le 1/2} \frac{\ptic{1}{n}{\mu_2}}{\pti{1}{n}{\mu_2}} \right]\nonumber\\ &\ge\frac12\mathrm{E}\left[ \frac{ \idx{(\ptic{1}{n}{\mu_2})^T\le 1/2} }{\pti{1}{n}{\mu_2}} \right]-\frac12\enspace .\label{sita_id}\\ &\phantom{wwwwwwwwwwwwww}(\mbox{by $(1-p)/p= 1/p-1$})\nonumber \end{align} Here we obtain from \eqref{eq_lower} that \begin{align} \lefteqn{ (\ptic{1}{n}{\mu_2})^T\le 1/2 }\nonumber\\ &\Leftrightarrow \pti{1}{n}{\mu_2}\ge 1-2^{-\frac{1}{T}}\nonumber\\ &\Leftarrow \left(1+\frac{n(\mu_2-\muhatt{1}{n})^2}{\vhatt{1}{n}}\right)^{-\frac{n-1}{2}-\alpha} \ge \frac{1-2^{-\frac{1}{T}}}{A_{n,\alpha}}\nonumber\\ &\Leftarrow \left(1+\frac{n(\mu_2-\muhatt{1}{n})^2}{\vhatt{1}{n}}\right)^{-\frac{n-1}{2}-\alpha} \ge \frac{\log 2}{A_{n,\alpha}T}\nonumber\\ &\phantom{wwwwwwwwwwwwwwwww}(\mbox{by $2^{x}\ge 1+x\log 2$})\nonumber\\ &\Leftrightarrow \frac{n(\mu_2-\muhatt{1}{n})^2}{\vhatt{1}{n}} \le \left(\frac{A_{n,\alpha}T}{\log 2}\right)^{\frac{1}{\frac{n-1}{2}+\alpha}}-1 \;=:\;C_T\enspace .\label{def_ct} \end{align} Therefore the expectation in \eqref{sita_id} is bounded from \eqref{eq_upper} as \begin{align} \lefteqn{ \mathrm{E}\left[ \frac{\idx{(\ptic{1}{n}{\mu_2})^T\le 1/2}}{\pti{1}{n}{\mu_2}} \right] }\nonumber\\ &\ge \!\!\! \iintt_{ \begin{array}{c} {\scriptstyle x\le \mu_2,\,s\ge 0,}\\ {\scriptstyle \frac{n(\mu_2-x)^2}{s}\le C_T} \end{array} } \!\!\! \frac{\mu_2-x}{\sqrt{s}} \left(1+\frac{n(\mu_2-x)^2}{s}\right)^{\frac{n}{2}+\alpha-1}\nonumber\\ &\phantom{wwwwwwwwwwwww}\cdot \sqrt{\frac{n}{2\pi}} \mathrm{e}^{-\frac{nx^2}{2}} \frac{s^{\frac{n-3}{2}}\mathrm{e}^{-\frac{s}{2}}}{2^{\frac{n-1}{2}}\Gamma(\frac{n-1}{2})} \mathrm{d} x \mathrm{d} s\nonumber\\ &= \frac{\sqrt{n}\mathrm{e}^{-\mu_2^2/2}}{\sqrt{\pi}2^{\frac{n}{2}}\Gamma(\frac{n-1}{2})} \!\!\! \iintt_{ \begin{array}{c} {\scriptstyle x\le\mu_2,\,s\ge 0,}\\ {\scriptstyle \frac{n(\mu_2-x)^2}{s}\le C_T} \end{array} } \!\!\! \mathrm{e}^{-\frac{n(\mu-x)^2}{2}+2\mu_2(\mu_2-x)} \nonumber\\ &\phantom{wi}\cdot (\mu_2-x) \left(1+\frac{n(\mu_2-x)^2}{s}\right)^{\frac{n}{2}+\alpha-1} s^{\frac{n}{2}-2}\mathrm{e}^{-\frac{s}{2}} \mathrm{d} x \mathrm{d} s\enspace .\nonumber \end{align} By letting \begin{align} (x,s)=\left( \mu_2-\sqrt{\frac{2zw}{n}},\, 2 z (1-w)\right)\enspace ,\nonumber \end{align} we obtain in a similar way to \eqref{term1} that \begin{align} \lefteqn{ \mathrm{E}\left[ \frac{ \idx{(\ptic{1}{n}{\mu_2})^T\le 1/2} }{\pti{1}{n}{\mu_2}} \right] }\nonumber\\ &\ge \frac{\mathrm{e}^{-\mu_2^2/2}}{2\sqrt{\pi n}\Gamma(\frac{n-1}{2})} \int_{0}^{\frac{1}{1+\frac{1}{C_T}}} \int_{0}^{\infty} \mathrm{e}^{-z}\mathrm{e}^{2\mu_2\sqrt{\frac{2zw}{n}}} \nonumber\\ &\qqua \cdot z^{\frac{n}{2}-1} (1-w)^{-1-\alpha}\mathrm{d} z \mathrm{d} w\nonumber\\ &\ge \frac{\mathrm{e}^{-\mu_2^2/2}}{2\sqrt{\pi n}\Gamma(\frac{n-1}{2})} \int_{0}^{\infty} \mathrm{e}^{2\mu_2\sqrt{\frac{2z}{n}}} \mathrm{e}^{-z}z^{\frac{n}{2}-1} \mathrm{d} z\nonumber\\ &\phantom{wi}\cdot \int_{0}^{\frac{1}{1+\frac{1}{C_T}}} (1-w)^{-1-\alpha}\mathrm{d} w\enspace . \qquad(\mbox{by $\mu_2<\mu_1=0$}) \nonumber \end{align} Here note that \begin{align} \int_{0}^{\frac{1}{1+\frac{1}{C_T}}} (1-w)^{-1-\alpha}\mathrm{d} w &= \begin{cases} \log (1+C_T),&\alpha=0,\\ \frac{(1+C_T)^{\alpha}-1}{\alpha},&\alpha>0. \end{cases}\nonumber \end{align} Then there exists a constant $B_{n,\alpha,\mu_2}$ such that \begin{align} \lefteqn{ \mathrm{E}\left[ \frac{ \idx{(\ptic{1}{n}{\mu_2})^T\le 1/2} }{\pti{1}{n}{\mu_2}} \right] }\nonumber\\ &\ge \begin{cases} B_{n,\alpha,\mu_2}\log (1+C_T),&\alpha=0,\\ B_{n,\alpha,\mu_2}((1+C_T)^{\alpha}-1),&\alpha>0. \end{cases}\label{impo_owari} \end{align} Finally by putting \eqref{bunkai_impo}, \eqref{sita_id} and \eqref{impo_owari} together we obtain for $\alpha=0$ that \begin{align} \lefteqn{ \mathrm{E}[\mathrm{Regret}(2T)] }\nonumber\\ &\ge \Delta_2\min\left\{ T-n,\, (1/2)B_{n,\alpha,\mu_2}\log (1+C_T)-1/2 \right\}.\nonumber \end{align} Eq.\,\eqref{cannot1} follows since $n\ge n_0$ is fixed and $C_T$ defined in \eqref{def_ct} is polynomial in $T$. Eq.\,\eqref{cannot2} for $\alpha>0$ is obtained in the same way. \end{proof} \section{Conclusion} We considered the stochastic multiarmed bandit problem such that each reward follows a normal distribution with an unknown mean and variance. We proved that Thompson sampling with prior $\pi(\mu_i,\sigma_i^2)\sim(\sigma_i^2)^{-1-\alpha}$ achieves the asymptotic bound if $\alpha<0$ but cannot if $\alpha\ge 0$, which includes reference prior $\alpha=0$ and Jeffreys prior $\alpha=1/2$. A future work is to examine whether TS with non-informative priors is risky or not for other multiparameter models where TS is used without theoretical analysis (see e.g., \emcite{thompson_empirical}). Since the analysis of this paper heavily depends on the specific form of normal distributions, it is currently unknown whether the technique of this paper can be applied to other models and this generalization remains as an important open problem.
\section{Introduction} \label{introduction} Observations made over the last 20 years have enabled the detection of several hundred exoplanets \citep[the first around a Solar mass star by][]{mayor_1995} and several thousand candidate systems \citep[identified, for instance, by the Kepler mission including the discovery of a system of six planets and a sub--Mercury sized planet, see][respectively]{lissauer_2011,barclay_2013}. Surveys of variability (detecting planetary transits) and radial velocity have also provided estimates of the mass and orbital radii of these exoplanets. Such surveys are most sensitive to giant ($\sim$Jovian mass) planets which orbit close to their parent stars, experience intense radiation ($10^3-10^5$ times that received by Jupiter, exacerbating problems involved with simplified radiative transfer schemes), and are termed `hot Jupiters'. The strong tidal forces experienced by these planets is thought to lead to rapid synchronisation of their rotation period with their orbital period (with the adoption of a reasonable dissipation parameter). This `tidal--locking' provides a strong constraint on the planetary rotation rate and means the planet has a permanent day and night side, experiencing net heating and cooling, respectively \citep[see][ and references therein]{baraffe_2010}. Furthermore, precise observations of the luminosity as a function of time and wavelength (transit spectroscopy) of a transiting star--planet system can be used to probe the planet's atmospheric conditions \citep[see][for review]{seager_2010}. Observations of the primary eclipse (when the planet transits in front of the star) have provided the detection of specific species in the atmospheres of hot Jupiters \citep[see for example][who detected potassium and sodium, respectively, in the atmosphere of Xo--2b]{sing_2011,sing_2012} as well as the detection of possible dust or hazes \citep[for example as found in HD 189733b,][]{pont_2012}. Additionally observations of both the primary and secondary eclipses (when the planet moves behind the star) have allowed the derivation of day and night side atmospheric temperatures \citep[for example $\sim$1250 K and $\sim$1000 K for HD 189733b as found by][]{knutson_2007,knutson_2009}. Moreover, using the full orbital luminosity phase--curve, the atmospheric temperature as a function of planetary longitude can be inferred \citep[see for example][]{knutson_2007,knutson_2009}. Analysis of these temperature `maps' has revealed offsets of the hottest part of the atmosphere or `hot spot' (at a given depth) from the sub-stellar points. This offset was predicted by \citet{showman_2002} as a consequence of the expected fast circulations induced by the large--scale heating. The presence of such fast winds has been suggested for HD 209458b using Doppler shifting of molecular CO bands \citep{snellen_2010}, where $\sim$km$\,$s$^{-1}$ wind speeds were derived. Although this is not a complete review of the observational results, the observations have produced several challenges for our theoretical models of planetary evolution. Many of these challenges require an understanding of the full three dimensional circulation, including the vertical transport. Firstly, comparison of the derived radii with the predictions of planetary interior models (as a function of age) has shown that some hot Jupiters appear inflated. \citet{guillot_2002} suggested that the vertical transport of $\sim1$ percent of the incident stellar flux, from the top of the atmosphere deep into the planet interior, could halt the planet's gravitational contraction sufficiently to explain the observations. \citet{showman_2002} then suggested that the required levels of kinetic energy could be generated by the large--scale forcing expected in hot Jupiter atmospheres. Secondly, as is evident for Solar system planets, significant abundances of scattering particles can dominate the global heat balance of a planet's atmosphere \citep[see for discussion][]{sanchez_lavega_2004}, which is difficult to capture using simplistic radiative transfer schemes. The possible presence of scattering particles \citep[suggested to be MgSiO$_3$ by][]{etangs_2008} in the observable atmospheres of hot Jupiters requires they be supported against gravitational settling and/or be replenished via circulations. Therefore, vertical transport modeled over a large range of pressures is vital to interpreting these observations, in addition to a non-grey radiative transfer scheme. Finally, comparison of day and night side temperatures has revealed a possible dichotomy separating hot Jupiters with efficient heat redistribution (from day to night side), which are generally more intensely irradiated, from those exhibiting less efficient heat redistribution. The efficiency of redistribution has been linked to the existence of a region of the hot Jupiter upper atmosphere where temperature increases with height, a thermal inversion, as inferred from model fitting for HD 209458b \citep{knutson_2008}, and thereby the presence of absorbing substances such as VO and TiO \citep{hubeny_2003,fortney_2008}. A correct description of all these processes requires a non--grey radiative transfer scheme coupled to a dynamical model of the atmospheric redistribution. The glimpses into the atmospheres of hot Jupiters provided by the observations, and the associated puzzles, have motivated the application of Global Circulation Models (GCMs), usually developed for the study of Earth's weather and climate, to hot Jupiters. GCMs are generally comprised of many components or modules which handle different aspects of the atmosphere. Many of these components are highly optimised for conditions on Earth, for example treatments of the surface boundary layer. To apply GCMs to planets other than Earth adaptation of the most fundamental components i.e. treating radiative transfer and dynamical motions, is required. Further adaptations to more detailed atmospheric process can then occur when merited by observations. GCMs have been successfully applied to model other Solar system planets \citep[see for example models of Jupiter, Saturn, Mars and Venus:][respectively]{yamazaki_2004,muller_wodarg_2006,hollingsworth_2011,lebonnois_2011}, but hot Jupiters present a very different regime. The latter receive significantly more radiation and rotate much more slowly than the giant planets in our Solar system. Therefore, the characteristic scales of atmospheric features such as the expected vortex size, the Rossby deformation radius and the elongation in the east--west direction of wind structures, the Rhines scale are both approximately the size of the planet \citep[proportionally much larger than for Solar system planets, see][for a review and comparison with Solar system planets]{showman_2011}. This effectively means that one might expect any `weather systems', or jet structures (i.e. prevailing circumplanetary flows) to be comparable in size with the horizontal extent of the atmosphere. Despite the exotic nature of the flow regime, the adaptation of GCMs to hot Jupiters has met with success as, for example, several models have been able to demonstrate that offsets in the `hot spot' are consistent with redistribution from zonal (longitudinal direction) winds \citep{showman_2009,dobbs_dixon_2008,dobbs_dixon_2009}. The progress of the modelling has been reviewed in \citet{showman_2008,showman_2011} and a useful summary of the different approaches taken can be found in \citet{dobbs_dixon_2012}. To interpret the observations of hot Jupiters the regime dictates the model should include solution to the full three--dimensional equations of motion for a rotating atmosphere coupled to a non--grey radiative transfer scheme. This will allow exploration of the consequences of realistic vertical transport and its interaction with the horizontal advection, and include the effect on the thermal balance of the atmosphere caused by frequency dependent opacities. Most GCMs applied to hot Jupiters solve the primitive equations of meteorology, involving the approximation of vertical hydrostatic equilibrium, and a `shallow--atmosphere' \citep[combining the constant gravity, `shallow--fluid' and `traditional' approximations, see][]{vallis_2006,white_2005}. The most sophisticated radiative transfer scheme within a GCM, applied to hot Jupiters, to date is that of \citet{showman_2009} which solves the primitive equations coupled to a simplified radiative transfer scheme based on the two-stream, correlated--$k$ method. However, the approximations involved in the primitive equations neglect the vertical acceleration of fluid parcels, and the effect of the vertical velocity on the horizontal momentum. More complete dynamical models, solving the full Navier--Stokes equations, have been applied to hot Jupiters by \citet{dobbs_dixon_2008,dobbs_dixon_2009,dobbs_dixon_2010,dobbs_dixon_2012}, but these models include a radiative transfer scheme more simplified than the method of \citet{showman_2009}. \citet{dobbs_dixon_2010} includes frequency dependent radiative transfer via the introduction of only three opacity bins (and generally runs for short elapsed model times). \citet{dobbs_dixon_2012} includes a treatment using a similar number of frequency bins to \citet{showman_2009}, but simply average the opacity in each bin as opposed to generating opacities via the correlated--k method. Therefore, calculations based on non--grey radiative transfer coupled to full three--dimensional equations of motion for a rotating atmosphere do not yet exist. Additionally, current models applied to hot Jupiters are still missing many other physical processes. Although not discussed in this paper, treatments of the magnetic fields, photochemistry and clouds or hazes, may well be required to create a model capable of meaningful predictions. We are beginning work on the incorporation of a photochemical network and the simple modelling of clouds into our model, but this will take some time to complete. The ENDGame (Even Newer Dynamics for General atmospheric modelling of the environment) dynamical core (the part of the GCM solving the discretised fluid dynamics equations of motion) of the UK Met Office GCM, the Unified Model (UM) is based on the non--hydrostatic deep--atmosphere equations \citep{staniforth_2003,staniforth_2008,wood_2013}, and does not make the approximations incorporated in the primitive equations \citep{white_2005}. The UM also includes a two--stream radiative transfer scheme with correlated--k method. This code has previously been adapted to studies of Jupiter \citep{yamazaki_2004}, but requires significant adaptation of the dynamical and radiative transfer schemes to be applied to hot Jupiters. We have previously presented the satisfactory completion of several Earth--like test cases of the dynamical core in \citet{mayne_2013}. Now, we have completed the adaptation of the dynamical core, and in this work present the first hot Jupiter test cases. We have completed the Shallow--Hot Jupiter \citep[SHJ, as prescribed in][]{menou_2009} and the HD 209458b test case of \citet{heng_2011}. Adaptation of the radiative transfer scheme is nearing completion and coupled models will be presented in a future work. With this paper, we begin a series in which we will present the details of the model developments and testing of the UM, as it is adapted for the study of exoplanets, as well as scientific applications and results. The structure of the paper is as follows. In Section \ref{model} we detail the model used including the equations solved and highlight the important details of our boundary conditions and numerical scheme. We also discuss the effect of canonical simplifications made to the dynamical equations. Section \ref{model} also includes an explanation of the parameterisations used and references to a more detailed description of the model and previous testing. Section \ref{test_cases} then describes the setups for two test cases we have run including the parameter values and temperature and pressure profiles. We also, in Section \ref{test_cases} demonstrate satisfactory completion of these tests. Section \ref{discussion} highlights some problems with the test cases and discusses future work. Finally, in Section \ref{conclusions} we include a summary of our conclusions. \section{Model} \label{model} The UM dynamical core called ENDGame is explained in detail in \citet{wood_2013}. The code is based on the non--hydrostatic, deep--atmosphere (NHD) equations of motion for a planetary atmosphere \citep{staniforth_2003,staniforth_2008}, including a varying (with height) gravity and a geometric height vertical grid. Uniquely, the code allows solution to the non--hydrostatic shallow--atmosphere \citep[NHS,][]{staniforth_2003,staniforth_2008} equations, or just the simple assumption of a constant gravity (to create a quasi--NHD system), within the same numerical scheme. \subsection{Overview of the numerical scheme} \label{numerics} The UM is a finite--difference code where the atmosphere is discretised onto a latitude--longitude grid (resolutions explained in Section \ref{test_cases}), using a staggered Arakawa--C grid \citep{arakawa_1977} and a vertically staggered Charney--Phillips grid \citep{charney_1953}. The code uses a terrain following height--based vertical coordinate\footnote{Although for this work we include no orography, and have no `surface'.}. The code is semi--Lagrangian and semi--implicit, where the latter is based on a Crank--Nicolson scheme. The code employs semi--Lagrangian advection where the values for advected quantities are derived at interpolated departure points, and are then used to calculate quantities within the Eulerian grid. For the semi--implicit scheme the temporal weighting between the $i$th and the $i+1$th state is set by the coefficient $\alpha$ which can vary between zero and one, and is set to 0.55 in this work. For each atmospheric timestep a nested iteration structure is used. The \textit{outer} iteration performs the semi--Lagrangian advection (including calculation of the departure points), and values of the pressure increments from the \textit{inner} iteration are back substituted to obtain updated values for each prognostic variable. The \textit{inner} iteration solves the Helmholtz (elliptical) problem to obtain the pressure increments, and the Coriolis and nonlinear terms are updated. The velocity components are staggered such that the meridional velocity is defined at the pole \citep[see][for a more detailed explanation]{mayne_2013}, but no other variable is stored at this location, thereby avoiding the need to solve for pressure at the poles of the latitude--longitude grid \citep{wood_2013}. \citet{thuburn_2004} show that mass, angular momentum and energy are much more readily conserved with a grid staggered such that $v$ and not $u$ is held at the pole. The stability afforded by the spatial and temporal discretisation removes the need for an explicit polar filter (although our diffusion operator has some aspects in common with a polar filter, see discussion in Section \ref{diffusion}). The code adopts SI units. A full description of the code can be found in \citet{wood_2013} and important features relating to the reproduction of idealised tests are reiterated in \citet{mayne_2013}. \subsection{Previous Testing} \label{testing} The UM undergoes regular verification for the Earth system, and \citet{wood_2013} completes several tests from the Dynamical Core Model Intercomparison Project\footnote{DCMIP, see \url{http://earthsystemcog.org/projects/dcmip-2012/}.}, and the deep--atmosphere baroclinic instability test \citep{ullrich_2013}, using the ENDGame dynamical core. We have also, as part of the adaptation to exoplanets completed several tests for an Earth--like model including the Held-Suarez test \citep{held_1994}, the Earth-like test of \citet{menou_2009} and the Tidally Locked Earth of \citet{merlis_2010}, the results of which are presented in \citet{mayne_2013}. Additionally, for each setup used we also complete a static, non--rotating, hydrostatic isothermal atmosphere test, ensuring that the horizontal and vertical velocities recorded are negligibly small and do not grow significantly with time (when run for a few million iterations). For simulations we have performed, the longest of which is many millions of iterations, mass and angular momentum, are conserved to better than $\lesssim 0.05$\% and $\sim 5$\%, respectively. In the UM mass is conserved via a correction factor applied after each timestep \citep[see][for details]{wood_2013}. \subsection{Equations solved by the dynamical core} \label{equations_solved} We model only a section, or spherical shell, of the total atmosphere and define the material below our inner boundary (discussed in more detail in Section \ref{boundary}) as the `planet' and the subscript ${\rm p}$ denotes quantities assigned to this region. The dynamical core solves a set of five equations: one for each momentum component, a continuity equation for mass and a thermodynamical energy equation, which are closed by the ideal gas equation. These equations are (using the ``Full'' equation set, see Table \ref{model_names} for explanation) \begin{align} \frac{Du}{Dt}=\frac{uv\tan\phi}{r}-\frac{uw}{r}+fv-f^{\prime}w-\frac{c_p\theta}{r \cos\phi}\frac{\partial \Pi}{\partial \lambda}+\textbf{D}(u),\\ \frac{Dv}{Dt}=-\frac{u^2\tan\phi}{r}-\frac{vw}{r}-uf-\frac{c_p\theta}{ r}\frac{\partial \Pi}{\partial \phi}+\textbf{D}(v),\\ \delta\frac{Dw}{Dt}=\frac{u^2+v^2}{r}+uf^{\prime}-g(r)-c_p\theta\frac{\partial \Pi}{\partial r},\\ \frac{D\rho}{Dt}=-\rho \left[ \frac{1}{r \cos \phi}\frac{\partial u}{\partial \lambda} \right. +\frac{1}{r \cos\phi}\frac{\partial (v\cos \phi)}{\partial \phi}\left. +\frac{1}{r^2}\frac{\partial (r^2w)}{\partial r} \right],\\ \frac{D\theta}{Dt}=\frac{Q}{\Pi}+\textbf{D}(\theta),\\ \Pi^{\frac{1-\kappa}{\kappa}}=\frac{R\rho\theta}{p_0}, \label{full_set} \end{align} respectively. The coordinates used are $\lambda$, $\phi$, $r$ and $t$, which are the longitude, latitude (from equator to poles), radial distance from the centre of the planet and time. The spatial directions, $\lambda$, $\phi$ and $r$, then have associated wind components $u$ (zonal), $v$ (meridional) and $w$ (vertical). $c_p$ is the specific heat capacity, $R$ is the gas constant and $\kappa$ the ratio of the $R/c_{p}$. $\delta$ is a `switch' (0 or 1) to enable a quasi--hydrostatic version of the equations \citep[not used in this work but detailed in][]{white_2005}. $p_0$ is a chosen reference pressure and $g(r)$ is the acceleration due to gravity at $(r)$ and is defined as \begin{equation} g(r)=g_{\rm p}\left( \frac{R_{\rm p}}{r}\right)^2, \label{grav_eg} \end{equation} where $g_{\rm p}$ and $R_{\rm p}$ are the gravitational acceleration and radial position at the inner boundary. $f$ and $f^{\prime}$ are the Coriolis parameters defined as, \begin{equation} f=2\Omega\sin\phi, \end{equation} and \begin{equation} f^{\prime}=2\Omega\cos\phi, \end{equation} where $\Omega$ is the planetary rotation rate. $\rho$ and $\theta$ are the prognostic variables of density and potential temperature, respectively. $\Pi$ is the Exner pressure (or function). $\theta$ and $\Pi$ are then defined, in terms of the temperature, $T$ and pressure, $p$, as \begin{equation} \theta = T\left(\frac{p_0}{p}\right) ^{\frac{R}{c_p}},\\ \end{equation} and \begin{equation} \Pi=\left(\frac{p}{p_0}\right)^{R/c_{p}} = \frac{T}{\theta},\\ \end{equation} respectively. The material derivative, $\frac{D}{Dt}$ is given by \begin{equation} \frac{D}{Dt}=\frac{\partial }{\partial t}+\frac{u}{r\cos\phi}\frac{\partial }{\partial \lambda}+\frac{v}{r}\frac{\partial }{\partial \phi}+w\frac{\partial }{\partial r}.\\ \end{equation} Finally, $Q$ and $\textbf{D}$ are the heating rate and diffusion operator (note that diffusion is not applied to the vertical velocity), respectively. The heating, in this work, is applied using a temperature relaxation or Newtonian cooling scheme discussed in Section \ref{radiative_transfer}. The diffusion operator is detailed in Section \ref{diffusion}. \subsubsection{Dynamical simplification and variants of the equations of motion} \label{dynamic_switches} A quartet of self--consistent governing dynamical equations conserving axial angular momentum, energy and potential vorticity are described in detail in \citet{white_2005}. These are the hydrostatic primitive equations or (hydrostatic) primitive equations (HPEs), quasi--hydrostatic equations (QHEs), the non--hydrostatic shallow--atmosphere (NHS) equations and the non--hydrostatic deep--atmosphere (NHD) equations. For this work we, using ENDGame, solve the NHD (which are detailed in Section \ref{equations_solved}), the quasi--NHD (with a constant gravity) and the NHS equations. \citet{white_2005} includes a full discussion of the assumptions made in each equation set, the most relevant (for this work) of which are included in Table \ref{model_names} alongside the validity criteria, and an estimate for the validity on HD 209458b (an example hot Jupiter). Table \ref{model_names} also includes the short reference names we have used to describe each setup. We adopt the nomenclature of \citet{white_2005}, where the `shallow--atmosphere' approximation implies constant gravity, in addition to the adoption of the `shallow--fluid' and `traditional' approximations. In this work we run simulations using the ``Full'' (NHD), ``Deep'' (quasi--NHD i.e. constant gravity) and ``Shallow'' (NHS) equations sets, where the primitive equations (HPEs) are included as illustrative of the codes we compare against. However, the GCMs we are comparing with use either pressure or $\sigma$ ($=p / p_{\rm surf}$, where $p_{\rm surf}$ is the pressure at the inner boundary, which is usually called the ``surface'' for terrestrial planets) as their vertical coordinate. The key point is that when we compare models using for instance the ``Shallow'' equations with the primitive equations, we are simply relaxing the assumption of hydrostatic equilibrium. Moving to the ``Deep'' equations then involves further relaxing the `shallow--fluid' and `traditional' approximation (but retaining a constant gravity) and finally, the ``Full'' equations include a further relaxation of the constant gravity approximation. \begin{table*} \caption{Key assumptions in each equation set, the local name used to describe the set, and the validity, both in general and for HD 209458b. $\Phi$ and $G$ are the geopotential and gravitational constant respectively. $M_{\rm p}$ and $M_{\rm atm}$ are the total masses of the planet below the inner boundary, and of the atmosphere, respectively. $z$ is the vertical height from the inner boundary. $H$ and $L$ are the vertical and horizontal sizes of the atmospheric domain. Finally, $N$ is the buoyancy (or Brunt-V\"ais\"al\"a) frequency, $N=\sqrt{-\frac{g(r)}{\rho_0}\frac{\partial\rho(r)}{\partial r}}$). SI units are used unless otherwise stated.} \label{model_names} \centering \begin{tabular}{ccccllcc} \hline\hline \multicolumn{4}{c}{Name}&\multicolumn{2}{l}{Approximation}&Formal Condition&HD 209458b\\ \hline \hline \multirow{7}{*}{Primitive}\rdelim\{{7}{0mm}[]&\multirow{6}{*}{Shallow}\rdelim\{{6}{0mm}[]&\multirow{3}{*}{Deep}\rdelim\{{3}{0mm}[]&\multirow{2}{*}{Full}\rdelim\{{2}{0mm}[]&Spherical geopotentials$^{(1)}$&$\Phi(\lambda,\phi,r)=\Phi (r)$&$\Omega^2r\ll g$$^{(2)}$&$10^{-2}\ll 10^{1}$\\ &&&&no self-gravity&$g(r)=\frac{GM_{\rm p}}{r^2}$&$M_{\rm atm}\ll M_{\rm p}$&$\sim 10^{23}\ll 10^{27}$\\ &&&&constant gravity&$g(r)=g_{\rm p}=\frac{GM_{\rm p}}{R_{\rm p}^2}$&$z\ll R_{\rm p}$&\rdelim\{{2}{4mm}[]\multirow{2}{*}{$\sim 10^{7} < 10^{8}$}\\ &&&&`shallow--fluid'&$r\rightarrow R_{\rm p}$ \& $\frac{\partial}{\partial r}\rightarrow\frac{\partial}{\partial z}$&$z\ll R_{\rm p}$&\\ &&&&\multirow{2}{*}{`traditional'}&$\frac{uw}{r}$, $\frac{vw}{r}$, $2\Omega w\cos\phi$ $\rightarrow 0$&\multirow{2}{*}{$N^2\gg \Omega^2$$^{(3)}$}&\multirow{2}{*}{$\sim 10^{-5}\gg 10^{-10}$}\\ &&&&&$\frac{u^2+v^2}{r}$, $2\Omega u\cos\phi$ $\rightarrow 0$&&\\ &&&&hydrostasy&$\frac{\partial \Pi}{\partial r}=-\frac{g}{c_p\theta}$ or $\frac{\partial p}{\partial r}=-\rho g$&$H \ll L$&$\sim 10^{7} < 10^{8}$\\ \hline \end{tabular} \tablebib{(1) For a full discussion on the impact of the spherical geopotentials approximation see \citet{white_2008}. (2) This condition neglects tidal deformation, essentially assuming the planetary gravitational field is well isolated (see discussion in text). (3) Condition from \citet{phillips_1968}, but may not be sufficient \citep[see discussion in][]{white_1995}.} \end{table*} The assumptions pertaining to gravity require some explanation. Firstly, the gravitational potential of the planet is assumed to be well isolated from external gravity fields \citep[this is not the case in some hot Jupiters, for instance Wasp--12,][]{li_2010}. Secondly, care must be taken over how to solve for the centrifugal force, and subsequently construct the gravitational potential. When modelling the Earth the acceleration due to gravity can be measured at the surface, $g_{\rm p}$. This is in effect the acceleration due to the \textit{apparent} gravity as it includes contributions from both the gravitational and centrifugal components \footnote{In reality the surface of the Earth has deformed such that the local \textit{apparent} gravity acts normal to the surface.}. This combined gravitational and centrifugal potential, or geopotential is then, in most cases, assumed to be spherically symmetric. This means, however, that the divergence of the resulting combined potential is not zero, as should be the case \citep[see][ for a detailed discussion of the spherical geopotentials approximation]{white_2008}. Additionally, this spurious divergence in the combined potential is increased if one adopts a constant gravitational acceleration throughout the atmosphere, as opposed to allowing it to fall via an inverse square law \citep{white_2012}. Calculating the acceleration due to the gravitational potential only, and solving explicitly, as part of the dynamical equations, for the centrifugal component, however, introduces spurious motions. For example a modeled hydrostatically balanced and statically stable atmosphere, at rest, would subsequently have to adjust to the \textit{apparent} gravity caused by the rotation, which generates a horizontal force, creating winds. For hot Jupiters the acceleration due to gravity cannot be measured, and there is no surface, in the same sense as on Earth or any other terrestrial planet. Therefore a value for $g_{\rm p}$ must be estimated using measurements of the total mass of the planet, $M_{\rm p}$ (derived from radial velocity measurements), and assuming this to be contained within a radius, $R_{\rm p}$ (practically the smallest available radius derived from observations of the primary eclipse). The precision to which $g_{\rm p}$ is estimated, or quoted, is much lower than the magnitude of the expected effect of the centrifugal component. Therefore, although formally, we absorb the centrifugal term into the gravity field, due to its negligible, relative magnitude, we prefer to state that, dynamically we neglect this term. Finally, most GCMs also neglect the gravity of the atmosphere itself. For hot Jupiters, whether the gravitational potential of the atmosphere can be neglected depends on the distribution of mass between the atmosphere itself and the `planet' below the inner boundary, whose mass defines $g_{\rm p}$. Formally, \begin{eqnarray} g&=&\frac{GM(r)}{r^2}\\ &=&\frac{G}{r^2}\left[ M_{\rm p}+M_{\rm atm}(r) \right]\\ &=&g_{\rm p}\left( \frac{R_{\rm p}}{r}\right)^2+\frac{GM_{\rm atm}(r)}{r^2}\\ &\sim&g_{\rm p}\left( \frac{R_{\rm p}}{r}\right)^2 \mbox{,}\, M_{\rm atm}(r) \ll M_{\rm p}, \label{grav} \end{eqnarray} where $M_{\rm atm}(r)=M(r)-M(R_{\rm p})$. The momentum and continuity equations differ depending on the assumptions made in each of the cases shown in Table \ref{model_names}. \citet{white_2005} explores, in detail, the form of the metric and differential operators. In Table \ref{eqn_sets} we express (in expanded form but omitting the diffusion terms) the relevant parts of the equations sets which are illustrative of the main differences, for the three equation sets we use (i.e. ``Full'', ``Deep'' and ``Shallow''), and also the primitive equations for comparison. \subsubsection{Consequences of approximations} \label{approx} Comparing the terms in the equations in Table \ref{eqn_sets} it is apparent that each progressive relaxation of an approximation acts to introduce extra `exchange' terms (and alter existing ones), or terms in each momentum equation involving the other components of momenta. Focusing on the $u$ and $v$ components of Table \ref{eqn_sets} the ``shallow--atmosphere'' approximation neglects the terms $uw / r$ and $2\Omega w\cos\phi$, and alters the term $uv\tan\phi / r$. Clearly, regardless of whether $w$ is small compared to $u$, this assumption, by definition, eliminates the exchange of vertical and zonal momentum present in a real atmosphere (similarly for the $v$ component). The omission of the metric and Coriolis terms is termed the `traditional' approximation, as explained in Table \ref{model_names}. Critically, the physical justification for the adoption of this approximation is weak, and it is largely taken with the ``shallow--fluid'' approximation to enable conservation of angular momentum and energy, not for physically motivated reasons. We present, in Table \ref{model_names} an expression for the validity of this expression, however, this is debatable and assumes a lack of planetary scale flows \citep[see discussion in][]{white_1995}. Given that planetary scale flows are expected for hot Jupiters, this approximation may well prove crucial to the reliability of the results of hot Jupiter models. \citet{white_1995} show that the term $2\Omega w\cos\phi$ in the zonal momentum equation may be neglected if $2\Omega H\cos\phi/U\ll 1$, which for HD 209458b gives $\sim0.1\ll 1.0$, suggesting it is marginally valid only for the regions of peak zonal velocity. The `traditional' approximation also removes terms from the vertical momentum equation involving $u$ and $v$, further inhibiting momentum exchange. Previous attempts have been made to isolate the effect of this approximation \citep[see for example][]{cho_2011}. \citet{tokano_2013} show that GCMs adopting the primitive equations do not correctly represent the dynamics of Titan's atmosphere (as well as indicating it may be problematic for Venus's atmosphere). Although \citet{tokano_2013} focus on the assumption of hydrostatic equilibrium, the term they indicate is dominant, $(u^2+v^2) / r$, is neglected as part of the `traditional' approximation. The lack of coupling between the vertical and horizontal momentum in the HPEs is exacerbated by the adoption of vertical hydrostatic equilibrium, which neglects the vertical acceleration of fluid parcels. Vertical velocities are still retained in the HPEs, derived from the continuity equation, but the lack of coupling between the vertical and horizontal components of momentum in the HPEs means these are unlikely to be realistic. As discussed in Section \ref{introduction} several key physical problems require a well modeled interaction of the vertical and horizontal circulations, and between the deep and shallow atmosphere. Modelling the atmosphere using the NHD equations therefore allows us to present a much more self--consistent and complete model of the atmospheric flow. However, vertical velocities are generally much smaller than the zonal or meridional flows \citep[up to two orders of magnitude smaller,][]{showman_2002}, and relaxation times (both radiative and dynamical) in the deeper atmospheres are generally orders of magnitude longer than those in the shallow atmosphere. Therefore, the effects of replacing the `exchange' terms in the dynamical equations may only be appropriate for simulations run much longer than usual. Additionally, the introduction of a non--constant (with height) gravity also subtly affects the stratification, and therefore the vertical transport. In an atmosphere in hydrostatic equilibrium the stratification is proportional to the gravitational acceleration, as the weight of the atmosphere above must be supported from below. Therefore, allowing gravity to vary (as described by Equation \ref{grav_eg}) from the value assumed at the `surface' or inner boundary effectively weakens it throughout the atmosphere, and therefore weakens the stratification reducing its inhibiting effect on vertical motions. In our HD 209458b test case over the vertical domain $\Delta g/g \sim 0.2$. In summary, the assumption of a `shallow--atmosphere' which includes the `shallow--fluid', `traditional' and constant gravity approximations, effectively neglects exchange between the vertical and horizontal momentum, and is likely to inhibit vertical motions, or produce inconsistent vertical velocities. Yet, vertical transport and its interaction with the horizontal advection is believed to be critical to understanding the major scientific questions regarding hot Jupiter atmospheres. In this work we find that the results of the test case, when run using the less simplified dynamical model, diverge from the literature results. This divergence is caused by the improved representation of vertical motions and exchange between the horizontal and vertical components of momentum (discussed in more detail in Section \ref{discussion}). \begin{sidewaystable*} \caption{List of momenta and continuity equations (where we have expanded the material derivative, $\frac{D}{Dt}$, but omitted diffusion or damping terms) for each of the named equation sets with approximations detailed in Table \ref{model_names}. The values of $g$ and $g_{\rm p}$ are discussed in Section \ref{equations_solved}.} \label{eqn_sets} \centering \fontsize{14}{16.8} \begin{tabular}{cccccccccccccl} \hline\hline \multicolumn{2}{l}{Variable}&\multicolumn{4}{c}{Material Derivative $\frac{D}{Dt}$}&\multicolumn{5}{c}{Terms}&Name\\ \hline \multirow{3}{*}{$u$}&\rdelim\{{2}{0mm}[]&$\left[ \frac{\partial u}{\partial t} \right. $&+$\frac{u}{r\cos\phi}\frac{\partial u}{\partial \lambda}$&$+\frac{v}{r}\frac{\partial u}{\partial \phi}$&$\left. +w\frac{\partial u}{\partial r}\right]=$&$+\frac{uv\tan\phi}{r}$&$-\frac{uw}{r}$&$+2\Omega v\sin\phi$&$-2\Omega w\cos\phi$&$-\frac{c_p\theta}{r \cos\phi}\frac{\partial \Pi}{\partial \lambda}$&\textit{Full \& Deep}\\ &&$\left[ \frac{\partial u}{\partial t}\right. $&+$\frac{u}{R_{\rm p}\cos\phi}\frac{\partial u}{\partial \lambda}$&$+\frac{v}{R_{\rm p}}\frac{\partial u}{\partial \phi}$&$\left. +w\frac{\partial u}{\partial z}\right]=$&$+\frac{uv\tan\phi}{R_{\rm p}}$&&$+2\Omega v\sin\phi$&&$-\frac{c_p\theta}{ R_{\rm p}\cos\phi}\frac{\partial \Pi}{\partial \lambda}$&\textit{Shallow \& Primitive}\\ \hline \multirow{3}{*}{$v$}&\rdelim\{{2}{0mm}[]&$\left[ \frac{\partial v}{\partial t}\right. $&+$\frac{u}{r\cos\phi}\frac{\partial v}{\partial \lambda}$&$+\frac{v}{r}\frac{\partial v}{\partial \phi}$&$\left. +w\frac{\partial v}{\partial r}\right]=$&$-\frac{u^2\tan\phi}{r}$&$-\frac{vw}{r}$&$-2\Omega u\sin\phi$&&$-\frac{c_p\theta}{ r}\frac{\partial \Pi}{\partial \phi}$&\textit{Full \& Deep}\\ &&$\left[ \frac{\partial v}{\partial t}\right. $&+$\frac{u}{R_{\rm p}\cos\phi}\frac{\partial v}{\partial \lambda}$&$+\frac{v}{R_{\rm p}}\frac{\partial v}{\partial \phi}$&$\left. +w\frac{\partial v}{\partial z}\right]=$&$-\frac{u^2\tan\phi}{R_{\rm p}}$&&$-2\Omega u\sin\phi$&&$-\frac{c_p\theta}{R_{\rm p}}\frac{\partial \Pi}{\partial \phi}$&\textit{Shallow \& Primitive}\\ \hline \multirow{4}{*}{$w$}&\rdelim\{{4}{0mm}[]&$\delta\left[ \frac{\partial w}{\partial t}\right. $&+$\frac{u}{r\cos\phi}\frac{\partial w}{\partial \lambda}$&$+\frac{v}{r}\frac{\partial w}{\partial \phi}$&$\left.+w\frac{\partial w}{\partial r}\right]=$&\multicolumn{2}{c}{$+\frac{u^2+v^2}{r}$}&$+2\Omega u \cos\phi$&$-g(r)$&$-c_p\theta\frac{\partial \Pi}{\partial r}$&\textit{Full}\\ &&$\delta\left[ \frac{\partial w}{\partial t}\right. $&+$\frac{u}{r\cos\phi}\frac{\partial w}{\partial \lambda}$&$+\frac{v}{r}\frac{\partial w}{\partial \phi}$&$\left. +w\frac{\partial w}{\partial r}\right]=$&\multicolumn{2}{c}{$+\frac{u^2+v^2}{r}$}&$+2\Omega u \cos\phi$&$-g_{\rm p}$&$-c_p\theta\frac{\partial \Pi}{\partial r}$&\textit{Deep}\\ &&$\delta\left[ \frac{\partial w}{\partial t}\right. $&+$\frac{u}{R_{\rm p}\cos\phi}\frac{\partial w}{\partial \lambda}$&$+\frac{v}{R_{\rm p}}\frac{\partial w}{\partial \phi}$&$\left.+w\frac{\partial w}{\partial z}\right]=$&&&&$-g_{\rm p}$&$-c_p\theta\frac{\partial \Pi}{\partial z}$&\textit{Shallow}\\ &&\multicolumn{4}{r}{$0=$}&&&&$-g_{\rm p}$&$-c_p\theta\frac{\partial \Pi}{\partial z}$&\textit{Primitive}\\ \hline \multirow{2}{*}{$\rho$}&\rdelim\{{2}{0mm}[]&$\left[ \frac{\partial \rho}{\partial t}\right. $&+$\frac{u}{r\cos\phi}\frac{\partial \rho}{\partial \lambda}$&$+\frac{v}{r}\frac{\partial \rho}{\partial \phi}$&$\left. +w\frac{\partial \rho}{\partial r}\right]=$&\multicolumn{2}{c}{$-\rho \left[ \frac{1}{r \cos \phi}\frac{\partial u}{\partial \lambda} \right. $}&\multicolumn{2}{c}{$+\frac{1}{r \cos\phi}\frac{\partial (v\cos \phi)}{\partial \phi}$}&$\left. +\frac{1}{r^2}\frac{\partial (r^2w)}{\partial r} \right]$&\textit{Full \& Deep}\\ &&$\left[\frac{\partial \rho}{\partial t}\right. $&+$\frac{u}{R_{\rm p}\cos\phi}\frac{\partial \rho}{\partial \lambda}$&$+\frac{v}{R_{\rm p}}\frac{\partial \rho}{\partial \phi}$&$\left. +w\frac{\partial \rho}{\partial z}\right]=$&\multicolumn{2}{c}{$-\rho \left[ \frac{1}{R_{\rm p} \cos \phi}\frac{\partial u}{\partial \lambda} \right. $}&\multicolumn{2}{c}{$+\frac{1}{R_{\rm p} \cos\phi}\frac{\partial (v\cos \phi)}{\partial \phi}$}&$\left. +\frac{\partial w}{\partial z} \right]$&\textit{Shallow \& Primitive}\\ \hline \hline \end{tabular} \end{sidewaystable*} \subsection{Boundary conditions} \label{boundary} A full discussion of the boundary conditions used is presented in \citet{wood_2013}, here we emphasis a few key characteristics. Given that hot Jupiters do not have a solid surface, we impose an inner boundary, which is frictionless (placed at $R_{\rm p}$). The inner and outer boundaries are rigid and impermeable \citep[to ensure energy and axial momentum conservations,][]{staniforth_2003}. As the boundaries are rigid they nonphysically act to reflect vertically propagating waves, such as acoustic or gravity waves, back into the domain. This is usually only significant during an initial `spin--up' period as initial transients are produced, in particular waves generated by the adjustment of the mass distribution in the atmosphere. To solve this problem the UM incorporates, into the upper boundary, a damping region (termed a `sponge' layer) high up at the top of the atmosphere to mitigate the spurious reflection of vertical motions at the upper boundary. Vertical damping of vertical velocities is incorporated using the formulation of \citet{melvin_2010} \citep[which follows][]{klemp_2008}, \begin{equation} w^{t +\Delta t}=w^{t \*}-R_{w}\Delta t w^{t +\Delta t},\\ \end{equation} where $w^t$ and $w^{t+\Delta t}$ are the vertical velocities at the current and next timestep, and $\Delta t$ the length of the timestep. The spatial extent and value of the damping coefficient ($R_{w}$) is then determined by the equation \begin{multline} R_{w} = \begin{cases} {\rm C} \sin^2\left( 0.5\pi(\eta-\eta_{\rm s}) \left( \frac{1.0}{1.0-\eta_{\rm s}}\right) \right) \mbox{,}\, &\eta \geq \eta_{\rm s}\\ 0 \mbox{,}\, & \eta < \eta_{\rm s},\\ \end{cases} \label{sponge} \end{multline} where, given the absence of orography, $\eta=z / H$ (i.e. non--dimensional height), $\eta_{\rm s}$ is the start height for the top level damping (set to $\eta_{\rm s}=0.75$) and $C$ is a coefficient. The value of $C$ is minimized for a given run. Usually, in Earth based studies one would place the sponge layer high above (or below) the region where the atmospheric flow is most active (i.e. the region of interest). However, for these test cases the top boundary intersects fast flowing features, and the sponge layer will potentially alter our solution there. While it may alter the solution this is more desirable than reflecting vertically propagating waves, artificially, back into the domain. The values assigned to the sponge layer are stated in Section \ref{hd209458b}. It is important to note that the damping coefficient $C$ represents the maximum damping present at the top boundary. Equation \ref{sponge} reduces the damping $\propto \sin^2$ as we move down from the upper boundary, meaning the practical damping felt by the vertical velocities reduces significantly from $C$. \subsection{Vertical coordinate and model comparison} \label{vert} In contrast to most other GCMs applied to hot Jupiters, which use $\sigma$ or pressure as the vertical coordinate, the UM uses geometric height coordinates. Ostensibly the choice of vertical coordinates should not alter the solution to a given equation set. However, due to large horizontal gradients in pressure, expected in the lower pressure regions of hot Jupiter atmospheres, surfaces of constant height do not align with surfaces of constant pressure (isobars). Therefore, to compare our model to a pressure--based model we must overcome three problems, namely, matching the boundary conditions and model domain, matching the vertical resolutions and comparing the results consistently. Generally, for both height--based and pressure--based models the inner boundary is at a set geopotential, and therefore (given the canonical assumption of spherical geopotentials) a fixed radial position, $r$. In general, as the inner boundary is deeper in the atmosphere pressure will not change significantly with time or horizontal position. Therefore, practically, if we set the pressure on our inner boundary to the value used in the pressure--based model our inner boundary conditions will be similar. However, for the upper boundary we use a constant height surface and pressure--based models use a constant pressure surface. The strong contrast in temperatures expected between the day and night side of hot Jupiters leads, in the upper atmosphere where the radiative timescale is short, to a significant gradient in pressure at a given height. At a given height the atmosphere will be hotter with higher pressure on the day side and cooler with lower pressures on the night side. If we are to completely capture the domain of a pressure--based code, we must set the position of our upper boundary so as to capture the minimum required pressure on the day side, and this height surface will sample lower pressures as it moves to the night side. This effectively means that we include an extra region of the atmosphere, over the domain modeled by a pressure--based code, being the region of the night side atmosphere at pressures lower than the minimum sampled by the pressure--based code (and by the height of our boundary on the higher pressure day side). The pressures, temperatures and densities of this material should, however, be small and therefore its dynamical effect be negligible (i.e. its angular momentum and kinetic energy contribution), as is shown by the agreement of our results with those from a pressure--based code (see Sections \ref{shj_results} and \ref{hd209458b_results}). However, we do have to alter the formulation of the radiative--equilibrium temperature--pressure ($T-p$) profiles in this region for stability (see Section \ref{hd209458b}), from that presented in \citet{heng_2011}. Additionally, as the pressure at a given height in an atmosphere will fluctuate in time it is impossible to exactly match the distribution of levels in a pressure--based models with one, such as the UM, based on geometric height. To provide a mapping between height and approximate pressure (or more specifically $\sigma$), for the SHJ test case we have completed a simulation using a uniform distribution of levels with an upper boundary high enough to capture the lowest required pressure. We then zonally and temporally--average the pressure structure. This allows us to distribute levels in height so as to sample $\sigma$ evenly, however, as the pressure will fluctuate we have increased the number of vertical levels (compared to the literature cases) to compensate. For HD 209458b, we have used uniform (in height) levels but, again, have increased the number relative to \citet{heng_2011} to compensate. We have altered our vertical resolution and level distributions and show in Section \ref{hd209458b_results} that it has a negligible effect on the results of the HD 209458b test case. Finally, to aid comparison of our results with literature pressure (or $\sigma$)--based models, we have interpolated the prognostic variables onto a pressure grid at each output. Horizontal averaging has then been performed along isobaric surfaces and the plots are presented with $\sigma$ or pressure as the vertical coordinate. \subsection{Diffusion, dissipation and artificial viscosity} \label{diffusion} In physical flows eddies and turbulence can cause cascades of kinetic energy from large--scale flows to smaller scales. At the smallest scales the kinetic energy is converted to thermal energy, heating the gas, due to the molecular viscosity of the gas. The resolutions possible with current models of planetary atmospheres (and other astrophysical models) do not reach the scales associated with molecular viscosity, and so a numerical scheme is required to mimic this dissipative process, as previously explained by many authors \citep{cooper_2005,cho_2008,menou_2009,heng_2011, bending_2013}. Some effective dissipation is provided by ``numerical viscosity'' inherent to the computational scheme itself. However, explicit schemes are included in different codes (both astrophysical and meteorological) to varying levels of accuracy or sophistication, and using differing nomenclature. Many astrophysical codes include an ``artificial viscosity'', where the controlling parameter can be altered to set the level of eddy or turbulent dissipation. Correctly formulated, an artificial viscosity includes the conversion of kinetic energy to heat via terms appearing in the momentum and thermal energy equation. For GCMs, and in meteorology, the term ``dissipation'' represents a similar scheme where losses of kinetic energy are accounted for in the thermal energy equation. Another scheme also regularly used in GCMs, is termed ``diffusion'', in this case a similar approach is used to remove kinetic energy, but this is not accounted for in the thermal energy equation. Such diffusion can be viewed as a numerical tool to remove grid scale noise. Although the operational\footnote{The version used by the UK Met Office for weather and climate prediction will use ENDGame from early 2014.} version of the ENDGame dynamical core includes no explicit diffusion, in our case, as with many other GCMs, we have incorporated a diffusion scheme. Whichever scheme is used the loss of kinetic energy can affect the characteristic flow and the maximum velocities achieved \citep{heng_2011,li_2010b}. It is possible to use known flows, such as in the boundary layer on Earth, to tune the form of this dissipation but this is not possible for hot Jupiters \citep[see discussion in][]{li_2010b}. Therefore, we do not ``tune'' our diffusion scheme to achieve a required wind speed, but for each of our test cases keep the diffusion constant for all simulations. Essentially, diffusion is used to provide numerical stability, although it will affect the results. Therefore, as with all other studies, the magnitude of our wind velocities are not robust predictions of the flow on a given hot Jupiter, rather the relative flows and patterns are the features to be interpreted. The scalar form of the diffusion operator $\textbf{D}(X)$ (which operates along $\eta$, or height as we have no orography, layers), is given by: \begin{multline} \textbf{D}(X)=\left( \frac{1}{r^2\cos\phi}\frac{\partial \eta}{\partial r}\right) \\ \left\{ \frac{\partial}{\partial \lambda}\left[ \frac{K_{\lambda}}{\cos\phi}\frac{\partial r}{\partial \eta}\frac{\partial }{\partial \lambda}\left( X \right) \right]+\frac{\partial}{\partial \phi} \left[ K_{\phi}\cos\phi \frac{\partial r}{\partial \eta}\frac{\partial}{\partial \phi} \left(X \right) \right] \right\},\\ \label{diffusion_eqn} \end{multline} where $X$ is the quantity to be diffused and $K_\lambda$ is given by \begin{equation} \frac{K_\lambda}{\cos\phi}=Kr^2\Delta\lambda^2\frac{\sin^2\left( \frac{\pi}{2}\cos\phi_{p} \right)}{\sin^2\left( \frac{\pi}{2}\cos\phi \right)} \end{equation} where $\phi_p=(\pi / 2)- (\Delta\Phi / 2)$ is the latitude of the row closest to the pole and $K_\phi=K_\lambda\left(\phi=0\right)(\Delta\phi^2 / \Delta\lambda^2)$. The value of $K$ is stated for each simulation in Table \ref{par_par}. In practice, as a further approximation, the diffusion operator is applied separately to each component of the vector field, as shown in Equations \ref{full_set} in Section \ref{equations_solved}. The construction of the diffusion operator allows the damping of the same physical scales as one approaches the equator (in practice this means that there is very little diffusion applied away from the polar regions and that small scale waves that could accumulate in the polar regions are removed) and also allows for variable resolution. Usually, polar filtering is achieved by applying multiple passes of an operator similar to that in Equation \ref{diffusion_eqn} from $\sim\pm 85^{\circ}$ to the poles, damping only in the zonal direction (as this is the scale which decreases towards the poles). In contrast, our diffusion operator is applied once across the entire globe and in both the zonal and meridional direction. We do not require an explicit polar filter, as used in other GCMs or previous versions of the UM. This is due to the changes in numerical scheme and the fact that our diffusion scheme will apply some damping, although significantly reduced, as would result from application of a polar filter. The diffusion is applied directly to the $u$, $v$ and $\theta$ fields for the SHJ test case \citep[][apply diffusion to relative vorticity and temperature using a $\sigma$ vertical coordinate]{menou_2009}. Whereas for the HD 209458b test case it is only applied to the $u$ and $v$ fields \citep[][apply diffusion to the $u$, $v$ and $T$ fields, again using a $\sigma$ vertical coordinate]{heng_2011}. One would ideally prefer to apply diffusion to the potential temperature for the HD 209458b test case, to match more closely the diffusion scheme of \citet{heng_2011}. However, firstly our results show some divergence from the results of \citet{heng_2011} when also applying diffusion to $\theta$, as shown and discussed in Section \ref{hd209458b_discuss}. Secondly, it is not actually clear that diffusing potential temperature along constant height surfaces (our scheme) is analogous to diffusing temperature along constant pressure surfaces \citep[scheme of][]{heng_2011}. We postpone a more complete discussion of this effect for a later work (Mayne et al, in preparation), and simply note here that the choice of diffusion scheme, target fields, and its interaction with the choice of vertical coordinate can potentially alter the results. \subsection{Radiative transfer} \label{radiative_transfer} Radiative transfer, for these tests, has been parameterised using a Newtonian cooling scheme \citep[as used for many models of hot Jupiters, e.g.][]{cooper_2005,menou_2009,heng_2011}. The heating rate in the thermodynamic equation stated in Section \ref{equations_solved} is, \begin{equation} Q=Q_{\rm Newton}=-\Pi\left(\frac{\theta-\theta_{\rm eq}}{\tau_{\rm rad}}\right), \label{newt_cool} \end{equation} where $\tau_{\rm rad}$ the characteristic radiative or relaxation timescale. $\theta_{\rm eq}$ is the equilibrium potential temperature and is derived from the equilibrium temperature ($T_{\rm eq}$) profile using \begin{equation} \theta_{\rm eq}^{i}=\frac{T_{\rm eq}}{\Pi^{i}}, \end{equation} where superscript $i$ denotes the current timestep. Practically, the potential temperature is adjusted explicitly within the semi--Lagrangian scheme using \begin{equation} \theta^{i+1}=\theta_{D}^{i}-\frac{\Delta t}{\tau_{\rm rad}}\left(\theta^{i}-\theta_{\rm eq}^{i}\right)_{D}, \end{equation} where the superscript $i+1$ denotes the next timestep and $\Delta t$ is the length of the timestep. The subscript $D$ denotes a quantity at the departure point of the fluid element \citep[see explanation in Section \ref{numerics} and][for a full discussion]{wood_2013} \footnote{From the equations in this section one can recover, $Q_{\rm Newton}=\frac{T_{\rm eq}-T}{\tau_{\rm rad}}$ and $T^{i+1}=T^{i}-\frac{\Delta t}{\tau_{\rm rad}}(T^{i}-T_{\rm eq})$ as shown, for example in \citet{heng_2011}.} We are currently completing the development of a two--stream, correlated--k radiative transfer scheme. This will allow us to run more realistic models and avoid the problems associated with simplified radiative transfer schemes \citep[for instance the omission of thermal re--emission of heated gas, and the separation between the temperature adjustment and heat capacity of a given atmospheric fluid elements, see][for discussion]{showman_2009}. \section{Test cases} \label{test_cases} We have performed simulations of a generic SHJ \citep[that prescribed in][]{menou_2009} and HD 209458b \citep[as prescribed in][]{heng_2011}. Table \ref{par_par} lists the general parameters common for all of the SHJ or HD 209458b simulations. \begin{table*} \caption{Value of the general (i.e. set for a given test case) parameters for the test cases.} \label{par_par} \centering \begin{tabular}{lcc} \hline\hline Quantity&SHJ&HD 209458b\\ \hline Horizontal resolution&\multicolumn{2}{c}{$144_\lambda$, $90_\phi$}\\ Standard vertical resolution&32&66\\ Timestep (s)&\multicolumn{2}{c}{1200}\\ Run length (Earth days)&\multicolumn{2}{c}{1200}\\ Sampling rate, $\Delta t_{\rm s}$ (days)&\multicolumn{2}{c}{10}\\ Initial inner boundary pressure, $p_{\rm s}$ (Pa)&$1\times 10^5$&$220\times 10^5$\\ \multirow{2}{*}{Radiative timescale, $\tau_{\rm rad}$ (s)}&\multirow{2}{*}{$\frac{\pi}{\Omega_{\rm p}}\sim1.5\times 10^5$}&\citet{iro_2005}, where $p< 10\times10^5$ Pa ($\sim1\times 10^{3-8}$)\\ &&$\infty$, where $p\geq 10\times10^5$ Pa\\ Initial temperature profile&Isothermal $1800$ K&$\frac{T_{\rm day}+T_{\rm night}}{2}$\\ Equator--pole temperature difference, $\Delta T_{\rm EP}$ (K)&300&\multicolumn{1}{l}{\rdelim\{{5}{0mm}[Modified \citet{iro_2005} profiles]}\\ Equatorial surface temperature, $T_{\rm surf}$ (K)&1600&\\ Lapse rate, $\Gamma_{\rm trop}$ (Km$^{-1}$)&$2\times 10^{-4}$&\\ Location of stratosphere ($z_{\rm stra}$, m \& $\sigma_{\rm stra}$)&$2\times10^6$, $\sim0.12$&\\ Tropopause temperature increment , $\Delta T_{\rm strat}$ (K)&10&\\ Rotation rate, $\Omega$ (s$^{-1}$)&$2.1\times 10^{-5}$&$2.06\times 10^{-5}$\\ Radius, $R_{\rm p}$ (m)&$10^8$&$9.44\times10^7$\\ Radius to outer boundary (m)&$3.29698\times 10^6$&$1.1\times 10^7$\\ Surface gravity, $g_{\rm p}$ (ms$^{-2}$)&8&9.42\\ Specific heat capacity (constant pressure), $c_{\rm p}$ (Jkg$^{-1}$K$^{-1}$)&13226.5&14308.4\\ Ideal gas constant, $R$ (Jkg$^{-1}$K$^{-1}$)$^{(1)}$&3779&4593\\ $K$, diffusion coefficient&0.495&0.158\\ \hline \end{tabular} \tablebib{(1) The $R$ value is varied between simulations to attempt to represent differences in the molecular weight of the modeled portion of the atmosphere.} \end{table*} For each simulation we have followed \citet{held_1994} and \citet{heng_2011} and run the simulations for 1200 days (here, and throughout this work, `days' refers to Earth days). The first 200 days are then discarded to allow for initial transients and `spin--up', which is sufficient to span several relaxation times for the entire atmosphere in the SHJ case and for the upper atmosphere down to a pressure of $\sim 10^5$ Pa (or a few bar) for HD 209458b \citep[using the radiative timescale of][]{iro_2005}. For the HD 209458b test case 1200 days is only sufficient to span $\sim 1$ radiative relaxation time throughout the radiative zone. Additionally, as the HD 209458b test case also includes a radiatively inactive region a significantly longer time is required to reach a statistical steady state \citep[for example][found after 5000 days the atmosphere had reached a steady state down to $3\times 10^5$ Pa or $\sim$3 bar]{cooper_2005}. The issue of whether the simulation has reached a statistically steady state will be discussed in more detail in Section \ref{hd209458b_discuss}. The solution from 200 to 1200 days is then used to create zonally and temporally averaged temperature and zonal wind plots, which we term `zonal mean' plots \citep[in a similar vein to][]{heng_2011}. As discussed in section \ref{vert}, to aid comparison with previous works we present plots using $\sigma$ (SHJ) and $\log(p)$ (HD 209458b) as our vertical coordinate, which we have created by interpolating the values from the geometric grids onto the isobaric surface required. The plots (throughout this work, for example Figure \ref{shj_compare}) feature contour lines (solid for positive and dashed for negative) that have been chosen, where applicable, to match those in \citet{heng_2011}. These are complemented by colour scales, where a greater number of divisions (than the line contours) are used to aid qualitative interpretation of the data\footnote{The values of the labels for the colour scales have been rounded to integer values. Additionally, the total range used for the colour scale is larger than the range of the data.}. The colour scales chosen have mostly been selected to match standard colour schemes in meteorology (i.e. blue--red for temperature). For wind plots we have adopted a blue--white--red colour scale where blue is retrograde or downdraft, i.e. negative wind, red is prograde or updraft, i.e. positive wind and white is positioned at zero\footnote{The splitting of the colour scales means that the colour scales need not be symmetric about zero.}. \subsection{Shallow--Hot Jupiter} \label{shj} \subsubsection{Test case setup} \label{shj_setup} The SHJ test is that prescribed by \citet{menou_2009}, a thin layer of a hypothetical tidally locked Jovian planet down to a depth of $1\times 10^5$ Pa or 1 bar. The equilibrium temperature profile is, \begin{equation} T_{\rm eq}=T_{\rm vert}+\beta_{\rm trop}\Delta T_{\rm EP}\cos(\lambda - 180^{\circ})\cos(\phi),\\ \end{equation} where $T_{\rm vert}$ is given by, \begin{align} T_{\rm vert}= \begin{cases} T_{\rm surf}-\Gamma_{\rm trop}(z_{\rm stra}+\frac{z-z_{\rm stra}}{2})\\ +\left( \left[ \frac{\Gamma_{\rm trop}(z-z_{\rm stra})}{2}\right] ^2+\Delta T_{\rm strat}^2\right) ^{\frac{1}{2}}\mbox{,}\, &z \leq z_{\rm stra}\mbox{,} \nonumber\\ T_{\rm surf}-\Gamma_{\rm trop}z_{\rm stra}+\Delta T_{\rm strat}\mbox{,}\, &z > z_{\rm stra} \mbox{.} \end{cases}\\ \label{T_vert} \end{align} and $\beta_{\rm trop}$ is defined as \begin{align} \beta_{\rm trop}&=\begin{cases} \sin\frac{\pi(\sigma-\sigma_{\rm stra})}{2(1-\sigma_{\rm stra})}\mbox{,}\, &z \leq z_{\rm stra} \,\mbox{ or }\, \sigma \geq\sigma_{\rm stra} \mbox{,} \nonumber\\ 0\mbox{,}\, &z > z_{\rm stra} \, \mbox{ or }\, \sigma < \sigma_{\rm stra} \mbox{.} \end{cases}\\ \end{align} The values for the parameters featured in these equations are presented in Table \ref{par_par}. The radiative relaxation timescale throughout the entire atmosphere is set to $\tau_{\rm rad}=\pi / \Omega_{\rm p}\sim 1.731$ days. We have run this test case using the ``Full'', ``Deep'' and ``Shallow'' equation sets (see Table \ref{model_names} for explanation), with the rest of the setup the same for each simulation. The number of vertical levels is 32 and the level top is placed at $3.29698\times 10^6$ m, no sponge layer was necessary and the diffusion has been applied to $u$, $v$ and $\theta$\footnote{We have performed a simulation incorporating a sponge and found no significant differences in the results from those presented in Figure \ref{shj_compare}.}. We started the atmosphere initially at rest and in vertical hydrostatic equilibrium using an isothermal temperature profile set at $1800$ K as used by \citet{heng_2011}. \subsubsection{Results} \label{shj_results} The resulting flow and temperature of the ``Shallow'' SHJ test case at the $\sigma=0.675$ surface after 346 days, as well as the zonal mean plots are shown alongside the figures from \citet{heng_2011} (using their finite--grid model) in Figure \ref{shj_compare}. We present the instantaneous temperature field at $\sigma=0.675$ instead of the quoted value of 0.7 in \citet{heng_2011} as this quoted value does not represent the actual value of the surface, but the half--level just above it (i.e. at lower sigma and greater height). Therefore, the real $\sigma$ value is half the vertical resolution below the quoted $\sigma$ value \citep[see][for a full discussion of this in regards to Earth--like tests]{mayne_2013}. Figure \ref{shj_compare} shows that, qualitatively, we match the broad characteristics of the flow. Figure \ref{shj_compare_deep} then shows the same plots but for the ``Full'' case (the ``Deep'' case is omitted as it is virtually identical to the ``Full''). Figure \ref{shj_compare_deep} shows an atmospheric structure broadly consistent with both the ``Shallow'' case and that of \citet{heng_2011}. As the atmosphere of the SHJ is only $1\times 10^5$ Pa or 1 bar in extent its height is $\sim4\times10^6$m, and as the planetary radius is $R_{\rm p}=10^8$ (see Table \ref{par_par}), it is unsurprising that no difference is found when relaxing the `shallow--atmosphere' approximation (see Table \ref{model_names}). Indeed the resulting flow is very similar in all cases. Some slight differences are present which will be discussed briefly in Section \ref{shj_discuss}, but for now we move on to a more physically interesting test case. \begin{figure*} \centering \includegraphics[width=8.5cm,angle=0,origin=c]{./Fig/heng_t_snap_shj.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/SHJ/shallow_temp_snap.png} \includegraphics[width=8.5cm,angle=0,origin=c]{./Fig/heng_t_shj.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/SHJ/shallow_temp_avg.png} \includegraphics[width=8.5cm,angle=0,origin=c]{./Fig/heng_u_shj.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/SHJ/shallow_uvel_avg.png} \caption{Figure showing the solutions to the SHJ test case. \textit{Left panels} are figures reproduced from \citet{heng_2011} using the finite--difference model (reproduced by permission of Oxford University Press), and the \textit{right panels} are results from this work for the ``Shallow'' case (see Table \ref{model_names} for explanation). The \textit{top row} shows the temperature field at $\sigma=0.675$ and 346 days. The \textit{middle} and \textit{bottom rows} show the zonal mean plots for temperature and wind respectively (i.e. zonally and temporally, from 200-1200 days, averaged).} \label{shj_compare} \end{figure*} \begin{figure} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/SHJ/deep_temp_snap.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/SHJ/deep_temp_avg.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/SHJ/deep_uvel_avg.png} \caption{Figure matching those described in Figure \ref{shj_compare} but for the ``Full'' case (see Table \ref{model_names} for explanation).} \label{shj_compare_deep} \end{figure} \subsection{HD 209458b} \label{hd209458b} \subsubsection{Test case setup} \label{hd209458b_setup} The test case for HD 209458b is a slightly adjusted version of that prescribed in \citet{heng_2011} \citep[similar to that described in][]{rauscher_2010}, where the temperature and relaxation profiles are taken from the radiative equilibrium models of \citet{iro_2005}. The domain encompasses a radiatively inactive region from $2.2\times 10^7$ to $1\times 10^6$ Pa (or 220 to 10 bar) \citep[where $\tau_{\rm rad}=\infty$, termed `inactive' in][]{heng_2011} with a radiative zone above this. As discussed in Section \ref{vert} due to the horizontal gradients in pressure in the upper atmosphere, as we are using a height based approach and matching a test case performed in pressure coordinates we are including, necessarily, an extra section of computational domain i.e. the low pressure night side region. We found for our non--hydrostatic code the model was extremely unstable on the night side in this very cool low pressure region, leading to exponential growth of vertical velocities under small perturbations. Additionally, we found that the discontinuities in temperature across the $1\times 10^6$ Pa (or 10 bar) boundary found in the profile described in \citet{heng_2011} also led to instability \citep[as discussed in][]{rauscher_2010}. Therefore, we have slightly adjusted the profiles of \citet{heng_2011}. The most significant change, a modest heating of around 150 K, is performed in the region above $10^{-3}$ bar. This region is not included in the model of \citet{heng_2011}, as their upper boundary is placed at this pressure. The altered temperature profiles are shown in Equations \ref{hd209_force_1} and \ref{hd209_force_2} and plotted in Figure \ref{hd209_tprof} (with the radiative and radiatively inactive regions also indicated). \begin{equation} T_{\rm night}=\begin{cases} T_{\rm night}^{\prime}\rvert^{p_{\rm high}}+100 {\rm K}\left(1.0-e^{-(\log(p)-\log(p_{\rm high}))}\right) \mbox{,}\,& p\geq p_{\rm high}\\ {\rm MAX}\left( T_{\rm night}^{\prime}\rvert^{p_{\rm low}}\times e^{0.10(\log(p)-\log(p_{\rm low}))}, 250\right)\mbox{,}\,& p< p_{\rm low}\\ T_{\rm night}^{\prime}\rvert^p\, &\mbox{otherwise} \end{cases} \label{hd209_force_1} \end{equation} \begin{equation} T_{\rm day}=\begin{cases} T_{\rm day}^{\prime}\rvert^{p_{\rm high}}-120.0 {\rm K}\left(1.0-e^{-(\log(p)-\log(p_{\rm high}))}\right) \mbox{,}\,& p\geq p_{\rm high}\\ {\rm MAX}\left( T_{\rm day}^{\prime}\rvert^{p_{\rm low}}\times e^{0.015(\log(p)-\log(p_{\rm low}))}, 1000\right)\mbox{,}\,& p< p_{\rm low}\\ T_{\rm day}^{\prime}\rvert^p\, &\mbox{otherwise} \end{cases} \label{hd209_force_2} \end{equation} $T_{\rm day}$ and $T_{\rm night}$ are the day and night side temperature profiles and $p$ is the pressure. $T_{\rm night}^\prime$ and $T_{\rm day}^\prime$ are the polynomial fits of \citet{heng_2011} to the day and night side profiles of \citet{iro_2005}, and $p_{\rm low}$ and $p_{\rm high}$ are $100$ and $1\times 10^6$ Pa respectively (or $1\times 10^{-3}$ and $10$ bar). \begin{figure} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90.0]{./Fig/profiles.png} \caption{Temperature--pressure profiles used for HD 209458b. The solid lines are from this work, and the dashed lines are the polynomial fits of \citet{heng_2011} to the models of \citet{iro_2005}. The blue lines are the night side profiles, the red lines the day side profiles (i.e. $T_{\rm night}$ and $T_{\rm day}$, respectively) and the green line is the initial profile. The horizontal dashed line demarks the radiatively inactive and radiative regions.} \label{hd209_tprof} \end{figure} The resulting profiles in Equations \ref{hd209_force_1} and \ref{hd209_force_2} are then combined to create a temperature map of the planet's atmosphere using, \begin{multline} T_{\rm eq}=\\ \begin{cases} \left[ T_{\rm night}^4+ (T_{\rm day}^4 - T_{\rm night}^4)\cos(\lambda-180^\circ)\cos\phi\right]^\frac{1}{4}\mbox{,}\,& 90^\circ\leq \lambda \leq 270^\circ\\ T_{\rm night}\mbox{,}\, &\mbox{otherwise.}\\ \end{cases} \label{hd209_temp} \end{multline} We have run this test case using the ``Full'', ``Deep'' and ``Shallow'' equations sets with the top boundary placed at $1.1\times 10^7$ m and use 66 vertical levels (distributed uniformly in height). For this test case we require a sponge layer and minimise this for each simulation, where $\eta_{\rm s}=0.75$ in all cases. $C$ is 0.20 for both the ``Deep'' and ``Shallow'' case but 0.15 for the ``Full'' case. The effect of both the sponge layer and the use of uniform vertical levels (as opposed to those sampling, for instance, uniform $\log(p)$) have been explored and are briefly discussed in \ref{hd209458b_discuss}. Each of the simulations has been initialised in hydrostatic equilibrium using a temperature profile midway between the day and night profiles (i.e. $(T_{\rm day}+T_{\rm night}) / 2$) and zero initial winds. As we are trying reproduce the results of a test case, we postpone a detailed exploration of the effect of varying initial conditions for later work (Mayne et al, in preparation). \subsubsection{Results} \label{hd209458b_results} In general our resulting large--scale, long--term flows and those of \citet{heng_2011} for HD 209458b are qualitatively very similar. In order to aid comparison Figure \ref{hd209_bench_heng} reproduces the results of \citet{heng_2011}. Figure \ref{hd209_bench_heng} shows snapshots of temperature and horizontal velocity for the same pressure levels (i.e. 213, 21$\,$6000, 4.69$\times 10^5$ and 21.9$\times 10^5$ Pa) as in \citet{heng_2011} at 1200 days as found using their spectral code\footnote{We do not compare to the finite--difference model as the full set of snapshots for this case are not presented in \cite{heng_2011}.}. Figure \ref{hd209_bench_heng} also shows the zonal mean plots for the finite difference model of \citet{heng_2011}. The same plots for our ``Shallow'' case are presented in Figure \ref{hd209_bench_shallow}. We note that \citet{heng_2011} uses the pressure unit of bar, whereas we use SI units, Pa (where $1$ bar is $1\times10^5$ Pa). \begin{figure*} \centering \includegraphics[width=8.5cm,angle=0.0,origin=c]{./Fig/hd209_heng_temp_mean.jpg} \includegraphics[width=8.5cm,angle=0.0,origin=c]{./Fig/hd209_heng_uvel_mean.jpg} \includegraphics[width=8.5cm,angle=0,origin=c]{./Fig/hd209_heng_213.jpg} \includegraphics[width=8.5cm,angle=0,origin=c]{./Fig/hd209_heng_21600.jpg} \includegraphics[width=8.5cm,angle=0.0,origin=c]{./Fig/hd209_heng_4_69e5.jpg} \includegraphics[width=8.5cm,angle=0.0,origin=c]{./Fig/hd209_heng_21_9e5.jpg} \caption{Figure showing results for the HD 209458b test case reproduced from \citet{heng_2011} (reproduced by permission of Oxford University Press). The \textit{top row} shows the zonal mean plots (i.e. zonally and temporally, from 200-1200 days, averaged, using bar as the unit of pressure) of temperature (\textit{left}) and zonal wind (\textit{right}). The \textit{middle and bottom rows} show the temperature (colour) and horizontal velocities (vectors) at pressures 213 (\textit{middle left}), 21$\,$600 (\textit{middle right}), 4.69$\times 10^5$ (\textit{bottom left}) and 21.9$\times 10^5$ Pa (\textit{bottom right}) after 1200 days.} \label{hd209_bench_heng} \end{figure*} \begin{figure*} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_temp_avg_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_uvel_avg_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_snap_213_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_snap_21600_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_snap_4_69e5_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_snap_21_9e5_1200.png} \caption{Figure matching those described in Figure \ref{hd209_bench_heng} but for our ``Shallow'' case. The zonal mean plots present pressure in Pa (SI unit, where $1$ bar=$1\times10^5$ Pa).} \label{hd209_bench_shallow} \end{figure*} Figures \ref{hd209_bench_deep} and \ref{hd209_bench_full} show the same plots as Figures \ref{hd209_bench_heng} and \ref{hd209_bench_shallow} but for the ``Deep'' and ``Full'' cases, respectively. Comparing the results of \citet{heng_2011} reproduced in Figure \ref{hd209_bench_heng} with our own results shown in Figures \ref{hd209_bench_shallow}, \ref{hd209_bench_deep} and \ref{hd209_bench_full}, shows good, qualitative, agreement. In all cases we produce a wide, in latitude, prograde equatorial jet extending throughout the upper atmosphere from about $5\times 10^5$ Pa ($5$ bar) to $100$ Pa (or 1 mbar), flanked by retrograde winds. The temperature distribution also matches across the radiative zone. The jet does sharpen slightly, in latitude, and move to higher altitudes and lower pressures, as well as reducing in magnitude, when moving to the more sophisticated equation sets (i.e. ``Shallow'' to ``Full''). The instantaneous slices through the atmosphere at 213 and 21$\,$600 Pa are also consistent across the figures presented. The 213 and 21$\,$600 Pa isobaric surfaces exhibit diverging flow at the lower pressures and the development of a circumplanetary jet, with an associated shift in the temperature distribution at the higher pressure of the two surfaces. The temperature distributions also show little variation ($\Delta T\lesssim 150$K) across all simulations, which is unsurprising given the short radiative timescale at these pressures. At the higher pressure of 21.9$\times 10^5$ Pa the flow, morphologically, is still very similar, however the flow of \citet{heng_2011} appears less coherent. Additionally, slightly larger differences in temperature (than those found at the lower pressures) across the simulations appear, for the deepest isobaric surface. The pole, at depths, in the radiatively inactive region appears to become warmer as we move to the more complete (i.e. ``Deep'' and ``Full'') dynamical equations. The isobaric slice which shows the most difference between simulations is at 4.69$\times 10^5$ Pa. Here the flow morphology of the instantaneous field at 1200 days is quite different across the simulations, as is the associated temperature structure. Both the ``Deep'' and ``Full'' cases show a counter rotating, or westward moving flow at all latitudes. There is also a shift in the temperature distribution, with the regions of lowest temperature shifted to lower longitudes (i.e. westward). Despite the differences in the instantaneous slices at 4.69$\times 10^5$ Pa, the overall flow morphology is qualitatively very similar through each of simulations. Moreover, the time averaged flow and temperature structure, for all simulations, shows very little difference, despite the differences in numerical scheme, initial conditions and the equations solved. \begin{figure*} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_temp_avg_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_uvel_avg_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_snap_213_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_snap_21600_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_snap_4_69e5_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_snap_21_9e5_1200.png} \caption{Figure matching those described in Figure \ref{hd209_bench_heng} but for our ``Deep'' case. The zonal mean plots present pressure in Pa (SI unit, where $1$ bar=$1\times10^5$ Pa).} \label{hd209_bench_deep} \end{figure*} \begin{figure*} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_temp_avg_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_uvel_avg_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_snap_213_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_snap_21600_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_snap_4_69e5_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_snap_21_9e5_1200.png} \caption{Figure matching those described in Figure \ref{hd209_bench_heng} but for our ``Full'' case. The zonal mean plots present pressure in Pa (SI unit, where $1$ bar=$1\times10^5$ Pa).} \label{hd209_bench_full} \end{figure*} In Figure \ref{hd209_bench_variants} we present the results from a subset of the simulations we have run demonstrating the relative invariance of the derived flow structure for this test case, over 1200 days. Here we term the standard simulations as those presented in Figures \ref{hd209_bench_shallow}, \ref{hd209_bench_deep} and \ref{hd209_bench_full}. We have then run a set of simulations where we have changed individual parameters or settings to explore their effect, using each of the ``Shallow'' and ``Full'' equation sets. Here we present only a subset in order to sample the whole `parameter space' with as few figures as possible. As discussed in Section \ref{diffusion} we apply diffusion to the $u$ and $v$ fields only for this test case, in order to simply reproduce a more consistent result with that of \citet{heng_2011}. The \textit{top left panel} for Figure \ref{hd209_bench_variants} shows the results for a simulations with exactly the same setup as the ``Shallow'' case shown in the \textit{top right} panel of Figure \ref{hd209_bench_shallow}, but with diffusion additionally applied to the $\theta$ field. The jet structure still persists but has shifted to higher pressures, sharpened in latitude and diminished, slightly diverging from the results of \citet{heng_2011} (as discussed in Section \ref{diffusion}). This change is likely due to the effect of diffusion of the potential temperature on the baroclinically unstable regions flanking the equatorial jet. The details and changes in the underlying mechanism which `pumps' the jet will be explored in a future publication (Mayne et al, in preparation). Despite the differences, the flanking retrograde jets are still present and the relative prograde to retrograde motions are similar to the previous simulations. As mentioned in Section \ref{vert} we also adopt uniformly distributed (in geometric height) vertical levels for our standard HD 209458b simulations, as opposed to those uniform in $\log(p)$ \citep[as adopted by][]{heng_2011}. The \textit{top right panel} of Figure \ref{hd209_bench_variants} shows the resulting flow for a simulations matching the ``Full'' case presented in the \textit{top right panel} of Figure \ref{hd209_bench_full} but with only the vertical level distribution altered. The non-uniform level distribution used has been chosen to sample the \textit{local minimum} atmospheric scaleheight. At each height, starting at the inner boundary, the smallest scaleheight (usually on the cooler night side) was found and three levels were placed within this scaleheight. The process was repeated till the height domain of the atmosphere ($1.1\times 10^7$ m) was reached (and resulted in a total of 78 vertical levels, compared to 66 for the standard simulations). Again, a similar flow morphology is found with a prograde jet flanked by retrograde jets, with only a modest sharpening of the jet apparent when compared to the standard ``Full'' case. Finally, we have also, as detailed in Section \ref{boundary} included a sponge layer in our upper boundary condition. Therefore, to explore the effect of this damping we have run two further simulations. The \textit{bottom right panel} of Figure \ref{hd209_bench_variants} shows a simulations where only the upper boundary has been altered from the standard ``Full'' case presented in Figure \ref{hd209_bench_full}, and is placed at $1.25\times 10^7$ m above the inner boundary (using 80 vertical levels to retain a similar vertical resolution). As we increase the size of the domain, our upper boundary moves to lower pressures, however, in Figure \ref{hd209_bench_variants} we only present the vertical section of the domain matching that encompassed by the standard ``Full'' case to aid comparison\footnote{The flow above this, at lower pressures, is just an obvious extension of the retrograde flow shown at the top of the figure.}. For this simulation the damping layer only becomes non--negligible for pressure lower than $<100$ Pa, i.e. above the plotted region. As before, the flow does not diverge significantly from what one would expect of the simulations both of \citet{heng_2011} and others in this work. Secondly, in the \textit{bottom left panel} we present a simulation matching the standard ``Shallow'' case, presented in \textit{top right panel} of Figure \ref{hd209_bench_shallow}, except the upper boundary has been placed at $6.7 \times10^6$ m above the inner boundary (using 40 vertical levels, again to retain a similar vertical resolution), and does not include a damping layer in the upper boundary condition. Once more, the flow is approximately what one might expect from inspection of our standard case and those of \citet{heng_2011}. The results presented in the \textit{bottom panels} of Figure \ref{hd209_bench_variants} indicate that the damping layer is not significantly altering our results besides its inclusion being physically preferable (by preventing reflection of gravity waves back into the domain at the upper boundary). Figure \ref{hd209_bench_variants} represents only a subset of the simulations we have run to explore the sensitivity to the parameters and numerical choices. However, all of the simulations not presented here show a similar qualitative flow structure, i.e. a prograde equatorial jet flanked by retrograde winds, over the 1200 day test period. \begin{figure*} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_uvel_avg_200_1200_diff_theta.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_uvel_avg_200_1200_sh.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_uvel_avg_200_1200_low.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_uvel_avg_200_1200_high.png} \caption{Figure showing the insensitivity of the zonally and temporally averaged zonal wind (ms$^{-1}$) to the different modelling choices. The simulations in the \textit{left panels} use the ``Shallow'' and the \textit{right panels} the ``Full'' equation set. The \textit{top left panel} shows a simulations where diffusion is applied to $\theta$ in addition to $u$ and $v$. The \textit{top right panel} shows a simulations with non--uniform vertical level placement to optimise sampling of the \textit{local} scaleheight. The \textit{bottom left panel} shows the results when the atmospheric height is decreased from $H=1.1\times 10^7$ m to $H=6.7\times 10^6$ m, and the \textit{bottom right panel} when it is increased to $1.25\times 10^7$ m (although only the overlapping pressure domain of these simulations with that of the models in \citet{heng_2011}, shown in Figure \ref{hd209_bench_heng}, is displayed to aid comparison).} \label{hd209_bench_variants} \end{figure*} The key conclusion one can draw from these results is that the general atmospheric structure is relatively invariant over 1200 days under a range of model choices. Therefore, the resulting zonal mean diagnostics plots for the HD 209458b test case \citep[as presented in][]{heng_2011} are qualitatively very similar for all models. When comparing our ``Shallow'' case with the primitive model of \citet{heng_2011} the agreement suggests that, for this test, the relaxation of the hydrostatic approximation and change in vertical coordinate (from $\sigma$ to height) is unimportant. Furthermore, although deviation is present in the snapshots and detail of the jet structures, further relaxation of the `shallow--fluid' and `traditional' approximations does not significantly alter the results (our ``Deep'' case). Finally, the additional relaxation of the constant gravity assumption (as represented by our ``Full'' case) also does not cause the long--term, large--scale atmospheric structure to change dramatically (i.e. the zonal mean plots). We have also shown that the results are relatively invariant to our numerical choices associated with diffusion, vertical resolution (and level placement) and the upper boundary sponge or damping layer. However, again slight differences in the detail of the flow structures are apparent. \section{Discussion} \label{discussion} Despite the general concordance of our results with literature results, and across our different model types, some differences are apparent which we briefly discuss in this section. The zonal and temporal averaging involved in the zonal mean plots is intended to provide a robust way to compare the long--term and large--scale structure of the model atmospheres. Therefore, by design these plots are relatively insensitive to the more detailed differences in the atmospheres. \subsection{Shallow--Hot Jupiter} \label{shj_discuss} As discussed in Section \ref{test_cases} we have placed our vertical levels for the SHJ test case at positions emulating the $\sigma$ levels described in \citet{heng_2011}. This process involved running a SHJ test case, with uniformly distributed vertical levels, to completion and zonally and temporally averaging the pressure structure to provide a mapping from height to pressure. During this process, the largest $\sigma$ value, i.e. the level closest to the inner boundary, leads to a very small (in vertical size) grid cell, which, even with a semi--implicit scheme, led to a numerical instability of the vertical velocity. Therefore, the lowest level was adjusted to create a larger (in vertical extent) bottom cell more numerically stable for a non--hydrostatic code. Although our results for the SHJ are qualitatively similar to those of \citet{heng_2011} some differences are apparent. Most notably, perhaps, is the fact that our jets (prograde or retrograde) do not intersect either the boundary. No sponge layer is incorporated in the upper boundary for this test, but the result is unchanged when it is. This slight discrepancy between our results and those of \citet{menou_2009} and \citet{heng_2011} is most likely caused by differences in domain or boundary conditions, as both boundaries intersect the flow features we are trying to capture. In this respect, i.e. likely dependency of the results on the domain and boundary conditions, the SHJ test is a poor benchmark. \subsection{HD 209458b} \label{hd209458b_discuss} As explained in Sections \ref{approx} and \ref{test_cases}, the prescribed test duration of 1200 days is only sufficient to span approximately one relaxation time for the deeper regions of the radiative zone. This, in addition to the fact it includes a radiatively inactive region which can only reach a relaxed or steady state through dynamical processes, suggests that 1200 days is insufficient for this case to reach a statistical steady state. Models based on the primitive equations have already shown that the deeper atmosphere will not reach a steady state in 1200 days. Both \citet{cooper_2005} and \citet{rauscher_2010} present evidence indicating that the atmosphere down to only $\sim 3\times 10^5$ Pa (or $\sim$3 bar) had relaxed in their models after 5000 and $\sim$600 days, respectively. \citet{rauscher_2010} additionally, explicitly show that the kinetic energy is still evolving in the deeper regions of their modeled atmosphere after 1200 days. Additionally, models from the literature which include a more complete dynamical description, have been run for much shorter times than 1200 days. For example \citet{dobbs_dixon_2010} run their simulations, which include the full dynamical equations, for only $\sim$ 100 days. As suggested by \citet{showman_2002} a downward flux of kinetic energy was found in models of HD 209458b by \citet{cooper_2005}, therefore energy is transported into the deeper radiatively inactive region. Energy is also injected by the compressional heating. As discussed in Section \ref{approx} if one compares the equation sets used in our different models, as presented in Table \ref{eqn_sets}, the terms affected as we move from a ``Shallow'' to a ``Deep'' and on to a ``Full'' equation set involve `exchange' between the components of momentum, and importantly the vertical and horizontal components. Moreover, relaxing the constant gravity assumption, in particular, acts to weaken the stratification of the atmosphere. Therefore, it is plausible that as one moves to a more complete dynamical description, one allows the transfer of energy and momentum between the upper radiative atmosphere with short relaxation time (see discussion in Section \ref{equations_solved}), with both the deeper longer timescale radiative and the even deeper radiatively inactive regions. A retrograde flow in the deep region of the atmosphere must arise through an equatorward meridional flow (by conservation of angular momentum) or by vertical transport of angular momentum by waves or eddies, and must be accompanied by a warming of the polar regions relative to the equator (by thermal wind balance), which itself can only arise through a meridional circulation with descent near the poles and ascent near the equator. Figure \ref{move_deep} shows the vertically and zonally averaged equator--to--pole temperature difference (in the sense $T_{\rm equator}-T_{\rm pole}$), and total kinetic energy, for the radiatively inactive region (i.e. $p\geq 1\times10^6$ Pa or 10 bar), for the HD 209458b test case and each equation set. Figure \ref{move_deep} shows evidence that the latitudinal temperature gradient in the deep atmosphere, and the kinetic energy, are approaching a steady state in the ``Shallow'' case. However, for both the ``Deep'' and ``Full'' cases the average latitudinal temperature gradient and total kinetic energy, are both still increasing by the end of the simulation. Additionally, Figure \ref{move_deep} shows that the average temperature difference between the equator and pole is larger in the ``Deep'' and ``Full'' cases than in the ``Shallow'' case, as is the total kinetic energy, in the radiatively inactive region. Figure \ref{move_deep}, therefore, gives a strong indication both that the more simplified equation sets poorly represent the dynamical evolution of the deep atmosphere, and that the more sophisticated cases require a longer time to reach a statistically steady state. The radiative timescale at the bottom of the radiative zone is $\tau_{\rm rad}\sim 500$ days, for HD 209458b. Therefore, given that below this the radiative timescale is infinite one would expect the timescale for relaxation of the radiatively inactive region to be $>>500$ days. The total elapsed time for the test cases performed in this work is 1200 days, suggesting it is unlikely the deep atmosphere will reach a relaxed state (an estimate supported by the data presented in Figure \ref{move_deep}). Given that the angular momentum, and kinetic energy budget of the atmosphere can potentially be dominated by the deepest regions of the atmosphere, and the relaxation time of the deeper layers is long (compared to model elapsed times), it suggests that partially relaxed solutions to the entire atmospheric flow may not be persistent equilibrium states. It has been shown, for models solving the HPEs, that the results of such simulations are invariant to initial conditions \citep[for discussion see][]{liu_2012}. However, as discussed in Section \ref{approx} the NHD equations include terms which act to exchange momentum between the vertical and horizontal motion. This exchange couples the shallow and deep atmosphere over long timescales meaning invariance to initial conditions cannot be proven until a statistical steady state \emph{throughout} the model domain is reached. The alteration of the flow as the deeper layers become activated may lead to the establishment of a different equilibrium state \citep[multiple equilibria are discussed in][]{liu_2012}, or it may move through a transient phase. Problematically, for models such as the UM, and more specifically the ENDGame dynamical core, which solve the NHD equations, the interaction with the deeper layers is extremely slow and therefore exploration of this element may require huge computer resources \citep[i.e. long integration times as mentioned in][]{showman_2008}. As a note of warning \citet{viallet_2008} demonstrate that for simulations of dwarf novae, where one side is strongly irradiated by the primary star, divergent flow is found, but no statistical steady state has been reached. \begin{figure} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/t_eq_all.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/ke_deep_all.png} \caption{Figure showing the zonally and vertically averaged equator--to--pole temperature difference (\textit{top panel}), and total kinetic energy (\textit{bottom panel}) for the radiatively inactive region (i.e. $p\geq 1\times10^6$ Pa or 10 bar), for the HD 209458b test case. The ``Shallow'', ``Deep'' and ``Full'' cases are shown as the solid, dotted and dashed lines, respectively.} \label{move_deep} \end{figure} For our simulations, the zonal mean plots all show a prograde equatorial jet, demonstrating insensitivity of the mechanism which produces this feature to the dynamical equations used, over 1200 days. However, given that the radiatively inactive region, for the ``Deep'' and ``Full'' cases is clearly still evolving, one might expect the lower pressure regions of the atmosphere to also demonstrate evolution. The zonal mean plots show that the prograde equatorial jet is thinned (in latitude), contracted in height and diminished in magnitude, in the ``Deep'' and ``Full'' cases when compared to the ``Shallow'' case. Looking in detail at the time evolution of the flow one finds a largely invariant structure throughout most of the atmosphere in the ``Shallow'' case, where exchange between the vertical and horizontal momentum is inhibited. However, both the ``Deep'' and ``Full'' cases exhibit a varying large--scale flow structures. Figure \ref{jet_decay} shows slices through the ``Full'' case at $1\times 10^5$ Pa (or 1 bar) at 100, 400, 800 and 1200 days (\textit{top left}, \textit{top right}, \textit{bottom left} and \textit{bottom right panels}, respectively). The slices in Figure \ref{jet_decay} show horizontal velocity (vectors) and the vertical velocity (colour scale). In this case (as is evident to a lesser degree in the ``Deep'' case) the large--scale flow is clearly still evolving. As the simulations runs the eastward jet, which is `spun--up' in the first tens of days, gradually degrades and westward flow encroaches across the equator. This effect is seen, to differing degrees, throughout the atmosphere and leads to the thinning and diminishing of the jet when performing a zonal average\footnote{We perform zonal averaging after our data have been transformed into pressure space to match the models of \citet{heng_2011}, and thereby avoid problems of comparison of quantities zonally averaged along different iso--surfaces \citep[as described in the appendix of][]{hardiman_2010}.}. It is intriguing, that the departure of our results from the results of \citet{heng_2011} is most apparent when the constant gravity approximation is relaxed. This change acts to weaken the stratification and thereby increase the efficiency of vertical transport via, for instance, gravity waves. Figure \ref{wvel_avg} shows the time averaged (from 200 to 1200 days) vertical velocities for the ``Shallow'', ``Deep'' and ``Full'' cases, as a function of pressure and either longitude (\textit{left panels}) or latitude (\textit{right panels}). In each case the field has been averaged in the horizontal dimension not plotted, i.e. if plotted as a function of latitude it has been zonally averaged. The meridional average is performed in a point--wise fashion, i.e. $\int\,vd\,\phi$ as opposed to $\int\,\cos\phi vd\,\phi$, to emphasise differences in vertical flow towards the polar regions. Figure \ref{wvel_avg} shows some significant differences in the vertical velocity profiles between the simulations. Firstly, the \textit{left panels} show the meridionally averaged updraft is stronger, broader (in longitude) and larger in vertical extent in the ``Full'' case. However, the ``Deep'' case appears similar to the ``Shallow''. Secondly, the \textit{right panels} show, for the zonally averaged vertical circulation, as we move from the ``Shallow'' to ``Deep'' to ``Full'' cases, the updraft at the equator, and over the poles, strengthens, and the fast flowing downdrafts flanking the jet (in latitude), become stronger. Similar to jets on Earth, the regions flanking the jet are baroclinically unstable and will, therefore, generate eddies and perturbations, such as atmospheric Rossby waves. The interaction of these perturbations with the mean flow provides a mechanism to move energy (and angular momentum) up--gradient, i.e. into the jet, and therefore sustain the jets against dissipation. \begin{figure*} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/jet_decay_100.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/jet_decay_400.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/jet_decay_800.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/jet_decay_1200.png} \caption{Figure showing the horizontal velocity (vector arrows) and vertical velocity (colour scale) for the ``Full'' case (see Table \ref{model_names} for explanation) at $1\times 10^5$ Pa (1 bar) and after 100 (\textit{top left}), 400 (\textit{top right}), 800 (\textit{bottom left}) and 1200 (\textit{bottom right}) days. Although the colour scales differ, the contour lines are the same for \textit{all panels}.} \label{jet_decay} \end{figure*} \begin{figure*} \centering \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_wvel_merid_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/shallow_wvel_zon_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_wvel_merid_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_gconst_wvel_zon_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_wvel_merid_200_1200.png} \hspace*{-0.7cm}\includegraphics[width=7.0cm,angle=90]{./Fig/HD209/deep_wvel_zon_200_1200.png} \caption{Figure showing vertical velocity, as a function of pressure, for the ``Shallow'', ``Deep'' and ``Full'' cases (see Table \ref{model_names} for explanation) as the \textit{top}, \textit{middle} and \textit{bottom panels} respectively. The \textit{left} and \textit{right panels} show vertical velocity as a function of longitude where a meridional average (performed in a point--wise fashion, i.e. $\int\,vd\,\phi$ as opposed to $\int\,\cos\phi vd\,\phi$, to emphasise differences in the vertical flow towards the polar regions) has been performed, and of latitude where a zonal average has been performed, respectively.} \label{wvel_avg} \end{figure*} \citet{showman_2011b} show that the jet pumping mechanism for hot Jupiters is unlikely to be similar to that relevant to Earth's mid--latitude jets, i.e. the poleward motion of atmospheric Rossby waves. In fact the likely culprit, given the planetary scale of Rossby waves for hot Jupiters, is the interaction between standing atmospheric waves and the mean flow. Such standing waves are planetary in scale, and therefore are certainly poorly represented by any model which adopts the `traditional' approximation \citep[as discussed in][]{white_1995}. Additionally, \citet{showman_2011b} show that the vertical transport of eddy momentum is a vital ingredient in the balance of superrotation at the equator. Therefore, it is clear that altering the efficiency of vertical transport will affect this mechanism, leading to a change in the balance of the pumping of the jet. Work is in progress to fully investigate this issue, which requires simulations run for a significantly longer integration time (Mayne et al, in preparation). \section{Conclusion} \label{conclusions} We have presented the first application of the UK Met Office global circulation model, the Unified Model, to hot Jupiters. In this work we have tested the ENDGame dynamical core (the component of a GCM which solves the equations of motion of the atmosphere) using a shallow--hot Jupiter \citep[SHJ, as prescribed in][]{menou_2009} and a HD 209458b test case \citep{heng_2011}. This work represents the first results of such test cases using a meteorological GCM solving the non--hydrostatic, deep--atmosphere equations. We have also completed the test case using the same code under varying levels of simplification to the governing dynamical equations. This work is complementary to the testing we have performed modelling Earth--like systems \citep{mayne_2013}. In this work we suggest that, when relaxing the canonical simplifications made to the dynamical equations, the deeper regions of the radiative atmosphere, and the radiatively inactive regions, do not reach a steady state and are still evolving throughout the 1200 day test case. We have found that moving to a more complete description of the dynamics activates exchange between the vertical and horizontal momentum, and the deeper and shallower atmosphere. This leads to a degradation of the eastward prograde equatorial jet, and could represent either the beginnings of a new equilibrium state or multiple states, which may be dependent on the initial conditions of the radiatively inactive region of the atmosphere. In a future work we will investigate longer integration times, and explore the effect of simplifications to the dynamical equations on examples of jet pumping mechanisms in these objects. These results suggest that the test cases performed are not necessarily good benchmarks for a model solving the non--hydrostatic, deep--atmosphere equations. We also aim to investigate the importance of the deeper atmosphere, and therefore, move the inner boundary for a HD 209458b simulation much deeper to $\sim 10^8$ Pa (or kbar) levels. This will require adaptation of the equation of state and increased flexibility in the prescription of $c_p$. These test cases have been performed using a Newtonian cooling scheme. As discussed in \citet{showman_2009} such a scheme does not include blackbody thermal emission of the gas itself, which can be significant when a region of heated material is advected into a region of net cooling. Using such a scheme the gas is just arbitrarily heated or cooled without taking into account its re-radiation into the surrounding area. In fact, as only the temperature is adjusted without knowledge of the specific heat capacity or quantity of material in a given cell (nor its optical properties) the energy deposited (or removed) from the system is unrepresentative. These problems can lead to regions where the heating or cooling is artificially high. To correct this we are adapting a non--grey radiative transfer scheme, under the two--stream approximation, which will be coupled to the UM dynamical core, ENDgame, under hot Jupiter conditions. The subsequent comparison to observations will be performed with a more physically meaningful model, once the coupling of our adapted schemes is complete. The UM GCM is a powerful tool with which to study the effect, on the predicted states of exoplanet atmospheres, of both the interaction between the observable and deep atmosphere, and canonically made approximations to the governing dynamical equations. The ability to alter the level of simplification of the underlying dynamical equations (as provided by the ENDGame dynamical core) will prove vital as we explore exotic climate regimes where assumptions based on Earth's atmosphere cannot \textit{a priori} be assumed valid. \begin{acknowledgements} Firstly, we would like to extend our gratitude to the referee whose thoughtful and clearly expounded comments greatly improved this manuscript. We would also like to thank Tom Melvin for his expert advice. This work is supported by the European Research Council under the European Communitys Seventh Framework Programme (FP7/2007-2013 Grant Agreement No. 247060) and by the Consolidated STFC grant ST/J001627/1.This work is also partly supported by the Royal Society award WM090065. The calculations for this paper were performed on the DiRAC Facility jointly funded by STFC, the Large Facilities Capital Fund of BIS, and the University of Exeter. \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} In the framework of the Standard Model of Particle Physics, the hadronic decay $D^0 \to K^0_S \pi^+ \pi^-$~ provides access to the measurement of the mixing parameters in the neutral $D$ meson system. Due to a cancellation via the GIM mechanism \cite{GIM} and CKM suppression \cite{Cabibbo, CKM}, mixing is suppressed in the neutral charm sector and therefore experimentally challenging. The Standard Model predicts CP violation in $D^0-\bar{D}^0$ mixing (indirect CP violation) to be $\leq 10^{-5}$ \cite{Cicerone}. More recent theoretical calculations \cite{Lenz} find sizable effects on the limit \cite{Cicerone} due to the corrections from the leading Heavy Quark Expansion (HQE) \cite{ManoharWise} contributions. Evidence for New Physics (NP) could be found if discrepancies between experimental observations and the predicted level of indirect CP violation in the Standard Model occur.\\ A measurement of the mixing parameters $x_D$ and $y_D$ as well as of the parameters $|q/p|$ and $\phi = arg(q,p)$, which govern indirect CP violation, will be performed based on a time-dependent amplitude-model analysis of the full LHCb dataset of 2011 and 2012 corresponding to an integrated luminosity of $3 \, \mathrm{fb^{-1}}$. This analysis will combine both prompt and semileptonically-tagged $D^0 \to K^0_S \pi^+ \pi^-$~ decays to obtain a large sample of high purity and with sensitivity at all $D^0$ decay times. Sensitivities of $0.23\%$ for $x_D$, of $0.17\%$ for $y_D$, of $0.2$ for $|q/p|$ and of $11.7^{\circ}$ for $\phi$ are expected for the combined dataset of $3 \, \mathrm{fb^{-1}}$ \cite{Implications}.\\ The current world-averages provided by the {\it{Heavy Flavour Averaging Group}} \cite{HFAG} are $x_D = (0.49^{+0.17}_{-0.18}) \, \%$, $y_D = (0.74 \pm 0.09) \, \%$, $|q/p| = (0.69^{+0.17}_{-0.14})$ and $\phi = (-29.6^{+8.9}_{-7.5}) ^{\circ}$ allowing for CP violation. \section{Theory} The hadronic decay $D^0 \to K^0_S \pi^+ \pi^-$ provides access to the measurement of the mixing parameters of the neutral $D$-meson system \begin{align} x_D &= \frac{2(m_1-m_2)}{\Gamma_1 + \Gamma_2}, \\ y_D &= \frac{\Gamma_1 - \Gamma_2} {\Gamma_1 + \Gamma_2}. \end{align} since the $D^0$ meson can either decay directly or indirectly after oscillating into a $\bar{D}^0$ affecting the $D^0$ decay time distribution.\\ The linear superposition of the flavour eigenstates $\Ket{D^0}$ and $\Ket{\bar{D}^0}$ yielding the mass eigenstates $\Ket{D_1}$ and $\Ket{D_2}$ \begin{align} \Ket{D_{1,2}} &= p \Ket{D^0} \pm q \Ket{\bar{D}^0} \end{align} implies indirect CP violation or CP violation in mixing if $|q/p| \neq 1$ while direct CP or CP violation in decay is manifested in a difference in the rates of a decay and its charge-conjugated decay, e.g. $\Gamma(D^0 \to K^0_S \pi^+ \pi^-) \neq \Gamma(\bar{D}^0 \to K^0_S \pi^+ \pi^-)$. In consequence, a time-dependent amplitude analysis is sensitive to the parameters $|q/p|$ and $\phi = arg(q,p)$ governing indirect CP violation whereas a time-integrated analysis of $D^0 \to K^0_S \pi^+ \pi^-$~ decays grants access to direct CP violation. \section{The LHCb experiment at the Large Hadron Collider} The analysed data are recorded by the LHCb experiment at the Large Hadron Collider (LHC) \cite{LHC_main, LHC_general, LHC_injector} at CERN. Pure proton beams are produced by stripping off the electron of hydrogen atoms and the protons are then initially accelerated by the linear collider (LINAC 2). Successively, the protons are accelerated further in the Booster, the Proton Synchrotron (PS) and the Super Proton Synchrotron (SPS) before being injected into the LHC and reaching their final collision energy.\\ The description of the LHCb detector is based on Ref. \cite{LHCb}. The LHCb detector is a single-arm spectrometer illustrated in Fig. \ref{Experiment:LHCb} where the collision point is chosen as the origin of a right-handed coordinate system depicted in Fig. \ref{Experiment:LHCb}. With a cross section of $10\%$ of all visible events inside the LHCb detector acceptance for producing charm quarks \cite{xsec}, the LHCb detector is perfectly suited for studies in the charm system.\\ The beam pipe is enclosed by the Vertex Locator (VELO) aligned such that the collision point of the protons is located in the centre of the $x$-$y$ plane of the VELO. Built of silicon strip sensors, the VELO provides measurements of track and vertex coordinates with high precision. Apart from the VELO, the tracking system consists of the Tracker Turicensis (TT), and three tracking stations (T1-T3) subdivided into the Inner and Outer Trackers, (IT) and (OT). The TT as well as the IT are composed of silicon microstrip sensors whereas the OT employs straw tubes. Embedded in a $4 \, \mathrm{Tm}$ dipole magnet, this system provides both track coordinate and momentum measurements. Approximately a third of all $K^0_S$ mesons decay inside the VELO acceptance - so-called long tracks (L) - and exhibit a better momentum resolution than so-called downstream tracks (D) of $K^0_S$ mesons decaying outside the VELO acceptance. A system of Ring Imaging Cherenkov Detectors (RICH) is used to obtain excellent separation between kaons and pions. The $\mathrm{C_4F_{10}}$ and aerogel radiators of RICH1 ensure particle identification for low-momentum charged particles whereas the $\mathrm{CF_4}$ radiator of RICH2 show a better performance for the high-momentum range. The shashlik calorimeter system composed of scintillating tiles and lead absorbers in the Electromagnetic Calorimeter (ECAL) and iron absorbers in the Hadronic Calorimeters (HCAL), respectively, provide identification and energy measurements of electrons, photons and hadrons. To identify electrons in the trigger, a Scintillator Pad Detector (SPD) and a Preshower Detector (PS) are installed in front of the ECAL. The muon system (M1-M5) uses multi wire proportional chambers filled with a gas mixture of $\mathrm{Ar:C0_2:CF_4}$ and in the inner region of M1 triple-GEM detectors to measure the spatial coordinates of muon tracks and between the M2 to M5 stations, iron blocks serve as absorbers. \begin{figure}[H] \centering \includegraphics[width=.85\textwidth]{LHCb.pdf} \caption[The LHCb detector]{The LHCb detector. The figure is taken from Ref. \cite{LHCb}.} \label{Experiment:LHCb} \end{figure} \section{Analysis} At LHCb, $D^0 \to K^0_S \pi^+ \pi^-$~ decays are accessible through various decay channels (charge conjugation is implied throughout). A sample of $D^{* \, +} \to D^0 \pi^+$ decays produced directly in the proton-proton collisions (prompt) has a high yield due to the high production cross section. Due to a cut on the $D^0$ decay time to suppress secondary $D^{* \, +} \to D^0 \pi^+$ decays, only $D^0$ decay times above $0.2 \, \mathrm{ps}$ are accessible. In comparison, the trigger efficiency of semileptonically-tagged $D^0 \to K^0_S \pi^+ \pi^-$~ decays is high for all $D^0$ decay times. A subsample from $\bar{B}^0 \to D^{* \, +} \mu^- \bar{\nu}_{\mu}$ decays is characterised by a cleaner signature due to the additional information from the $D^*$ decay than the sample from $B^- \to D^0 \mu^- \bar{\nu}_{\mu}$ decays. In the following, the discussion will be restricted to the analysis of $D^0 \to K^0_S \pi^+ \pi^-$~ decays originating from $B^- \to D^0 \mu^- \bar{\nu}_{\mu}$ decays.\\ The complete LHCb dataset is passed through an offline preselection algorithm after triggering to select specific decay types. The hardware trigger exploits the transverse momentum of the muon whereas the software-based second stage requires the muon candidate to pass criteria on momentum, transverse momentum, the track $\chi^2/dof$ and impact parameter amongst others. The dataset is then reduced using selection criteria which were revised to increase the efficiency especially at the boundaries of the Dalitz plane and to flatten the efficiency across the Dalitz plane. A uniform efficiency in the Dalitz variables $m^2_{K^0_S\pi^+}$, $m^2_{K^0_S\pi^-}$ and $m^2_{\pi^+\pi^-}$ is required to minimise efficiency corrections and thus also the systematic uncertainty corresponding to the correction. The events selected by the preselection are required to pass through an offline selection and are then used to train a multivariate classifier to further supress background events. A comparison of the previous and revised preselection efficiencies for $B^- \to D^0 (\to K^0_S \, \mathrm{(LL)} \, \pi^+ \pi^-) \mu^- \bar{\nu}_{\mu}$~ decays and the trigger efficiency for various topological trigger lines measured relative to the $B^- \to D^0 (\to K^0_S \, \mathrm{(LL)} \, \pi^+ \pi^-) \mu^- \bar{\nu}_{\mu}$~ candidates having passed the preselection are illustrated in Fig. \ref{Stripping_eff} and in Fig. \ref{Hlt2_eff}, respectively.\\ As implied in Fig. \ref{Stripping_eff} and in Fig. \ref{Hlt2_eff}, a slight increase of the efficiency for $B^- \to D^0 (\to K^0_S \, \mathrm{(LL)} \, \pi^+ \pi^-) \mu^- \bar{\nu}_{\mu}$~ candidates was achieved but variations in the preselection efficiency as a function of $m^2_{K^0_S\pi^+}$ and $m^2_{\pi^+\pi^-}$, respectively, are still observed. In case of the $B^- \to D^0 (\to K^0_S \, \mathrm{(DD)} \, \pi^+ \pi^-) \mu^- \bar{\nu}_{\mu}$~, the variations introduced by the trigger lines are large and thus a variation in efficiency is inevitable leading to sizeable efficiency corrections. \begin{figure}[H] \centering \begin{subfigure}{.45\textwidth}% \includegraphics[width=1.\textwidth, angle=0]{Charm_Comparison_EffMC_D02KSPiPiLL1.pdf} \captionof{figure}{Versus $m^2_{K^0_S\pi^+}$} \end{subfigure}% \begin{subfigure}{.45\textwidth}% \includegraphics[width=1.\textwidth, angle=0]{Charm_Comparison_EffMC_D02KSPiPiLL2.pdf} \captionof{figure}{Versus $m^2_{\pi^+\pi^-}$} \end{subfigure}% \caption{Efficiencies of the previous and revised preselection relative to a phase-space Monte-Carlo simulation sample of $D^0 \to K^0_S \, \mathrm{(LL)} \,\pi^+ \pi^-$~ decays originating from $B^- \to D^0 \mu^- \bar{\nu}_{\mu}$ decays studied on a Monte-Carlo simulation sample for 2011 running conditions.} \label{Stripping_eff} \end{figure} \begin{figure}[H] \centering \begin{subfigure}{.45\textwidth}% \includegraphics[width=1.\textwidth, angle=0]{CharmWG_Comparison_Hlt2_vs_Stripping20r1p1_EffMC_D02KSPiPiLL1.pdf} \captionof{figure}{Versus $m^2_{K^0_S\pi^+}$} \end{subfigure}% \begin{subfigure}{.45\textwidth}% \includegraphics[width=1.\textwidth, angle=0]{CharmWG_Comparison_Hlt2_vs_Stripping20r1p1_EffMC_D02KSPiPiLL2.pdf} \captionof{figure}{Versus $m^2_{\pi^+\pi^-}$} \end{subfigure}% \caption{Efficiencies of various topological trigger requirements relative to the $D^0 \to K^0_S \, \mathrm{(LL)} \,\pi^+ \pi^-$~ candidates originating from $B^- \to D^0 \mu^- \bar{\nu}_{\mu}$ decays having passed the revised preselection studied on a Monte-Carlo simulation sample for 2011 running conditions.} \label{Hlt2_eff} \end{figure} \begin{wrapfigure}{r}{0.45\textwidth} \centering \includegraphics[width=.45\textwidth]{datatzplot.pdf} \caption{Dalitz plot for $B^- \to D^0 (\to K^0_S \pi^+ \pi^-) \mu^- \bar{\nu}_{\mu}$~ decays produced from generator level Monte-Carlo simulation using the \mbox{\ensuremath{{\displaystyle B}\!{\scriptstyle A}\!{\displaystyle B}\!{\scriptstyle {A\!R}}}\,\,} 2010 model \cite{Babar2010}.} \label{Dalitz_Toy} \end{wrapfigure} $D^0 \to K^0_S \pi^+ \pi^-$~ decays can be formulated as quasi two-body decays via intermediate resonances where the parametrisation of the line shapes is model-dependent. The intermediate P- and D-wave resonances are the $\rho(770), f_2(1270), \omega(782)$ mesons for the $\pi^+\pi^-$ channel, the $K^*(892)^+, K^*_2(1430)^+$ mesons for the $K^0_S\pi^{+}$ mode and the $K^*(892)^-, K^*_2(1430)^-, K^*(1680)^-$ mesons for the $K^0_S\pi^{-}$ contribution \cite{Babar2010}. With exception of the $\rho(770)$ meson which is described by a Gounaris-Sakurai distribution, the intermediate P- and D-wave resonances are modelled by a relativistic Breit-Wigner distribution \cite{Babar2010}. The S-wave contribution in the $K^0_S\pi^{\pm}$ channel corresponding to the $K^*_0(1430)^{\pm}$ meson is described by the LASS parametrisation whereas the K-matrix formalism is applied to formulate the $\pi^+\pi^-$ S-wave contributions \cite{Babar2010}. Changes to the model might be indispensable due to the achievable sensitivity. The interference of the listed amplitudes can be visualised in the Dalitz plane spanned by $m^2_{K^0_S\pi^+}$ and $m^2_{K^0_S\pi^-}$ as illustrated in Fig. \ref{Dalitz_Toy} for a sample of generator-level Monte-Carlo simulation sample of $B^- \to D^0 (\to K^0_S \pi^+ \pi^-) \mu^- \bar{\nu}_{\mu}$~ decays.\\ The time-dependent amplitude-model analysis fit will use GooFit \cite{GooFit} - a parallel fitting framework for Graphics Processing Units (GPUs) implemented in CUDA - which provides a significant speed-up compared to conventional Central Processing Units (CPUs). GooFit \cite{GooFit} is written especially for maximum-likelihood fits or time-dependent amlitude-model amplitude analyses and thus the most common line shape models like relativistic Breit-Wigner, Gounaris-Sakurai and the LASS parametrisation are available. \section{Conclusion} A measurement of the mixing parameters $x_D$ and $y_D$ as well as of the parameters $|q/p|$ and $\phi = arg(q,p)$, which govern indirect CP violation, will be performed based on a time-dependent amplitude-model analysis of the full LHCb dataset of 2011 and 2012 corresponding to an integrated luminosity of $3 \, \mathrm{fb^{-1}}$ using both prompt and semileptonically-tagged $D^0 \to K^0_S \pi^+ \pi^-$~ decays. \section*{Acknowledgements} We express our gratitude to our colleagues in the CERN accelerator departments for the excellent performance of the LHC. We thank the technical and administrative staff at the LHCb institutes. We acknowledge support from CERN and from the national agencies: CAPES, CNPq, FAPERJ and FINEP (Brazil); NSFC (China); CNRS/IN2P3 and Region Auvergne (France); BMBF, DFG, HGF and MPG (Germany); SFI (Ireland); INFN (Italy); FOM and NWO (The Netherlands); SCSR (Poland); MEN/IFA (Romania); MinES, Rosatom, RFBR and NRC ``Kurchatov Institute'' (Russia); MinECo, XuntaGal and GENCAT (Spain); SNSF and SER (Switzerland); NAS Ukraine (Ukraine); STFC (United Kingdom); NSF (USA). We also acknowledge the support received from the ERC under FP7. The Tier1 computing centres are supported by IN2P3 (France), KIT and BMBF (Germany), INFN (Italy), NWO and SURF (The Netherlands), PIC (Spain), GridPP (United Kingdom). We are thankful for the computing resources put at our disposal by Yandex LLC (Russia), as well as to the communities behind the multiple open source software packages that we depend on.
\section{Autocoding of Fault Detection Semantics} \begin{multicols}{2} \begin{figure*} \centering \includegraphics[scale=0.7]{system} \caption{Simulink Model Input For Credible Autocoding} \label{complete} \end{figure*} \end{multicols} In this section, we describe the autocoding of the fault detection semantics of a running example. The running example is a fault detection system as specified in section \ref{theory} combined with a LQR controller that has two integrators. We first point out that on the abstraction level of a computer program, the notion of a continuous-time differential equation like in (\ref{3a}) no longer applies. The running example, including the plant, need to be in discrete-time. In fact, for analysis purposes, the plant can be treated as another C program. We have the following discrete-time linear state-space systems \be \ba{l} \dps x_{c} (k+1) = A_{c} x_{c}(k) + B_{c} y(k), \cr \dps u(k) = C_{c} x_{k} + D_{c} y(k), \ea \label{sys:01} \ee \be \ba{l} \dps \hat{x}(k+1)=\hat{A} \hat{x}(k) + B u + Ly, \cr \dps r(k)=y(k)-\hat{y}(k), \ea \label{sys:02} \ee and \be \ba{l} \dps x(k+1) = A x (k) + Bu(k), \cr \dps y(k) = Cx(k), \ea \label{sys:03} \ee representing the controller, the detector, and the plant respectively. After discretization, system matrices change. However, with an abuse of notation, we use the same symbols for the discrete-time model of the system and the observer in \eqref{sys:02}--\eqref{sys:03}. We have also a discretized version of the error dynamics from (\ref{3a}), \be \dps e(k+1) = \hat{A} e(k) + E f(k) \label{sys:04} \ee where $f(k)$ represent sampled fault signal and $\hat{A}=A-LC$. The Simulink model of the controller and the fault detection system is displayed in Figure \ref{complete}. Additionally, there are also fault detection semantics in the model. They are expressed using the annotation blocks as displayed in red in Figure \ref{complete}. The annotation blocks are converted into ACSL annotations for the output code. The four Ellipsoid observer blocks represent four ellipsoid invariants. Two are for the plant states $x$ and two are for the detector states $\hat{x}$. The semantics of the plants (faulty and nominal) are expressed using the Plant annotation blocks in Figure \ref{complete}. We do not express the ellipsoid sets for the error states from (\ref{45a}) and (\ref{46a}) on the input Simulink model. The reason for this is explained in section (\ref{why}). However, we point out that the credible autocoding process will generate two ellipsoid sets on the error states $e=x-\hat{x}$, one for the nominal plant and the other for the faulty. They are just expressed in the annotations of the generated code output. Not shown in Figure \ref{complete} is the ellipsoid observer block expressing a bound on the input signal $y_{c}$. This bound is an assumption that gets transformed into an ACSL assertion. \subsection{Ellipsoid sets on the error states $e$} \label{why} In this section, we discuss the reason behind not choosing the the two ellipsoid sets from (\ref{45a}) and (\ref{46a}) to be expressed in the the input Simulink model, directly (instead we approximate it from the bounds on system and observer states in the code). Consider the following two dynamics \be \dps \hat{x}(k+1) = \hat{A} \hat{x} (k) + B u(k) + LCx(k), \label{sys:error1} \ee and \be \dps \hat{x}(k+1) = A x(k) + Bu(k) + LC e(k). \label{sys:error2} \ee Note that (\ref{sys:error1}), (\ref{sys:error2}), and (\ref{sys:02}) are equivalent. In the credible autocoding process, any ellipsoid invariant that is inserted into the code from the input model is propagated forward through the code using ellipsoid calculus~\cite{ellipsoid97}. Methods such as computing the affine transformation of an ellipsoid are very useful here, since semantically speaking, most of the generated code is consisted of assigning affine expressions to variables. For example, if an ellipsoid set $\mathcal{E} \l x, P \r$ is the pre-condition, and the ensuing block of code is semantically $x:=A x$, then the autocoder generates the post-condition $\mathcal{E} \l x, \l A P^{-1}A^{\m{T}} \r^{-1} \r$. On the actual C code, the propagation steps are much smaller so there are a sequence of intermediate ellipsoid invariants between $\mathcal{E} \l x, P \r$ and $\mathcal{E} \l x, \l A P^{-1}A^{\m{T}} \r^{-1} \r$. Let $\tilde{x}=\dps \lb \ba{c} x \cr x_{c} \ea \rb$ be the closed-loop system states, and assume that closed-loop stability analysis yields an ellipsoid invariant set $\mathcal{E} \l \tilde{x}, P_{0} \r$. Given $\mathcal{E} \l \tilde{x}, P_{\tilde{x}} \r$, the prototype autocoder can generate $P_{u}$, $P_{x}$ such that $\mathcal{E} \l u, P_{u} \r$ and $\mathcal{E} \l x, P_{x}\r$. Given $\mathcal{E} \l u , P_{u} \r$, $\mathcal{E} \l x, P_{x}\r$, and the dynamics in (\ref{sys:error1}), one can compute an ellipsoid invariant $\mathcal{E} \l \hat{x}, P_{\hat{x}} \r$ for the detector states $\hat{x}$ by solving a linear matrix inequality. Solving the LMI also yields the relaxation multipliers $\alpha>0$ and $\gamma>0$ for the quadratic inequalities in $\mathcal{E} \l u , P_{u} \r$ and $\mathcal{E} \l x, P_{x}\r $. These multipliers are used to generate an ellipsoid invariant on $e$ in the following way. Given the ellipsoid invariants $\mathcal{E} \l \hat{x}, P_{\hat{x}} \r $, $\mathcal{E} \l x, P_{x}\r$, and the error states $e= \lb \ba{cc} I & -I \ea \rb \dps \lb \ba{c} x \cr \hat{x} \ea \rb$, a correct ellipsoid invariant on $e$ is $\mathcal{E} \l e, P_{e} \r$, where \be \dps P_{e} = \l \lb \ba{cc} I & -I \ea \rb P_{x,\hat{x}}^{-1} \lb \ba{cc} I & -I \ea \rb^{\m{T}} \r^{-1} \label{sproc} \ee with $P_{x,\hat{x}} = \lb \ba{cc} \gamma P_{x} & 0 \cr 0 & \l 1-\alpha-\gamma \r P_{\hat{x}} \ea \rb$. By choosing to express the ellipsoid invariants $\mathcal{E} \l \hat{x}, P_{\hat{x}} \r $, $\mathcal{E} \l \tilde{x}, P_{\tilde{x}}\r$ on the input Simulink model, the autocoder can automatically produce ellipsoid invariants on the error states $e$. Alternatively, if we choose to instead to express the ellipsoids on $e$ and $\mathcal{E} \l \tilde{x}, P_{\tilde{x}}\r$ on the Simulink model, then the autocoder encounters the problem of computing an ellipsoid invariant for $\hat{x}$ given the dynamics in (\ref{sys:error2}), with the assumptions of $\mathcal{E} \l \tilde{x}, P_{\tilde{x}} \r$ and $ \mathcal{E} \l e, P_{e} \r$. This is infeasible since $A$ is not stable in this example. Without an invariant set for $\hat{x}$, there are no safety bounds on the program variables that correspond to $\hat{x}$. In the first option, the annotations generated by the credible autocoding process can guarantee both the safety property (the ellipsoid bounds on the variables correspond to $x_{c}$ and $\hat{x}$) and the liveness property of the fault detection system i.e. the two ellipsoids on the error states. In the second option, they can only guarantee the latter. \subsection{Ellipsoid sets in the Simulink model} To generate the ellipsoid invariants for the credible autocoding, we have \be \dps \hat{x}(k+1) = \hat{A} \hat{x}(k) + \hat{B} \hat{u}(k) \label{fddyn} \ee with $\tilde{B}=LC$, $\hat{A} = A - LC$, $\dps \hat{u} =\lb \ba{c} u \cr x \ea \rb$, and $\hat{B} = \lb \ba{cc} B & \tilde{B} \ea \rb$. Given that the closed-loop ellipsoid set $\mathcal{E} \l \tilde{x}, P_{\tilde{x}}\r $ implies $\mathcal{E} \l u, P_{u} \r$ and $\mathcal{E} \l x ,P_{x}\r$ for some matrices $P_{u}$, $P_{x}$ by the affine transformation of ellipsoid set. With $\mathcal{E} \l u, P_{u} \r$, $\mathcal{E} \l x ,P_{x}\r$, and the detector dynamics in (\ref{sys:error1}), we have the following results for computing an ellipsoid invariant on $\hat{x}$. \begin{lemma} Let $\hat{u} = \lb \ba{c} u \cr x \ea \rb $ and assume that $\hat{u}$ belongs to the set $\lc \hat{u} | \hat{u}^{\m{T}} P_{1} \hat{u} \leq 1 \rc$. If there exist a symmetric positive-definite matrix $P$ and a positive scalar $\alpha$ that satisfies the following linear matrix inequality \be \dps \lb \ba{cc} \hat{A}^{\m{T}} P \hat{A} - P + \alpha P & \hat{A}^{\m{T}} P \hat{B} \cr \hat{B}^{\m{T}} P \hat{A} & \hat{B}^{\m{T}} P \hat{B} - \alpha P_{1} \ea \rb \prec 0 \label{lmi02} \ee then the set $\lc \hat{x} | \hat{x}^{\m{T}} P \hat{x} \leq 1 \rc$ is invariant with respect to (\ref{fddyn}). \label{lemma02} \end{lemma} First we manually compute the invariant sets for the closed-loop system. Once for the faulty plant and one more for the nominal plant using similar techniques as lemma \ref{lm1}. For the closed-loop analysis, we assume the command input $y_{c}$ is bounded. From the obtained closed-loop invariant sets, the autocoder can generate two ellipsoid invariants on $\hat{u}$. With $\mathcal{E} \l \hat{u}_{i}, P_{\hat{u}_i}, i=N,F\r$, and $N,F$ denotes respectively nominal or faulty. Now we apply lemma \ref{lemma02} twice to obtain the two ellipsoid sets $\mathcal{E} \l \hat{x}_{i}, P_{\hat{x}_{i}} \r, i=N,F$ on $\hat{x}$. As discussed before, we insert the obtained ellipsoid invariants $\mathcal{E} \l \tilde{x}_{n}, P_{\tilde{x}_i} \r,i=N,F$ and $\mathcal{E} \l \hat{x}_{i}, P_{\hat{x}_i} \r,i=N,F$ on the detector states $\hat{x}$ into the Simulink model. The ellipsoid invariants on $e$ are automatically computed by the credible autocoder using (\ref{sproc}), thus does not need to be expressed on the Simulink model. \subsection{Prototype Refinements and the Annotations} For the automatic transformation of the semantics of the fault detection and controller system in Figure \ref{complete} into useful ACSL annotations, we have further refined the prototype autocoder to be able handle the following issues: \begin{enumerate} \item Generate different sets of closed-loop semantics based on different assumptions of the plant. \item Formally expressing the faults to be able to reason about them in the invariant propagation process. \end{enumerate} The main change made to the prototype is a new capability to generate multiple different sets of closed-loop semantics based on the assumptions of the different plant semantics. For example, in the generated ACSL annotated code in listing \ref{acsl01}, there are two ellipsoid sets parameterized by the the ACSL matrix variables \emph{QMat\_1} and \emph{QMat\_2}. They express the closed-loop ellipsoid invariant sets $\mathcal{E} \l \hat{x}_i, P_{\hat{x}_i}\r,i=N,F$. The matrix variables are assigned the correct values using the ACSL functions \emph{mat\_of\_$n$x$n$\_scalar}, which takes in $n^2$ number of real-valued arguments and returns an array of size $n \times n$. For brevity's sake, the input arguments to the ACSL functions in listing \ref{acsl01} are truncated. The ellipsoid sets are grouped into two different set of semantics using the ACSL keyword \emph{behavior}. One set of semantics assumes a nominal plant and the other assumes the faulty. Each set of semantics are linked to their respective plant models by the behavior name. The pre-conditions displayed in listing \ref{acsl01} are ellipsoid invariant sets on the observer states $\hat{x}$ defined by the ACSL variables \emph{QMat\_3} and \emph{QMat\_4}. The post-conditions are generated using the invariant propagation process as described in section \ref{why}. The annotation statement \emph{PROOF\_TACTIC} is a non-ACSL element that the prototype autocoder generates to assist the automatic verification of the invariants. For example, to formally prove that the post-conditions in listing \ref{acsl01} is true given the pre-conditions, the automatic analyzer knows from the \emph{PROOF\_TACTIC} statement to apply the affine transformation strategy. Subsection \ref{autoverify} has more details on the automatic verification of the annotations. \begin{lstlisting}[caption={ACSL Expressing Multiple Sets of Closed-loop Semantics},label={acsl01}] /*@ logic matrix QMat_1=mat_of_8x8_scalar(...); */ /*@ logic matrix QMat_2=mat_of_8x8_scalar(...); */ /*@ logic matrix QMat_3 = mat_of_6x6_scalar(...) */ /*@ logic matrix QMat_4 = mat_of_6x6_scalar(...) */ /*@ behavior nominal_ellipsoid: requires in_ellipsoidQ(QMat_3, vect_of_6_scalar(observer_states[0], observer_states[1],observer_states[2], observer_states[3],observer_states[4], observer_states[5])); ensures in_ellipsoidQ(QMat_41, vect_of_12_scalar(observer_states[0]..., _io_->xhat[0]...)); @ PROOF_TACTIC (use_strategy (AffineEllipsoid)); */ /*@ behavior faulty_ellipsoid: requires in_ellipsoidQ(QMat_4, vect_of_6_scalar(observer_states[0], observer_states[1],observer_states[2], observer_states[3],observer_states[4], observer_states[5])); ensures in_ellipsoidQ(QMat_42, vect_of_12_scalar(observer_states[0]..._io_->xhat[0]..)); @ PROOF_TACTIC (use_strategy (AffineEllipsoid)); */ { for (i1 = 0; i1 < 6; i1++) { _io_->xhat[i1] = observer_states[i1]; } } \end{lstlisting} The semantics of the plant models are expressed using the C functions \emph{faulty\_plant} and \emph{nominal\_plant} in the ghost code statements denoted by the \emph{ghost} keyword. Ghost code statements are ACSL statements that are similar to the actual C code in every aspect except they are not executed and they are restricted from changing the state of any variables in the code. The semantics of the plant model are connected with their respective set of ellipsoid invariants through the usage of assertions. For example, in listing \ref{acsl02}, we have the plant states \emph{faulty\_state} and \emph{nominal\_state}, which are declared in the ghost code statements. They are linked to the same variable \emph{\_io\_-$>$xp} from the code in the two ACSL behaviors using the equal relation symbol $==$. \begin{lstlisting}[caption={ACSL Expressing the Semantics of the Plants with C code},label={acsl02}] /*@ ghost double faulty_state[6]; ghost double nominal_state[6]; */ /*@ behavior nominal_ellipsoid: assumes _io_->xp==nominal_state requires in_ellipsoidQ(QMat_1, vect_of_8_scalar(..._)); ensures in_ellipsoidQ(QMat_7, vect_of_14_scalar(...)); @ PROOF_TACTIC (use_strategy (AffineEllipsoid)); */ /*@ behavior faulty_ellipsoid: assumes _io_->xp==faulty_state requires in_ellipsoidQ(QMat_2, vect_of_8_scalar(...)); ensures in_ellipsoidQ(QMat_8, vect_of_14_scalar(...)); @ PROOF_TACTIC (use_strategy (AffineEllipsoid)); */ { for (i1 = 0; i1 < 6; i1++) { xp[i1] = _io_->xp[i1]; } } . . . /*@ ghost faulty_plant(_io_->u,faulty_plant_state); ghost nominal_plant(_io_->u,faulty_plant_state); */ \end{lstlisting} Finally the key property of fault detection is generated by a special annotation block that indicates to the autocoder, the specific ghost variable that we want to express ellipsoid invariants for. In this case, the ghost variable of interest is the one that corresponds to the error states $e=x-\hat{x}$. The variable is a ghost variable since the error states $e$ do not explicitly correspond to any variables generated in the code. The autocoder inserts the definition of the error states in the form of another ghost code statement as shown in listing \ref{acsl03}. As in the previous two code snippets, there are two different set of ellipsoid invariants on the variables error\_states[i1], which is based on initial plant model assumptions of either faulty or nominal. \begin{lstlisting}[caption={ACSL Expressing Invariant Sets on the Error States},label={acsl03}] /*@ ghost for (i1=0; i1<6; i1++) { error_states[i1]=x[i1]-observer_states[i1]; } */ /*@ behavior nominal_ellipsoid: requires in_ellipsoidQ(QMat_30, vect_of_12_scalar(x[0],...,observer_states[0],...)); ensures in_ellipsoidQ(QMat_31, vect_of_6_scalar(error_states[0],...)); @ PROOF_TACTIC (use_strategy (AffineEllipsoid)); */ /*@ behavior faulty_ellipsoid: requires in_ellipsoidQ(QMat_32, vect_of_12_scalar(x[0],...,observer_states[0],...)); ensures in_ellipsoidQ(QMat_33, vect_of_6_scalar(...)); @ PROOF_TACTIC (use_strategy (AffineEllipsoid)); */ { for (i1 = 0; i1 < 6; i1++) { observer_states[i1] = _state_->observer_states_memory[i1]; } for (i1 = 0; i1 < 6; i1++) { x[i1] = _io_->x[i1]; } } \end{lstlisting} \subsection{Automatic Verification} \label{autoverify} In ~\cite{wjf13arxiv}, a \emph{backend} to the prototype autocoder is developed. It takes as input the annotated C code generated by the autocoder and outputs a certificate of validity in the form of proofs of correctness for the annotations. The Frama-C/WP platform~\cite{framac}~\cite{wp} reads the annotated code and generates logic properties which truth is equivalent to the correctness of the annotations. The Why3 tool~\cite{boogie11why3} converts these properties into a format readable by the interactive theorem prover PVS~\cite{PVS-CADE92}. The annotation statement \emph{PROOF\_TACTIC} provides our framework with the necessary information to generate the proof to these properties and check them in PVS. While the prototype autocoder needed to be extended to handle fault detection software, the nature of the annotations on the C code and the type of logic reasoning used to prove them remained the same. As such, this backend is readily available to verify the correctness of the generated annotations. However this verification is currently done under the assumption that computations occurring in the program return the exact real value, and not the floating point approximation, of their result. This extension is left for future research. \section{Conclusion} In this paper, we have presented a framework that can rapidly generate fault detection code with a formal assurance of high-level fault detection semantics such as stability and correct fault detection. The properties are formally expressed using ellipsoid invariants. The framework, dubbed credible autocoding, can generate the fault detection code as well as the invariant properties for the code. Moreover, the generated invariant properties can be be verified in using semi-automatic theorem prover. We have demonstrated that the credible autocoding prototype that was previously applied to control systems can be extended to fault detection systems with some additions. We applied the prototype tool to an example of observer-based fault detection system running with a LQR controller of a $3$ degrees-of-freedom helicopter. The prototype was able to autocode the fault detection semantics successfully. In this paper we only consider a simple output observer for fault detection to demonstrate the autocoding steps. However, the idea can be extended to more complicated fault detection methods. \subsection*{{\normalsize\b#1}}} \renewcommand{\t}{^{\mbox{\tiny\sf T}}} \newcommand{^{\mbox{\tiny\sf T}}}{^{\mbox{\tiny\sf T}}} \newcommand{\ensuremath{\mathcal{B}_{\varepsilon}}}{\ensuremath{\mathcal{B}_{\varepsilon}}} \newcommand{\ensuremath{\textup{int}\,}}{\ensuremath{\textup{int}\,}} \newcommand{\tu}[1]{\,\,\textup{#1}\,\,} \newcommand{\ensuremath{\textup{cl}\,}}{\ensuremath{\textup{cl}\,}} \newcommand{\ensuremath{\textup{bd}\,}}{\ensuremath{\textup{bd}\,}} \newcommand{\ensuremath{x_{0}}}{\ensuremath{x_{0}}} \newcommand{\ensuremath{\operatorname{co}}}{\ensuremath{\operatorname{co}}} \newcommand{\ensuremath{\operatorname{cone}}}{\ensuremath{\operatorname{cone}}} \newcommand{\ensuremath{\textup{sgn}\,}}{\ensuremath{\textup{sgn}\,}} \newcommand{\varepsilon}{\varepsilon} \newcommand{\ensuremath{\operatorname{aff}}}{\ensuremath{\operatorname{aff}}} \newcommand{\ensuremath{\operatorname{rank}}}{\ensuremath{\operatorname{rank}}} \newcommand{\begin{enumerate}}{\begin{enumerate}} \newcommand{\end{enumerate}}{\end{enumerate}} \newcommand{\newcommand}{\newcommand} \newcommand{\renewcommand}{\renewcommand} \newcommand{\begin{verbatim}}{\begin{verbatim}} \newcommand{\end{verbatim}}{\end{verbatim}} \DeclareMathOperator{\fl}{\mathcal{F}} \newcommand{\textit}{\textit} \newcommand{\llbracket}{\llbracket} \newcommand{\rrbracket}{\rrbracket} \newcommand{\Rightarrow}{\Rightarrow} \newcommand{\Leftarrow}{\Leftarrow} \newcommand{\bq}{\textbf} \newcommand{\m}{\textrm} \newcommand{\bb}{\mathbb} \newcommand{\til}{\texttildelow} \newcommand{\be}{\begin{equation}} \newcommand{\ee}{\end{equation}} \newcommand{\dps}{\displaystyle} \renewcommand{\l}{\left(}\renewcommand{\r}{\right)} \newcommand{\lc}{\left\{}\newcommand{\rc}{\right\}} \newcommand{\lb}{\left[}\newcommand{\rb}{\right]} \newcommand{\ba}[1]{\begin{array}{#1}} \newcommand{\ea}{\end{array}} \newcommand{\ra}{\rightarrow} \newcommand{\li}{\left |} \newcommand{\ri}{\right |} \newcommand{\pde}[2]{\frac{\partial #1}{\partial #2}} \newcommand{\ode}[2]{\frac{d #1}{d #2}} \newcommand{\odee}[3]{\frac{d^{#3} #1}{d #2^{#3}}} \newcommand{\pdee}[3]{\frac{\partial^{#3} #1}{\partial #2^{#3}}} \newcommand{\bn}{\begin{enumerate}} \newcommand{\en}{\end{enumerate}} \newcommand{\bt}{\begin{theorem}} \newcommand{\et}{\end{theorem}} \newcommand{\y}[1]{\lambda_{#1}} \newcommand{\ninf}{{\oplus}^{-\infty}} \newcommand{\pinf}{{\oplus}^{+\infty}} \newcommand{\nninf}{{\otimes}^{-\infty}} \newcommand{\ppinf}{{\otimes}^{+\infty}} \newcommand{\ir}{\mathbb{I}\mathbb{R}} \newcommand{\closure}[2][3]{{}\mkern#1mu\overline{\mkern-#1mu#2}} \newcommand{\ep}{\mathcal{E}_{P}} \newcommand{\mr}{\mathcal{M}_{r}} \newcommand{\mfa}{\mathcal{M}_{f,a}} \newcommand{\mfp}{\mathcal{M}_{f,p}} \newcommand{\mt}{\m{T}} \newcommand{\F}{\mathbb{F}} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{prop}[theorem]{Proposition} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{rem}[theorem]{Remark} \title{\LARGE \bf Credible Autocoding of Fault Detection Observers } \author{Timothy E. Wang$^{1}$, Alireza Esna Ashari$^{2}$, Romain J. Jobredeaux$^{3}$, and Eric M. Feron$^{4} \thanks{*This article was prepared under support from NSF Grant CNS - 1135955 ``CPS: Medium: Collaborative Research: Credible Autocoding and Verification of Embedded Software (CrAVES)'', NASA Grant NNX12AM52A ``Validation Elements For Loss-of-Control Recovery Operations (VELCRO)'', the Army Research Office under MURI Award W911NF-11-1-0046, and ANR ASTRID project VORACE. \thanks{$^{1}$Timothy Wang is a PhD student in the Department of Aerospace Engineering, Georgia Tech, Atlanta, GA 30332, USA {\tt\small <EMAIL>}}% \thanks{$^{2}$Alireza Esna Ashari is a Post-doctoral Researcher in the Department of Aerospace Engineering, Georgia Tech, Atlanta, GA 30332, USA {\tt\small <EMAIL>}}% \thanks{$^{3}$Romain Jobredeaux is a PhD student in the Department of Aerospace Engineering at Georgia Tech {\tt\small <EMAIL>}}% \thanks{$^{4}$Eric Feron is the Dutton/Ducoffe Professor with the Department of Aerospace Engineering, Georgia Tech Atlanta, GA 30332, USA {\tt\small <EMAIL>}}% } \begin{document} \maketitle \thispagestyle{empty} \pagestyle{empty} \begin{abstract} In this paper, we present a domain specific process to assist the verification of observer-based fault detection software. Observer-based fault detection systems, like control systems, yield invariant properties of quadratic types. These quadratic invariants express both safety properties of the software such as the boundedness of the states and correctness properties such as the absence of false alarms from the fault detector. We seek to leverage these quadratic invariants, in an automated fashion, for the formal verification of the fault detection software. The approach, referred to as the credible autocoding framework~\cite{wjf13arxiv}, can be characterized as autocoding with proofs. The process starts with the fault detector model, along with its safety and correctness properties, all expressed formally in a synchronous modeling environment such as Simulink. The model is then transformed by a prototype credible autocoder into both code and analyzable annotations for the code. We demonstrate the credible autocoding process on a running example of an output observer fault detector for a 3 degree-of-freedom (3DOF) helicopter control system. \end{abstract} \textbf{Keywords: Fault Detection, Software Verification, Credible Autocoding, Aerospace Systems, Formal Methods, ACSL}. \input{intro.tex} \input{fault.tex} \input{autocoding.tex} \input{autoverify.tex} \input{conclu.tex} \bibliographystyle{IEEEtran} \section{Fault detection problem formulation} In this paper we focus on observer-based fault detection of dynamic systems. Such methods need the system to be modeled by differential equations. In this paper we design the fault detection observer for a three-degrees-of-freedom laboratory helicopter. The system is modeled by nonlinear equations. Such a model can be linearized around the operating point of the system as follows \begin{eqnarray} \dot{x}(t)&=&A x(t)+B u(t)+ E f(t), \label{1}\\ y(t)&=&C x(t), \label{2} \end{eqnarray} where $x(t)\in \Re^6$ and $u(k) \in \Re^2$ are the state vector and the known input vector at time $t$, respectively. Also, $y(t) \in \Re^{3}$ is the output vector . $A$, $B$ and $C$ are state transition, input and output matrices, respectively: \small \begin{eqnarray*} A\!\!\!\!\!\!&=&\!\!\!\!\!\!\begin{pmatrix} 0 & 0 & 0 & 1 & 0 & 0\\ 0 & 0 & 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 0 & 0 & 1\\ 0 & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & 0\\ 0 & \frac{(2 m_f L_a-m_w L_m) g}{2 m_f L_a^2+2 m_f L_h^2+m_w L_m^2} & 0 & 0 & 0 & 0 \end{pmatrix}, \\ B\!\!\!\!\!\!&=&\!\!\!\!\!\!\begin{pmatrix} 0 & 0\\ 0 & 0\\ 0 & 0\\ \frac{L_a K_f}{(m_w L_w^2+2 m_f L_a^2)} & \frac{L_a K_f}{m_w Lw^2 +2 m_f L_a^2}\\ \frac{K_f}{2 m_f L_f} & \frac{-K_f}{2 m_f L_f}\\ 0 & 0 \end{pmatrix}, \\ C\!\!\!\!\!\!&=&\!\!\!\!\!\!\begin{pmatrix} 1 & 0 & 0 & 0 & 0 & 0\\ 0 & 1 & 0 & 0 & 0 & 0\\ 0 & 0 & 1 & 0 & 0 & 0 \end{pmatrix}. \label{8} \end{eqnarray*} \normalsize System parameters are given in \cite{Quanser}. $f(t) \in \Re^{n_f}$ in \eqref{1} represents an additive fault to the system that should be detected. No prior knowledge on this input signal is available. The value of $f(t)$ is zero for nominal (fault-free) system. The aim of the fault detection is to raise an alarm whenever this value differs significantly from zero (faulty system). For that purpose, an observers based fault detection method is used. Here, we do not develop new fault detection methods. Instead we focus on the software implementation of an observer-based fault detection method. We consider an actuator degradation fault for this system. Such a fault changes the behavior and the steady state of the system, and can be modeled as additive fault. The effect of the degradation can be modeled by replacing $u(t)$ in \eqref{1} with $\bar{u}(t)$ where \begin{eqnarray} \bar{u}(t)= X u(t). \end{eqnarray} Hence, we obtain the fault matrix bellow, defined in \eqref{1} \begin{eqnarray} E=B(I-X). \end{eqnarray} \subsection{Output observer design for fault detection} \label{theory} To explain the autocoding process, we select the simplest observer-based method \cite{ding2008model,isermann2005fault}. The detector observes the system, received input and output data and compares the data with the nominal response the system is suppose to have. Consider the full-order state observer bellow \begin{eqnarray} \dot{\hat{x}}(t)&=&A \hat{x}(t)+B u(t)+ L (y(t)-C \hat{x}(t)), \label{3}\\ \hat{y}(t)&=&C \hat{x}(t). \label{3a} \end{eqnarray} Using this observer, we generate a residual signal, comparing the estimated output \eqref{3a} with the measured one \begin{equation} r(t)=y(t)-\hat{y}(t). \label{4} \end{equation} We compare the residual signal $r(t)$ against a predefined threshold. If the threshold is reached, a fault alarm is raised. In order to explain how the method works and how the observer should be designed, we introduce estimation error $e(t)=x(t)-\hat{x}(t)$ and calculate the error dynamics \begin{eqnarray} \dot{e}(t)&=&(A-LC)e(t)+E f(t), \label{5}\\ r(t)&=&Ce(t). \label{5a} \end{eqnarray} From \eqref{5}--\eqref{5a}, $r(t)$ goes to zero if $f(t)$ is zero and the observer matrix $L$ is chosen so that $A-LC$ is stable. Note that $L$ is the only design parameter for this observer. In practice, usually we do not to raise a fault alarm if $f(t)$ is too small. Hence we suppose $\|f(t)\|>\sigma$ is a fault that must be detected. Consequently $\|r(t)\|>r_{th}$ raises a fault alarm, where $r_{th}$ is the threshold corresponding to $\sigma$. \section{A formal method to verify fault detection} \label{veri} The theory behind fault detection methods is presented in Section \ref{theory}. However, there always exist a semantic gap between theory and real implementation. The methods in Section \ref{theory} should be implemented in form of software, either in graphical environments of control design such as Simulink or as computer codes in computer languages such as Matlab or C. Due to the computation errors and replacing real numbers by floating numbers in digital computers, there exist a difference between implemented method in software form and the ideal results in theory. We aim at annotating the software so that an expert or a machine can track the operation of the software and verify that the design criteria are satisfied at software level. The idea developed in \cite{feron2010control} to formally document the stability of closed-loop systems is now extended to fault detection methods. In order to certify the fault detection software, we need to certify particular properties of the observer \begin{enumerate} \item Stability: the error dynamics is stable, i.e. $e(t)$ in \eqref{5} is around origin when system is in nominal case and stays bounded in faulty case. \item Fault detection: the residual $r(t)$ correctly detects the fault. In other words $r(t)$ dos not reach a predefined threshold if $f(t)$ is sufficiently small. \end{enumerate} To verify these properties, we use Lyapunov theory which was shown to be a very good mechanism to generate easy-to-use, formal code annotations (see \cite{feron2010control}). Note that checking the place of observer poles is not desired for software verification, because we need a mechanism to verify that each line of the code keeps invariant properties we prove in theory. Also, computer scientists and other engineers are familiar with invariant properties while understanding the connection between the place of system poles and convergent of control software needs a control theory background \cite{feron2010control}. On the other hand, formal verification methods in computer science work based on invariant sets. For that purpose, we start from informal specifications in Section \ref{theory} and translate them into formal specification as follows. These properties show how the software variables can change so that the pre-defined filter specifications are verified. Suppose the system is in nominal mode. Considering a Lyapunov function $V(t)=e^T(t)Pe(t)$, where $P$ is a positive definite matrix. We can show that the $e(t)$ remains in a predefined invariant ellipsoid \begin{eqnarray} \mathcal{E}_n=\{e(t)\in \Re^n | e^T(t)Pe(t) \leq \zeta \},\label{45a} \end{eqnarray} for all $t \in \Re$ if the observer is stable. Here, $\zeta\geq0$ is a scalar. For the faulty mode we can introduce a similar ellipsoid around the new equilibrium point. However, we do not know the new equilibrium point, as the fault is supposed to be completely unknown. But in practice, $f(t)$ is bounded. Suppose that $\|f(t)\|<\sigma$. We introduce \begin{eqnarray} \mathcal{E}_f=\{e(t)\in \Re^n | e^T(t)Pe(t) \leq \bar{\zeta} \}.\label{46a} \end{eqnarray} In \eqref{46a} $\bar{\zeta}$ is \begin{eqnarray} \bar{\zeta}(t)&=& \max_e e^T(t)Pe(t) \nonumber\\ & & s.t. \;\;\; \dot{e}(t)=(A-LC)e(t)+E f(t) \nonumber\\ & & and \;\;\; \|f(t)\|<\sigma.\label{47a} \end{eqnarray} Hence, we have two ellipsoids to which the value of the Lyapunov function may belong. As far as $\forall t$, $V(t) \in \mathcal{E}_n$, the system is in nominal mode and the observer is stable. On the other hand if $\forall t$, $V(t) \in \mathcal{E}_f$, the system is in faulty mode and the observer is stable. If $\exists t, V(t) \not\in \mathcal{E}_f$ the detector is unstable. Figure \ref{Lyapunov} shows the nominal and faulty regions in the error space. \begin{figure}[htbp] \begin{center} \includegraphics [scale=0.25,angle=270]{Lyapunov1} \end{center} \caption{Blue ellipsoid shows $\mathcal{E}_n$ and yellow ellipsoid demonstrates $\mathcal{E}_f$ in error space. Red ellipsoid is a fault scenario.} \label{Lyapunov} \end{figure} What remains is to relate the threshold on the residual signal to the value of $\zeta$. Here we do not explain the details. However, the following Lemma (or similar Lemmas) helps to calculate $\bar{\zeta}$. \begin{lemma} \label{lm1} given the system \begin{eqnarray} \dot{x}(t)=Ax(t) + E d(t),\;\; x(t)=0, \label{48} \end{eqnarray} for a given constant $\rho>0$ where $P$ is a positive definite symmetric matrix we have \begin{eqnarray} x^T(t)Px(t)<\rho \|d(t)\|, \label{49} \end{eqnarray} if there exist $Q>0$ so that \begin{eqnarray} \begin{pmatrix} A^TQ+QA &Q E & P^{1/2}\\ E^TQ & -\rho I &0\\ P^{1/2} & 0 & -\rho I \end{pmatrix}<0, \label{50} \end{eqnarray} \end{lemma} \subsection*{{\normalsize\b#1}}} \renewcommand{\t}{^{\mbox{\tiny\sf T}}} \newcommand{^{\mbox{\tiny\sf T}}}{^{\mbox{\tiny\sf T}}} \newcommand{\ensuremath{\mathcal{B}_{\varepsilon}}}{\ensuremath{\mathcal{B}_{\varepsilon}}} \newcommand{\ensuremath{\textup{int}\,}}{\ensuremath{\textup{int}\,}} \newcommand{\tu}[1]{\,\,\textup{#1}\,\,} \newcommand{\ensuremath{\textup{cl}\,}}{\ensuremath{\textup{cl}\,}} \newcommand{\ensuremath{\textup{bd}\,}}{\ensuremath{\textup{bd}\,}} \newcommand{\ensuremath{x_{0}}}{\ensuremath{x_{0}}} \newcommand{\ensuremath{\operatorname{co}}}{\ensuremath{\operatorname{co}}} \newcommand{\ensuremath{\operatorname{cone}}}{\ensuremath{\operatorname{cone}}} \newcommand{\ensuremath{\textup{sgn}\,}}{\ensuremath{\textup{sgn}\,}} \newcommand{\varepsilon}{\varepsilon} \newcommand{\ensuremath{\operatorname{aff}}}{\ensuremath{\operatorname{aff}}} \newcommand{\ensuremath{\operatorname{rank}}}{\ensuremath{\operatorname{rank}}} \newcommand{\begin{enumerate}}{\begin{enumerate}} \newcommand{\end{enumerate}}{\end{enumerate}} \newcommand{\newcommand}{\newcommand} \newcommand{\renewcommand}{\renewcommand} \newcommand{\begin{verbatim}}{\begin{verbatim}} \newcommand{\end{verbatim}}{\end{verbatim}} \DeclareMathOperator{\fl}{\mathcal{F}} \newcommand{\textit}{\textit} \newcommand{\llbracket}{\llbracket} \newcommand{\rrbracket}{\rrbracket} \newcommand{\Rightarrow}{\Rightarrow} \newcommand{\Leftarrow}{\Leftarrow} \newcommand{\bq}{\textbf} \newcommand{\m}{\textrm} \newcommand{\bb}{\mathbb} \newcommand{\til}{\texttildelow} \newcommand{\be}{\begin{equation}} \newcommand{\ee}{\end{equation}} \newcommand{\dps}{\displaystyle} \renewcommand{\l}{\left(}\renewcommand{\r}{\right)} \newcommand{\lc}{\left\{}\newcommand{\rc}{\right\}} \newcommand{\lb}{\left[}\newcommand{\rb}{\right]} \newcommand{\ba}[1]{\begin{array}{#1}} \newcommand{\ea}{\end{array}} \newcommand{\ra}{\rightarrow} \newcommand{\li}{\left |} \newcommand{\ri}{\right |} \newcommand{\pde}[2]{\frac{\partial #1}{\partial #2}} \newcommand{\ode}[2]{\frac{d #1}{d #2}} \newcommand{\odee}[3]{\frac{d^{#3} #1}{d #2^{#3}}} \newcommand{\pdee}[3]{\frac{\partial^{#3} #1}{\partial #2^{#3}}} \newcommand{\bn}{\begin{enumerate}} \newcommand{\en}{\end{enumerate}} \newcommand{\bt}{\begin{theorem}} \newcommand{\et}{\end{theorem}} \newcommand{\y}[1]{\lambda_{#1}} \newcommand{\ninf}{{\oplus}^{-\infty}} \newcommand{\pinf}{{\oplus}^{+\infty}} \newcommand{\nninf}{{\otimes}^{-\infty}} \newcommand{\ppinf}{{\otimes}^{+\infty}} \newcommand{\ir}{\mathbb{I}\mathbb{R}} \newcommand{\closure}[2][3]{{}\mkern#1mu\overline{\mkern-#1mu#2}} \newcommand{\ep}{\mathcal{E}_{P}} \newcommand{\mr}{\mathcal{M}_{r}} \newcommand{\mfa}{\mathcal{M}_{f,a}} \newcommand{\mfp}{\mathcal{M}_{f,p}} \newcommand{\mt}{\m{T}} \newcommand{\F}{\mathbb{F}} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{prop}[theorem]{Proposition} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{rem}[theorem]{Remark} \section{Introduction} The safety of dynamic systems has attracted attention over years. Many studies on fault detection of safety-critical systems are reported in recent years \cite{hwang2010survey,gertler1988survey,frank1990fault,frank1997survey,isermann1997supervision,campbell2004auxiliary,esna2011auxiliary, niemann2000design,niemann1997robust,esna2010active}. Most of those studies detection developed observer-based fault detection methods, which are suitable for online fault detection in the case of abrupt faults. The observer provides analytic redundancy for the dynamics of the system. Comparing the input-output data with the nominal data we obtain from the model of the system, we conclude whether or not the system is in nominal mode. Possible faults are modeled as additive inputs to the system. Such an additive fault changes the nominal relations between inputs and outputs. A summary of the recent developments in this domain can be found in \cite{ding2008model,patton1999robust,isermann2005fault}. Nowadays, observer-based fault detection methods are usually implemented as software on digital computers. However, there is usually a semantic gap between fault detection theory and software implementation of those methods. Computation errors may cause incorrect results. Also, engineers with little or no background in control may need to test and modify the software. Thus there is a need to express fault detection semantics at the level of software. Additionally, such an endeavor can help verify systematically that the software works correctly based on the theory and the initial design. In this paper, we present an automated process of applying control-theoretic techniques towards the verification of observer-based fault detection software. We extend our previous works~\cite{wjf11arxiv},~\cite{wjfdasc11},~\cite{heber},~\cite{wjf13arxiv} for controller systems to fault detection systems. The idea of using domain-specific knowledge in software verification is not new. However, the application of system and control theory to control software verification can be traced back to relatively recent works like~\cite{feroncsm10} and~\cite{ferretesop04}. In these papers, the authors presented a manual example of documenting a controller program with a quadratic invariant set derived from a stability analysis of the state-space representation of the controller. Since then, we have progressed towards creating an automated framework that can rapidly obtain and transform high-level functional properties of the control system into logic statements that are embedded into the generated code in the form of comments. The usefulness of these comments comes from their potential usage in the automatic verification of the code. We will refer to the logic statements as ``annotations'' and the generated code with those comments as ``annotated code''. We named the framework \emph{credible autocoding} as it is a process to rapidly generate the software as well as the annotations that guarantee some functional properties of the system. The realization of the framework is a prototype tool that we have built and applied to control systems such as a controller for the 3 degrees of freedom helicopter Quanser~\cite{Quanser}. For this paper, we have further refined the prototype to handle the addition of a fault detection system running along with the Quanser controller. The paper is organized as follows: first we introduce credible autocoding framework and the details of its implementation; this is followed by a mathematical description of the fault detection method used in our running example; finally, we describe the credible autocoding process that has been further extended for fault detection systems and the automatic verification of the annotated code produced by the credible autocoder. \section{Credible Autocoding} Credible autocoding is an automatic or semi-automatic process that transforms a system that is initially expressed in a language of high-level of abstraction, along with the mathematical proofs of its good behavior, into code, annotated with said mathematical proof. The initial level of abstraction could be a differential equation of the system and the final level could be the software binary. For the prototype implementation of our framework, we picked Simulink as the starting point and C code as the final output. Regardless of the input and output languages, the main contribution from this prototype is the automatic translation of Lyapunov type stability proofs into axiomatic semantics for the output code. Axiomatic semantics is an approach to reason about the correctness of program that traces back to the works of Charlie Hoare~\cite{hoareaxiom69}. In this approach, the semantics or mathematical meanings of a piece of code is defined through how the piece of code modifies certain logic predicates on the variable(s) of the code. The basics of axiomatic semantics are demonstrated here using two examples. In figure \ref{code01}, we have a piece of C code that computes the square of $x$ and assigns the answer to the variable $x$. Notice in the comments or annotations that precede the code, we have two logic predicates $x<=0$ and $x>=0$, preceded by the symbols ``@ requires'' and ``@ ensures''. They represent properties of $x$ that we claim to be true, respectively before and after the execution of the line of code. The keyword \emph{requires} denotes a \emph{pre-condition} and the keyword \emph{ensures} keyword denotes a \emph{post-condition}. The pre and post-conditions, together with the statement they surround, form a \emph{Hoare Triple}: they express a contract of sorts, namely, that for any execution of the program, regardless of what has happened elsewhere, if the pre-condition is true before the statement is executed, then the post-condition will be true after its execution. In this example, it is trivial to see that if the variable $x$ is non-positive before the execution of $x:=x*x$ then it will be non-negative afterwards. However we stress here that any such contract inserted as annotations in the code needs to be formally proven before it is said to be valid for the code. \begin{figure} \centering \begin{lstlisting} /* @ requires x<=0 @ ensures x>=0 */ { x=x*x; } \end{lstlisting} \label{code01} \caption{C code with ACSL} \end{figure} Consider the C implementation of a 1-dimensional linear state-space system in figure \ref{code02}. The state-transition matrix $A$ is 0.98 and the input matrix $B$ is 0.02. Unlike the previous example, this piece of code contains an infinite loop. In the case of infinite loops, the difficulty lies in finding a property that will hold both before, throughout and after the execution of the loop. Such properties are called \emph{loop invariants}. They are closely related to invariant sets. We can express such a property with a contract on the body of the loop where the pre- and post- conditions need to be identical. It is often the case, for even trivial invariant, that verifying their correctness is a non-trivial task. Comparatively speaking though, verifying the correctness of a given invariant in an automatic fashion is still a more tractable task than finding the invariants of a program automatically. For even simple examples, it can be impossible for general automatic decision procedures to compute invariants for the code without the application of domain specific knowledge. In addition, there is also an assertion $input*input<1$ denoted by the keyword \emph{assumes}. The difference between assertions and properties is that assertions are the assumptions that we make without given any proofs for it. In this case, we are going to assume the magnitude of the input variable is bounded by $1$. Unlike the property $x*x<=1$, the assertion $input*input<1$ cannot be checked for its correctness based only on the information available from the code. For this example or any other linear state-space systems, we can apply domain specific knowledge, namely Lyapunov-based theories, to compute an ellipsoid invariant set $\mathcal{E}\l x, P \r=\lc x | x^{\m{T}} P x \leq 1 \rc$. A collection of these type of quadratic stability results with an efficient computational solutions can be found in~\cite{boydlmi94}. For this example, $\mathcal{E} \l x,1\r$ forms a valid invariant set. Through the credible autocoding framework, this invariant property can be rapidly transformed into contracts for the code, and thus, in theory, makes the process of automatic verification of the generated code more feasible. \begin{figure} \centering \begin{lstlisting} /* @ assumes input*input<1; @ requires x*x<=1 @ ensures x*x<=1 */ { while (1) { x=0.98*x+0.02*input; } } \end{lstlisting} \label{code02} \caption{C code with ACSL} \end{figure} The code annotations in the two examples are expressed in the ANSI C Specification Language (ACSL) \footnote[1]{The prototype credible autocoder also produces annotations in ACSL}. In the latter sections, the autocoded fault detection semantics are also expressed in ACSL. For more details, interested readers can refer to ~\cite{baudinacsl08}.
\section{Introduction} On one hand, it is well documented in the modern literature that the conventional laws of capillarity are not adequate when applied to fluids confined by porous materials \cite{Bear}. On the other hand, the technical development of sciences allows us to observe phenomena at length scales of a very few number of nanometres. This nanophysics allows to infer applications in numerous fields, including medicine and biology. Iijima often cited as the discoverer of carbon nanotubes \cite{Iijima}, was fascinated by Kr\"atschmer \emph{et al}' Nature paper \cite{Kr}, and decided to launch out into a detailed study of nanomaterials. The recent applications revealed new behaviors that are often surprising and essentially different from those usually observed at macroscopic scale but also at microscopic scale \cit {Harris}. Nonetheless, simple models proposing qualitative behaviors need to be developed in the different fields of nanosciences; our aim is to investigate the fluid-solid interaction in static as well as in dynamic conditions by differential calculus in continuum mechanics. \newline As it was pointed out in experiments, the density of liquid water changes in narrow pores \cite{Ball}; an analytic asymptotic expression was obtained with an approximation of London potentials for liquid-liquid and solid-liquid interactions, which yields the surface interaction energy \cit {Gouin 2}. With the aim to propose an analytic expression of the density for liquid films with a nanometer thickness near a solid wall, we add together the interaction energy at the solid surface to a density-functional at the liquid-vapour interface and a square-gradient functional which represents the volume free energy of the fluid \cite{Gouin 1,Gouin 3}. The obtained functional allows to get a differential equation and boundary condition which yield the density profile in cylindrical nanotubes. \newline For shallow water, the flows of liquids on solids are mainly represented by using the Navier-Stokes equations associated with the adherence condition at the walls \cite{oron}. Recent experiments in nanofluidics seem to prove that the adherence condition is often disqualified \cite{De Gennes}. So, we can draw consequences which differ from results of classical models; the model we are presenting reveals an essential difference between the flows of microfluidics and those of nanofluidics \cite{Tabeling}. The simple laws of scales cannot be only taken into account. The film-solid interactions are accounted for in terms of the disjoining pressure. This concept of disjoining pressure has been introduced by Derjaguin in 1936 as the difference between the pressure in a phase adjacent to a surface confining it and the pressure in the bulk of this phase \cite{Derjaguin, Gouin 6}. We have previously seen that the gradient of thickness along layers creates a gradient of disjoining pressure that induces driving forces along the layer \cite{Gouin 7}. Moreover, we had noticed that the stability criterion of the flow issued from the equation of motions fits with the results of Derjaguin's school \cite{Gouin 7}. \newline The liquid flows through nanotubes also depend on the wetting conditions on the wall. Some phase transitions can appear and drastically change the liquid flows through nanotubes. Since fifteen years the literature is abundant about nanotube technology and flows inside nanotubes \cite{Mattia}. \newline The aim of this paper is not to redo the literature but to emphasis on liquid compressibility near solid walls in nanoscale conditions. We consider a nanotube made up of a cylindrical hollow tube whose diameter is of some nanometres. The length of the nanotube is microscopic and the edge of the cylinder is a solid made out of carbon or other materials \cite{Harris}. The nanotube is immersed in a liquid or a vapour made up of the same fluid. The fluid fills the interior of the nanotube. The fluid is modeled by a van der Waals fluid \cite{Korteweg,van der Waals} for which the surdeformations are taken into account (we called \emph{capillary fluid} or \emph{Cahn and Hilliard fluid} \cite{Cahn,Gouin 1}). The volume free energy of the fluid is a function not only of the density but also of the gradient of density. The conditions on the wall take account of the fluid density at its immediate proximity \cite{Gouin 2}. We first recall the equations of equilibrium and motion of capillary fluids \cite{Gouin 1}. These fluids can be modeled on the interfaces as the fluids in the immediate vicinity of solid media \cite{Gouin 2}. In liquid or vapour phase, it is possible to express the chemical potential with a development in a linear form taking account of the isothermal sound velocity values in the bulks \cite{Gouin 6}. The main expansions of the free energy in liquid and vapour phases are deduced. In nanofluidics, the interactions between fluid and solid wall dominate over the hydrodynamic behaviour of the fluid \cite{Ball}. Boundary conditions are embedding effects; they are expressed thanks to a surface energy associated with a molecular model in mean field theory \cit {Gouin 3}. In the case of cylindrical geometry, a differential equation with respect to the fluid density is obtained. Then, the profile of the fluid density in a cylinder can be deduced. The result is applied to nanotubes when the diameter ranges from a little number of nanometres to one hundred nanometres. Depending on the disjoining pressure between liquid and vapour bulks and on the wettability of the nanotube wall, we can forecast when the fluid inside the nanotube is liquid or vapour; the wall effect is dominant. The case of liquid and vapour separated by an interface is not possible when the nanotube diameter is smaller than one hundred nanometres and a liquid is generally found inside the nanotube. \newline Recently, it was showed, by using nonequilibrium molecular dynamics simulations, that liquid flow through a membrane composed of an array of aligned carbon nanotubes is four to five orders of magnitude faster than it would be predicted from conventional fluid-flow theory \cite{Majumder}. These high fluid velocities are possible because of a frictionless surface at the nanotube wall \cite{Ma}. Majunder \emph{et al} quote slip lengths on the order of microns for their experiments with nanometer size pores \cite{Majumder}. The extremely large slip lengths measured in carbon nanotubes greatly reduce the fluidic resistance and nanoscale structures could mimic extraordinarily fast flow possible in biological cellular channels \cite{Bonthuis}. By calculating the variation of water viscosity and slip length as a function of the nanotubes diameter, the results can be fully explained in the context of continuum fluid mechanics \cite{Thomas}.\newline In this paper we recalculate the flows through nanotubes by using a Navier-Stokes equation but, due to the slip condition and the Navier length, it is possible to forecast an important difference between classical Poiseuille flows and flows through nanotubes. The calculations associated with the physicochemical quality of the nanotube allow to forecast if the fluid phase inside the tube is constituted of liquid or vapour. A spectacular effect must appear when the mother bulk outside the nanotube is constituted of vapour; in this case, the volume flow through the nanotube is multiplied by a factor of the order of one million with respect to the Poiseuille flow and the velocity field through the nanotube may be very important. \section{Equation of motion and boundary conditions for a capillary fluid} \subsection{Case of conservative fluid} The second gradient theory \cite{Forest,Germain}, conceptually more straightforward than the Laplace theory, can be used to elaborate a theory of capillarity \cite{Ono}. The theory can also be used to investigate domains where the fluids are strongly inhomogeneous as in the immediate vicinity of solid walls where intermolecular forces are dominant between fluid and solid with respect to fluid interactions. By this simple way, the only change with respect to compressible fluids is that the specific internal energy is not only a function of the density $\rho $, of the specific entropy $s$, but also of $\func{grad}\rho $. Consequently, the specific internal energy $\varepsilon $ characterizes both the compressibility and the surdeformation of the fluid, \begin{equation*} \varepsilon =f(\rho ,s,\beta ),\quad \mathrm{{where}\quad }\beta =(\func{gra }\rho )^{2}. \end{equation*} We recall the main results of \emph{capillary fluids} already obtained in the literature \cite{Gouin 1,Gouin 4,Rowlinson}: The equation of conservative motions of such \emph{capillary fluids} is \begin{equation} \rho \,\mathbf{a}=\func{div}\mathbf{\mathbf{\sigma }}-\rho\, \func{grad \Omega , \label{2} \end{equation where $\mathbf{a}$ denotes the acceleration vector, $\Omega $ the extraneous force potential and $\mathbf{\mathbf{\sigma }}$ the total stress tensor. The total stress tensor is \begin{equation} \mathbf{\mathbf{\sigma }}=-p\,\mathbf{I}-\lambda \,(\func{grad}\rho )(\func grad}\rho )^{T}\ \ \mathrm{{or}\ }\ \sigma _{ij}=-p\ \delta _{ij}-\lambda \,\rho _{,i}\rho _{,j}\,,\ i,j\in \{1,2,3\} \label{3a} \end{equation where $\,^T$ denotes the transposition, wit \begin{equation*} \quad\lambda\equiv 2\,\rho \,\varepsilon _{\beta }^{\prime }\qquad \mathrm{{and \qquad }p\equiv \rho ^{2}\varepsilon _{\rho }^{\prime }-\rho\,\func{div}(\lambda \,\func grad}\rho ). \label{sigma} \end{equation* It should be noted that $\varepsilon _{s}^{\prime }$ is the Kelvin temperature expressed as a function of $\rho $, $s$ and $\beta $. It appears that only the scalar $\lambda $ accounts for surdeformation effects. As \varepsilon $ does, the scalar $\lambda$ depends on $\rho, s$ and $\beta $. For the surface tension study based on the gas kinetic theory, Rocard obtained the expression (\ref{3a}) for the stress tensor but with $\lambda$ constant \cit {Rocard}. If $\lambda $ is constant, the specific energy $\varepsilon $ reads \begin{equation*} \varepsilon (\rho ,s,\beta )=\alpha (\rho ,s)+\frac{\lambda}{2\rho }\ \beta , \end{equation* and the \emph{second gradient} term $\displaystyle {\lambda}\, \beta/(2\,\rho)$ is simply added to the specific internal energy $\alpha (\rho ,s)$ of the classical compressible fluid. The pressure of the compressible fluid is $P \equiv \rho ^{2}\alpha _{\rho }^{\prime }$ and the temperature is $T \equiv\alpha _{s}^{\prime }$. Consequently, \begin{equation*} p=P-\lambda \left(\frac{\beta }{2}+\rho\, \Delta \rho\right). \end{equation* For the thermodynamical pressure $P$, Rocard and others authors use the van der Waals pressure \begin{equation*} P=\rho\, \frac{R\,T }{1-b\rho }-a\rho ^{2} \end{equation* or other similar laws \cite{Rocard}. It should be noted that if $\lambda$ is constant, there exits a relation independent of $\func grad}\,\rho$ between $T ,\, \rho $ and $s$. \subsubsection{Case $\lambda $ constant} Equation (\ref{3a}) yields \begin{equation*} {\sigma }_{ij}=-P\,\delta _{ij}+\lambda \,\left\{ \left( \frac{1}{2}\ \rho _{,k}\rho _{,k}+\rho \rho ,_{kk}\right) \delta _{ij}- \rho _{,i}\rho _{,j}\right\}. \end{equation*} Let us denote $\omega =\Omega -\lambda \,\Delta \rho $, then Eq. (\ref{2}) reads \begin{equation} \rho \,\mathbf{a}+\func{grad}P+\rho \func{grad}\omega =0. \label{motion2} \end{equation This relation is similar to the perfect fluid case; the term $\omega $ involves all capillarity effects. From ${\sigma }_{ij{,j}}=-P_{,i}+\lambda \,\rho \rho _{,ijj}\, $ and by neglecting the extraneous forces, we obtain: \begin{equation*} \rho\,\mathbf{\mathbf{a}}+\func{grad}P=\lambda \,\rho \,\func{grad \Delta \rho . \end{equation*} \subsubsection{Thermodynamic form of the equation of motion} Commonly - and not only when $\lambda$ is constant - the equation of motion \ref{2}) can be written in a thermodynamic form \begin{equation} \mathbf{a} =\theta \func{grad}s-\func{grad}(h+\Omega ) ,\quad \mathrm{with \quad h= \varepsilon+\frac{p}{\rho}\, . \label{thermotion} \end{equation} In the non-capillarity case ($\varepsilon ^{\prime }_{\beta } =0$ or \varepsilon =\alpha (\rho ,s)$), Eq. (\ref{thermotion}) is well-known. When $T$ is constant, Eq. (\ref{thermotion}) yield \begin{equation} \mathbf{a} +\func{grad}(\pi+\Omega )=0 ,\quad {\rm with} \quad \pi= h-T\,s\,. \label{thermotion1} \end{equation} The potentials $h$ and $\pi$ are the \emph{generalized enthalpy and the chemical potential} of the capillary fluid. \subsection{Case of viscous fluid} In the case of viscous fluids, the equation of motion includes not only the stress tensor $\mathbf{\mathbf{\sigma }},$ but also the viscous stress tensor $\mathbf{\mathbf{\sigma }}_{v}$ written in the form: \begin{equation*} \bm{\mathbf{\sigma }}_{v}=\eta \ \mathtt{tr}\,\bm{D}+2\,\kappa \,\bm{D}, \end{equation* where $\bm D$ is the deformation tensor, symmetric gradient of the velocity field and $\eta $ and $\kappa$ are constant in the viscous linear case. Of course in second gradient theory, it would be coherent to add terms accounting for the influence of higher order derivatives of the velocity field to the viscous stress tensor $\mathbf{\mathbf{\sigma }}_{v}$; the surdeformation of density comes from wall effects but the variations of velocity are negligible and the second derivatives are not taken into account. Equation (\ref{2}) is modified by adding the forces associated with the viscosity and we obtain \begin{equation*} \rho \,\mathbf{a}=\func{div}(\mathbf{\mathbf{\sigma }}+\mathbf{\mathbf \sigma }}_{v})-\rho \func{grad}\,\Omega \,. \end{equation* For viscous fluid, Eq. (\ref{motion2}) reads \begin{equation} \rho \,\mathbf{a}+\func{grad}P+\rho \func{grad}(\Omega -\lambda \,\Delta \rho )-\func{div}\mathbf{\mathbf{\sigma }}_{v}=0. \label{viscous motions} \end{equation} \subsection{Boundary conditions at a solid wall} The forces acting between liquid and solid range over a few nanometres but can be simply described by a special surface energy. This energy is not the total interfacial energy which results from the direct fluid/solid contact; another energy results from the distortion in the fluid density profile near the wall. For a solid wall not too curved at a molecular scale, the total surface free energy $\varphi$ is developed as \cit {Gouin 2}: \begin{equation} \varphi (\rho_{_S})=-\gamma _{1}\rho_{_S} +\frac{1}{2}\,\gamma _{2}\,\rho_{_S}^{2}. \label{surface energy} \end{equation Here $\rho_{_S}$ denotes the limit value of the fluid density at the surface $(S)$; the constants $\gamma _{1}$, $\gamma _{2}$ as the constant $\lambda $ are positive. In the mean field approximation of molecular theory they are: \begin{equation*} \gamma _{1}=\frac{\pi c_{ls}}{12\delta ^{2}m_{l}m_{s}}\;\rho _{sol},\quad \gamma _{2}=\frac{\pi c_{ll}}{12\delta ^{2}m_{l}^{2}}, \label{coefficients} \end{equation*} with \begin{equation*} \lambda =\frac 2\pi c_{ll}}{3\sigma _{l}\,m_{l}^{2}}, \end{equation*} where $m_{l}$ et $m_{s}$ denote the molecular masses of fluid and solid, respectively, $\rho _{sol}$ is the solid density; other constants come from London potentials of liquid-liquid and liquid-solid interactions expressed in the form \begin{equation*} \left\{ \begin{array}{c} \displaystyle\;\;\;\;\;\;\varphi _{ll}=-\frac{c_{ll}}{r^{6}}\;,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ when\ }r>\sigma _{l}\;\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{and }\;\ \varphi _{ll}=\infty \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ when \ r\leq \sigma _{l}\,,\ \\ \displaystyle\;\;\;\;\;\;\varphi _{ls}=-\frac{c_{ls}}{r^{6}}\;,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ when\ }r>\delta \;\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{and }\;\ \varphi _{ls}=\infty \RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ when \ r\leq \delta \;,\ \end{array \right. \end{equation* where $c_{ll}$ et $c_{ls}$ are two positive coefficients associated with Hamaker constants, $\sigma _{l}$ and $\sigma _{s}$ denote the molecular diameters for the fluid and the solid, $\delta =\frac{1}{2}\,(\sigma _{l}+ \sigma _{s})$. The boundary condition for the density at the solid wall $(S)$ associated with the free surface energy (\ref{surface energy}) is calculated in \cite{Gouin 3}: \begin{equation} \lambda \left( \frac{d\rho }{dn}\right) _{|_{S}}+\varphi ^{\prime }(\rho_{_S})\ =0, \label{cl1} \end{equation where $n\,$ is the external normal direction to the fluid. \section{The chemical potential in liquid and vapour phases} The chemical potential of a compressible fluid at temperature $T$ is denoted by $\mu _{_{0}}$. Due to the equation of state $P\equiv P(\rho ,T)$, it is possible to express $\mu _{_{0}}$ as a function of $\rho $ (and $T$). At a given temperature, the volume free energy $g_{_{0}}$ associated with $\mu _{_{0}}$ verifies $g_{_{0}}^{\prime }(\rho )=\mu _{_{0}}(\rho )$. Due to the fact \mu _{_{0}}$ and $g_{_{0}}$ are defined to an additive constant, we add the conditions \begin{equation*} \mu _{_{0}}(\rho _{l})=\mu _{_{0}}(\rho _{v})=0\quad {\mathrm{and}\quad g_{_{0}}(\rho _{l})=g_{_{0}}(\rho }_{v}{)}=0, \end{equation* where $\rho _{l}$ and $\rho _{v}$ are the fluid densities in the liquid and vapour bulks corresponding to the plane liquid-vapour interface at temperature $T$.\newline The expressions of the two thermodynamical potentials $\mu _{_0}$ and g_{_0} $ can be expended at the first order near the liquid and vapour bulks, respectively \begin{equation*} \mu _{_{0}}(\rho )=\frac{c_{l}^{2}}{\rho _{l}}\left( \rho -\rho _{l}\right) \qquad \mathrm{{and}\qquad }\mu _{_{0}}(\rho )=\frac{c_{v}^{2}}{\rho _{v} \left( \rho -\rho _{v}\right) , \end{equation* \begin{equation*} g_{_{0}}(\rho )=\frac{c_{l}^{2}}{2\rho _{l}}\left( \rho -\rho _{l}\right) ^{2}\qquad {\mathrm{and}\qquad g_{_{0}}(\rho )=\frac{c_{v}^{2}}{2\rho _{v} \left( \rho -\rho _{v}\right) ^{2}}, \end{equation* where $c_{l}$ and $c_{v}$ are the isothermal sound velocities in the liquid and vapour bulks \cite{Gouin 6}. Consequently, at temperature $T$, it is possible to obtain the connection between the liquid bulk of density $\rho _{l_b}$ and the vapour bulk of density $\rho _{v_{b}}$ corresponding to curved interfaces (as for spherical bubbles and droplets \cite{Isola1,Isola2}): we call them the mother bulk densities. These equilibria do not obey the Maxwell rule \cite{Aifantis}, but the values of the chemical potential in the two mother bulks are equal: \begin{equation} \mu _{_{0}}(\rho _{l_b})=\mu _{_{0}}(\rho _{v_{b}}). \label{equcp} \end{equation Consequently, we define $\mu _{l_{b}}(\rho )$ and $\mu _{v_{b}}(\rho )$ at temperature $T$ as: \begin{equation*} \mu _{l_{b}}(\rho )=\mu _{_{0}}(\rho )-\mu _{_{0}}(\rho _{l_b})\equiv \mu _{_{0}}(\rho )-\mu _{_{0}}(\rho _{v_{b}})=\mu _{v_{b}}(\rho ). \end{equation*} An expansion to the first order near the liquid and vapour bulks yields \begin{equation*} \mu _{l_{b}}(\rho )=\frac{c_{l}^{2}}{\rho _{l}}\left( \rho -\rho _{l_b}\right) \qquad \mathrm{{and}\qquad }\mu _{v_{b}}(\rho )=\frac{c_{v}^{2}}{\rho _{v} \left( \rho -\rho _{v_{b}}\right) \end{equation* and due to relation (\ref{equcp}), \begin{equation*} \frac{c_{l}^{2}}{\rho _{l}}\left( \rho _{l_b}-\rho _{l}\right) =\frac{c_{v}^{2 }{\rho _{v}}\left( \rho _{v_{b}}-\rho _{v}\right) \end{equation* which clarifies the connection between $\rho _{l_b}$ and $\rho _{v_{b}}$ \newline To the chemical potential $\mu _{l_{b}}(\rho)\equiv \mu _{v_{b}}(\rho)$ at temperature $T$, we associate the volume free energies $g_{l_{b}}(\rho )$ and $g_{_{_{v_{b}}}}(\rho )$ that are null for $\rho _{l_b}$ and $\rho _{v_{b}}$, respectively: \begin{equation*} g_{l_{b}}(\rho )=g_{_{0}}(\rho )-g_{_{0}}(\rho _{l_b})-\mu _{_{0}}(\rho _{l_b})(\rho -\rho _{l_b}), \end{equation* \begin{equation*} g_{_{_{v_{b}}}}(\rho )=g_{_{0}}(\rho )-g_{_{0}}(\rho _{v_{b}})-\mu _{_{0}}(\rho _{v_{b}})(\rho -\rho _{v_{b}}). \end{equation*} The free energies $g_{l_{b}}(\rho )$ and $g_{_{v_{b}}}(\rho )$ are the reference free energies associated with the liquid and vapour mother bulks. The reference free energy $g_{l_{b}}(\rho )$ differs from g_{_{v_{b}}}(\rho )$ by a constant. Moreover, the volume free energies are expanded as \begin{equation*} g_{l_{b}}(\rho )=\frac{c_{l}^{2}}{2\rho _{l}}\left( \rho -\rho _{l_b}\right) ^{2}\qquad \mathrm{{and}\qquad }g_{v_{b}}(\rho )=\frac{c_{v}^{2}}{2\rho _{v} \left( \rho -\rho _{v_{b}}\right) ^{2}. \end{equation*} \section{Liquid and vapour densities in a nanotube} A nanotube is constituted of a hollow cylinder of length size $\ell $ and of small diameter $d=2R$, ($d/\ell \ll 1$). We consider solid walls with a large thickness with regards to molecular dimensions such that the surface energy verifies an expression in form (\ref{surface energy}). We assume that a capillary fluid is a convenient model to represent fluids inside the nanotube. \newline At equilibrium, far from the nanotube tips and by neglecting the external forces, the profile of density is solution of Eq. (\ref{thermotion1}) with $\mathbf{a}=0$ and $\pi =\mu _{_{0}}-\lambda \,\Delta \rho $ : \begin{equation} \lambda \,\Delta \rho =\mu _{_{0}}(\rho )-C, \label{densityequ} \end{equation where $C$ is an additional constant. The value of $ C$ is associated with the density value in the mother bulk outside the nanotube (where $\Delta \rho =0$) \cite{Derjaguin}. Consequently, the reference density value $\rho _{_{ref}}$ may be chosen as $\rho _{l_b}$ or $\rho _{v_{b}}$.\newline We consider the cases when exclusively liquid or vapour fill up the nanotube; the mother bulk can be as well liquid as vapour. The profile of density is given by the differential equation: \begin{equation} \lambda \,\left( \frac{d^{2}u}{dr^{2}}+\frac{1}{r}\frac{du}{dr}\right) \frac{c_{_{ref}}^{2}}{\rho _{_{ref}}}\ u=0,\qquad \mathrm{with}\quad u=\rho -\rho _{_{ref}}, \label{densityprofile1} \end{equation where $c_{_{ref}}$ is the isothermal sound velocity associated with liquid mother bulk (respectively vapour mother bulk) when the liquid phase fills up the nanotube (respectively vapour phase). In cylindrical coordinates, $r$ denotes the radial coordinate. The reference length is \begin{equation*} \delta _{_{ref}}=\sqrt{{\,\lambda \,\rho _{_{ref}}}/{c_{_{ref}}^{2}}}\,. \end{equation* We denote by $x$ the dimensionless variable such that $r=\delta _{_{ref}}\,x$. Equation (\ref{densityprofile1}) reads \begin{equation*} \frac{d^{2}u}{dx^{2}}+\frac{1}{x}\frac{du}{dx}-\ u=0. \label{densityprofile2} \end{equation* The solutions of Eq. (\ref{densityprofile2}) in classical expansion form $u=\sum_{n=0}^{\infty }a_{n}x^{n}$ yield: \begin{equation*} \sum_{n=2}^{\infty }n^{2}\,a_{n}\,x^{n-2}-a_{n-2}\,x^{n-2}=0\quad \Longrightarrow \quad n^{2}\,a_{n}=a_{n-2}\,. \end{equation* Due to the symmetry at $x=0$, the odd terms are null and consequently, \begin{equation*} u=a_{_{0}}\,\sum_{p=0}^{\infty }\ \frac{1}{4^{p}\,(p\,!)^{2}}\ x^{2p}\,. \end{equation* The series has an infinite radius of convergence. Let us define the quantities \begin{eqnarray*} \qquad \qquad f(x) &=&\sum_{p=0}^{\infty }\ \frac{1}{4^{p}\,(p!)^{2}}\ x^{2p}, \\ \qquad \qquad h(x) &\equiv &f^{\prime }(x)=\sum_{p=1}^{\infty }\ \frac{2p} 4^{p}\,(p!)^{2}}\ x^{2p-1}, \\ \qquad \qquad k(x) &\equiv &f^{\prime \prime }(x)=\sum_{p=1}^{\infty }\ \frac 2p\,(2p-1)}{4^{p}\,(p!)^{2}}\ x^{2p-2}. \end{eqnarray* Consequently, $u=a_{_{0}}\,f(r/\delta _{_{ref}})$. The boundary condition \ref{cl1}) at $x=R/\delta _{_{ref}}$ is: \begin{equation*} \lambda \,\frac{du}{dx}= \gamma _{1}-\gamma _{2}\,\rho \qquad {\rm or} \qquad a_{_{0}}=\frac{\delta _{_{ref}}\,\left( \gamma _{1}-\gamma _{2}\,\rho _{_{ref}}\right) }{\lambda \,h\left( \frac{R}{\delta _{_{ref}}}\right) +\gamma _{2}\,\delta _{_{ref}}\,f\left( \frac{R}{\delta _{_{ref}}}\right) } \end{equation* and the density profile reads \begin{equation*} \rho =\rho _{_{ref}}+\frac{\delta _{_{ref}}\,\left( \gamma _{1}-\gamma _{2}\,\rho _{_{ref}}\right) }{\lambda \,h\left( \frac{R}{\delta _{_{ref}} \right) +\gamma _{2}\,\delta _{_{ref}}\,f\left( \frac{R}{\delta _{_{ref}} \right) }\ f\left( \frac{r}{\delta _{_{ref}}}\right) . \end{equation* Let us consider the free volume energy $g_{\rho _{ref}}(\rho )$, null in the mother bulk of density $\rho _{_{ref}}$ (where $\rho _{_{ref}}=\rho _{l_b}$ or $\rho _{_{ref}}=\rho _{v_{b}}$) chosen as the reference mother bulk; the free energy $\phi $ per unit of volume in a inhomogeneous fluid is \begin{equation*} \phi =g_{\rho _{ref}}(\rho )+\frac{\lambda }{2}\,\left( \frac{d\rho }{dr \right) ^{2} \end{equation* and consequently, if $\phi $ is expressed as a function of $r$, \begin{eqnarray} \phi (r) &=&g_{\rho _{ref}}\left( \rho _{_{ref}}+\frac{\delta _{_{ref}}\,\left( \gamma _{1}-\gamma _{2}\,\rho _{_{ref}}\right) f\left( \frac{r}{\delta _{_{ref}}}\right) }{\lambda \,h\left( \frac{R}{\delta _{_{ref}}}\right) +\gamma _{2}\,\delta _{_{ref}}f\left( \frac{R}{\delta _{_{ref}}}\right) }\right) + \notag \\ &&\,\frac{\lambda }{2}\left( \frac{\left( \gamma _{1}-\gamma _{2}\,\rho _{_{ref}}\right) \,h\left( \frac{r}{\delta _{_{ref}}}\right) }{\lambda \,h\left( \frac{R}{\delta _{_{ref}}}\right) +\gamma _{2}\,\delta _{_{ref}}\,f\left( \frac{R}{\delta _{_{ref}}}\right) }\right) ^{2}\ .\label{freenergy} \end{eqnarray} \subsection{Impossibility of a two-phase fluid in a nanotube} \begin{figure}[h] \begin{center} \includegraphics[width=6cm]{1.eps} \end{center} \caption{\textbf{Two-phase fluid in a nanotube}: The nanotube is simultaneously filled with two phases liquid and vapour of the same fluid. The two phases (a) and (b) are separated by a cylindric \emph{material} interface.} \label{Fig. 1} \end{figure} Let us consider a nanotube simultaneously filled with liquid and vapour of the same fluid; an interface appears between the liquid and vapour phases. By reason of symmetry the liquid-vapour interface is a material surface represented by a cylindrical surface with the same axis as the nanotube. The interface has a positive surface energy $\gamma _{lv}$ increasing the free energy of the fluid inside the nanotube. \\ \emph{First case:} \emph{domain} (a) \emph{is liquid and domain} (b) \emph{is vapour.} An approximation allows us to compare the energy of the only liquid phase and the energy of the two-phase fluid when the liquid is in contact with the nanotube wall. The free energies associated with the wall are approximatively equal. We have the relation $g_{l_{b}}(\rho )\equiv g_{v_{b}}(\rho )+g_{l_{b}}(\rho _{v_{b}})$ with $- g_{l_{b}}(\rho _{v_{b}})=P(\rho _{v_b})-P(\rho _{l_{b}})= \Pi (\rho _{l_b})$; the term $\Pi (\rho _{l_b})$ is called \emph{the disjoining pressure} relatively to the mother bulk $\rho_{l_b}$ \cite{Derjaguin,Gouin 6}. The difference between the free energy per unit length $E_1$ of the liquid phase and the free energy per unit length $E_2$ of the two-phase fluid is approximatively \begin{equation*} E_2-E_1=\pi \left( -e_{1}^{2}\,\Pi (\rho _{l_b})+2e_{1}\gamma _{lv}\right) , \end{equation* where $e_{1}$ denotes the radius of the domain (b) of vapour delimited by the interface. Consequently, the free energy for the two-phase fluid in the nanotube is smaller than for only the liquid phase if e_{1}\geq 2\gamma _{lv}/\Pi (\rho _{l_b})$.\newline \emph{Second case:} \emph{domain} (a) \emph{is vapour and domain} (b) \emph{is liquid.} We can also compare the energy of the only vapour phase and the energy of the two-phase fluid when the vapour is in contact with the nanotube wall. The free energies associated with the wall are approximatively equal. The difference between the free energy per unit length $E_3$ of the vapour phase and the free energy per unit length $E_4$ of the two-phase fluid is approximatively \begin{equation*} E_4-E_3=\pi \left( e_{2}^{2}\,\Pi (\rho _{l_b})+2e_{2}\gamma _{lv}\right) , \end{equation* where $e_{2}$ denotes the radius of the domain (b) of liquid delimited by the interface. Consequently, the free energy in the nanotube is smaller for the two-phase fluid than for the vapour phase if e_{2}\geq 2\gamma _{lv}/(-\Pi (\rho _{l_b})$.\newline As an example, we consider the case of water at $20^{\degree}$ Celsius, in \textbf{c.g.s.} units, the interfacial free energy $\gamma _{lv}=72$. If |\Pi (\rho _{l_b})|=10^{7}$ (or 10 atmospheres), corresponding to an important absolute value of the disjoining pressure, we obtain $e_{1}=e_{2}\geq R_0 = 14.4\times 10^{-6}=144$\thinspace\ nm$\, = 0.144\,\mu $m. Consequently, the nanotube is filled with only one phase if its radius verifies the inequality $R < R_0$. The limit radius $R_0$ corresponds to the radius of a microscopic tube. Consequently, we have just to compare the free energies of the liquid and the vapour phases filling up the nanotube. \subsection{Liquid phase in the nanotube} We consider the case when the fluid phase in the nanotube is liquid; the liquid density is close to $\rho _{l_b}$ (and $\rho _{l_b}\simeq \rho _{l}$). We choose $g_{l_{b}}$ as \emph{reference level of volume free energy}. By taking account of Eq. (\ref{freenergy}) and $\delta _{l}=\sqrt \lambda \rho _{l}/c_{l}^{2}}$, we get \begin{equation} \phi(r)=\frac{c_{l}^{2}}{2\rho _{l}}\left( \frac{\delta _{l}\,\left( \gamma _{1}-\gamma _{2}\,\rho _{l_b}\right) f\left( \frac{r}{\delta _{l}}\right) } \lambda \,h\left( \frac{R}{\delta _{l}}\right) +\gamma _{2}\,\delta _{l}f\left( \frac{R}{\delta _{l}}\right) }\right) ^{2}+ \frac{\lambda }{2 \left( \frac{\left( \gamma _{1}-\gamma _{2}\,\rho _{l_b}\right) h\left( \frac{r}{\delta _{l}}\right) }{\lambda \,h\left( \frac{R}{\delta _{l} \right) +\gamma _{2}\,\delta _{l}\,f\left( \frac{R}{\delta _{l}}\right) \right) ^{2}. \label{freenergy1} \end{equation Due to the fact that $c_{l}^{2}\delta _{l}^{2}/2\rho _{l}=\lambda /2$, Eq. \ref{freenergy1}) reads \begin{equation*} \phi (r)=\frac{\lambda }{2}\,\left( \gamma _{1}-\gamma _{2}\,\rho _{l_b}\right) ^{2}\frac{f\left( \frac{r}{\delta _{l}}\right) ^{2}+h\left( \frac{r}{\delta _{l}}\right) ^{2}}{\left( \lambda \,h\left( \frac{R}{\delta _{l}}\right) +\gamma _{2}\,\delta _{l}\,f\left( \frac{R}{\delta _{l}}\right) \right) ^{2}}\ . \end{equation* The total free energy per unit of length in the nanotube is: \begin{equation*} E_{l_b}(\rho _{l_b})=2\pi \left( \int_{0}^{R}\phi (r)\,r\,dr+R\left( -\gamma _{1}+\frac{\gamma _{2}}{2}\,\rho _{_{R}}\right) \rho _{_{R}}\right) , \end{equation* where \begin{equation*} \rho _{_{R}}=\rho _{l_b}+\frac{\delta _{l}\,\left( \gamma _{1}-\gamma _{2}\,\rho _{l_b}\right) }{\gamma _{2}\,\delta _{l}\,f\left( \frac{R}{\delta _{l}}\right) +\lambda \,h\left( \frac{R}{\delta _{l}}\right) }\,f\left( \frac{R}{\delta _{l}}\right) . \end{equation* Consequently, if we denote $r=\delta _{l}\,x$ and $n=R/\,\delta _{l}$, we obtain the total free energy per unit of surface in the nanotube $F_{l_b}(\rho _{l_b})\equiv E_{l_b}(l)/2\pi R$ in the form \begin{equation} F_{l_b}(\rho _{l_b})=\frac{\delta _{l}c_{l}^{2}}{2\rho _{l}}\frac{\left( \gamma _{1}-\gamma _{2}\,\rho _{l_b}\right) ^{2}}{n}{{\int_{0}^{n}}}\frac{f\left( x\right) ^{2}+h\left( x\right) ^{2}}{\left( \gamma _{2}\,f\left( n\right) \frac{\lambda }{\delta _{l}}\,h\left( n\right) \right) ^{2}}\,x\,dx+\left( -\gamma _{1}+\frac{\gamma _{2}}{2}\,\rho _{_{R}}\right) \rho _{_{R}}. \label{free energy1} \end{equation} \subsection{Vapour in the nanotube} We consider the case when the fluid phase in the nanotube is vapour; the vapour density is close to $\rho _{v_{b}}$ (and $ \rho _{v_{b}}\simeq \rho _{v}$). For the reference level of volume free energy we obtain for a density close to $\rho _{v}$ \begin{equation*} g_{l_{b}}(\rho )={\frac{c_{v}^{2}}{2\rho _{v}}\left( \rho -\rho _{v}\right) ^{2}}-\Pi(\rho _{l_b}). \end{equation* The density in the nanotube is close to the vapour density $\rho _{v_{b}}$ and consequently we neglect the surface free energy of the wall. With \gamma _{1}-\gamma _{2}\,\rho _{v_{b}}\approx \gamma _{1}$ we get \begin{equation*} \phi (r)\approx \psi (r)-\Pi (\rho _{l_b})\ \RIfM@\expandafter\text@\else\expandafter\mbox\fi{\ with}\ \ \ \psi (r)=\frac \lambda }{2}\,\gamma _{1}^{2}\ \frac{f\left( \frac{r}{\delta _{v}}\right) ^{2}+h\left( \frac{r}{\delta _{v}}\right) ^{2}}{\left( \lambda \,h\left( \frac{R}{\delta _{v}}\right) +\gamma _{2}\,\delta _{v}\,f\left( \frac{R} \delta _{v}}\right) \right) ^{2}}\ , \end{equation*} where $\delta _{v}=\sqrt{\lambda \rho _{v}/c_{v}^{2}}$. The total free energy per unit of length in the nanotube is: \begin{equation*} E_{l_b}(\rho {_{v_{b}}})=2\pi \left( \int_{0}^{R}\psi (r)\,r\,dr-\Pi (\rho _{l_b} \frac{R^{2}}{2}\right). \end{equation* If we denote $r=\delta _{v}\,x$ and $N=R/\delta _{v}$, we obtain the total free energy per unit of surface in the nanotube $F_{l_b}(\rho _{v_{b}})\equiv E_{l_b}(v)/2\, \pi\, R$ in the form \begin{equation} F_{l_b}(\rho {_{v_{b}}})=\frac{\delta _{v}\,c_{v}^{2}}{2\rho _{v}}\frac{\gamma _{1}^{2}}{N}{{\int_{0}^{N}}}\frac{f\left( x\right) ^{2}+g\left( x\right) ^{2 }{\left( \gamma _{2}\,f\left( N\right) +\frac{\lambda }{\delta _{v}} \,g\left( N\right) \right) ^{2}}\,x\,dx\ -\frac{N\,\delta _{v}}{2}\,\Pi (\rho _{l_b})\, . \label{free energy2} \end{equation} \subsection{Numerical application in the case of water} \begin{figure}[h] \begin{center} \includegraphics[width=9cm]{2.eps} \end{center} \caption{\textbf{Nanotube of diameter 4 nm}: When we change the value of the disjoining pressure $\Pi(\protect\rho _{l_b})$, the total free energy in a nanotube filled up with liquid water phase is smaller than the total free energy of a nanotube filled up with vapour water phase. Consequently, the nanotube is always filled up with a liquid water phase.} \label{Fig. 2} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[width=9cm]{3.eps} \end{center} \caption{\textbf{Nanotube of diameter 20 nm}: When the disjoining pressure is positive and strong enough and if the wall is strongly hydrophobic, the nanotube can be filled up with a water vapour phase. For an hydrophillic wall the nanotube is always filled up with a liquid water phase.} \label{Fig. 3} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[width=9cm]{4.eps} \end{center} \caption{\textbf{Nanotube of diameter 100 nm}: The tube corresponding to a diameter of about $0.1 \,\mu$m is a microtube. The quality of the wall and the effect of the disjoining pressure are in competition. If the disjoining pressure is strong enough and if the wall is moderately hydrophobic, the tube is filled up with vapour water phase. When the disjoining pressure is negative, the tube is always filled up with liquid water phase.} \label{Fig. 4} \end{figure} In \textbf{c.g.s.} units the physical constants of water are\ $c_{v}=3.7\ \times 10^{4};\ c_{l}=1.478\ \times 10^{5};\ \rho _{v}=1.7\ \times 10^{-6}; \rho _{l}=0.998;\lambda =1.17\ \times 10^{-5};\ \gamma _{2}=54.$\newline We have obtained the free energy values for liquid and vapour phases. Relations (\ref{free energy1}-\ref{free energy2}) depend on both the wetting quality of the wall and on the value of the disjoining pressure. For convenient materials, we can numerically compare the free energy of the liquid with the free energy of the vapour. We consider the case when the water fluid is in contact with different nanotube walls. The $x$-axis corresponds to the value of $\gamma_1$ associated with the hydrophobicity or hydrophillicity of the wall. The value of $\gamma_2$ depends only on the fluid. When $\gamma_1$ is small enough, the wall of the nanotube is hydrophobic and when $\gamma_1$ is large enough, the wall of the nanotube is hydrophillic. The $y$-axis corresponds to the value of the total free energy in the nanotube per unit surface of the wall. The case when the nanotube wall is strongly hydrophobic corresponds to $\gamma_1 < 30$. In all the graphs the straight line parallel to the $x -axis corresponds to the free energy per unit surface of the nanotube filled up with a vapour phase; the oblique curve corresponds to the free energy per unit surface of the nanotube filled up with a liquid phase. The graphs corresponding to the values of the two free energies (\ref{free energy1}-\ref{free energy2}) allow to foresee if the nanotube is filled up with liquid or with vapour. They are presented in Figures 2 to 4.\newline When the disjoining pressure is negative, the nanotube is filled up with a liquid water phase for all diameters of the nanotube. When the nanotube wall is hydrophobic, for a large radius and a strong positive disjoining pressure, the nanotube can be filled up with a vapour phase. As a result, the case of vapour phase filling up the nanotube is less usual than the case of liquid water phase.\newline In all the cases, we note that the volume free energy of the phase in the nanotube is negligible with respect to the surface free energy of the wall. \section{Permanent viscous motions in a nanotube} "Fluid flow through nanoscopic structures, such as carbon nanotubes, is very different from the corresponding flow through microscopic and macroscopic structures. For example, the flow of fluids through nanomachines is expected to be fundamentally different from the flow through large-scale machines since, for the latter flow, the atomistic degrees of freedom of the fluids can be safely ignored, and the flow in such structures can be characterized by viscosity, density and other bulk properties. Furthermore, for flows through large-scale systems, the no-slip boundary condition is often implemented, according to which the fluid velocity is negligibly small at the fluid/wall boundary. Reducing the length scales immediately introduces new phenomena into the physics of the problem, in addition to the fact that at nanoscopic scales the motion of both the walls and the fluid, and their mutual interaction, must be taken into account" \cite{Rafii}. In this section, we consider the permanent and laminar motions of viscous capillary liquid in a nanotube. Because the liquid is heterogeneous, for capillary fluids, the liquid stress tensor is not anymore scalar and the equations of hydrodynamics are not valid. However, the results obtained for viscous flows \cite{Landau} can be adapted at nanoscale. \newline As in \cite{Rocard}, we assume that the kinematic viscosity coefficient $\nu =\kappa /\rho $ only depends on the temperature. In the equation of motions, the viscosity term is \begin{equation*} \ ({1}/{\rho })\,\RIfM@\expandafter\text@\else\expandafter\mbox\fi{div }\mathbf{\sigma }_{v}=2\nu \,\left[ \ \RIfM@\expandafter\text@\else\expandafter\mbox\fi{div } \bm D}\,+\,{\bm D}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ grad \{\thinspace Ln}\,(2\,\kappa )\}\ \right] , \end{equation* where ${\bm D}$ is the velocity deformation tensor and ${\bm D}$ grad\{Ln ( 2\,\kappa $)\} is negligible with respect to div\thinspace ${\bm D}$. \newline We denote the velocity by $\mathbf{V}=(0,0,w)^{T}$ where $w$ is the velocity component in direction of the nanotube axis. When we neglect the external forces (as gravity), the liquid nano-motion verifies Eq. (\ref{viscous motions}) written in the form \begin{equation} {\mathbf{a}}+\RIfM@\expandafter\text@\else\expandafter\mbox\fi{grad}[\,\mu _{o}(\rho )-\lambda \,\Delta \rho \,]=\nu \,\Delta {\mathbf{V}}\ \ \mathrm{with}\ \ \Delta {\mathbf{V}}\simeq \begin{bmatrix} 0,0,\Delta \end{bmatrix} ^T . \label{motion0} \end{equation \newline Simple fluids slip on a solid wall only at a molecular level \cite{Churaev} and consequently, in classical conditions, the kinematic condition at solid wall is the adherence condition $(z=0\;\Rightarrow \;w=0)$. Recent papers in nonequilibrium molecular dynamic simulations of three dimensional micro-Poiseuille flows in Knudsen regime reconsider micro-channels: the influence of gravity force, surface roughness, surface wetting condition and wall density are investigated. The results point out that the no-slip condition can be observed for Knudsen flow when the surface is rough. The roughness is a dominant parameter when the slip of fluid is concerned. The surface wetting condition substantially influences the velocity profiles \cite{Tabeling}. But it is not the case for smooth surfaces. The relation between wall shear stress and slip velocity is the key for characterizing the slip flow. With water flowing through hydrophobic thin capillaries, there are some qualitative evidences for slippage \cite{Blake}. De Gennes said: "the results are unexpected and stimulating and led us to think about unusual processes which could take place near a wall. They are connected with the thickness of the film when the thickness is of an order of the mean free path" \cite{De Gennes}.\newline When the free mean path $L$ is smaller than $d$, the Knudsen number \emph{Kn} is smaller than 1. That is the case for liquid where the mean free path $L$ is of the same order than the molecular diameter. For example in the case of liquid water, \emph{Kn} ranges between $0.5$ and $10^{-2}$ while the nanotube radius ranges between $1$ nm and $50$ nm. The adherence boundary condition at a surface, commonly employed with the Navier-Stokes equation assuming a zero flow, is physically invalid and a slip regime occurs; the boundary condition must be changed to take account of the slippage at the solid surfaces. \newline For gases, the mean free path is of order of one hundred molecular diameters and consequently the flow regime is only valid for large nanotubes. For thin nanotubes the rarefied gas regime must be considered; but the calculation in slip regime may give an idea of the change of flow with respect to the Navier-Stokes regime also for gases. Nonetheless, we note that the vapour phase in the tube occurs for large nanotube relevant from the microfluidic case and the slip condition using continuum mechanics is realistic. In fluid/wall slippage, the condition at solid wall writes \begin{equation} w=L_{s}\frac{\partial w}{\partial r}\qquad {\mathrm{ at}}\ \ r=R, \label{slip} \end{equation where $L_{s}$ is the so-called Navier length \cite{Landau}. The Navier length is expected to be independent of the thickness of the nanoflow and may be as large as a few microns \cite{Tabeling}. In the following, the dynamics of liquid nanoflows is studied in the case of nonrough nanotubes. Consequently,\newline \emph{i)} The equation of motion writes in the form (\ref{motion0}), \newline \emph{ii)} The boundary conditions take account of the slip condition (\re {slip}). \newline When the liquid nanoflow thickness is small with respect to transverse dimensions of the wall, it is possible to simplify Eq. (\ref{motion0}) which governs the viscous flow; so, when $d\ll \ell $, \newline \textit{iii) }We consider a laminar flow: the velocity component along the wall is large with respect to the normal velocity component to the wall which is negligible.\newline \textit{iv) }For permanent motion, the equation of continuity reads: \begin{equation*} \left(\mathtt{grad}\,\rho\right)^T \mathbf{V}+\rho _{\ }\mathtt{div}\ \mathbf{V}=0. \end{equation* The velocity vector $\mathbf{V}$ mainly varies along the direction orthogonal to the wall and grad\thinspace $\rho $ is normal to $\mathbf{V}$. The density is constant along each stream line ($\overset{\mathbf \centerdot }}{\rho }=0\Longleftrightarrow \mathtt{div}\,\mathbf{V}=0$); the trajectories are drawn on isodensity surfaces and $w=w(r)$. Due to the solid wall effect, the density in the tube is closely constant out from a boundary layer of approximatively one nanometer \cite{Gouin 6}; consequently, we consider the approximation of an incompressible liquid in the tube.\newline \emph{v)} Due to the geometry of the tube, for permanent motion, the acceleration is null. Equations (\ref{viscous motions}) or (\ref{motion0}) separate as: $\bullet \ \ $ The first part along the $z$-coordinate, \begin{equation} \frac{\partial P}{\partial z}=\kappa \,\Delta \,w\qquad \mathrm{with}\qquad \Delta \,w=\frac{1}{r}\frac{d}{dr}\left( r\ \frac{dw}{dr}\right), \label{axequ} \end{equation} \indent $\bullet \ \ $ The second part in the plane orthogonal to the tube axis, \begin{equation} \frac{\partial }{\partial r}\,(\lambda \,\Delta \rho -\mu _{_{0}})=0. \label{tgpart} \end{equation Equation (\ref{tgpart}) yields the same equation (\ref{densityequ}) as at equilibrium. Equation (\ref{axequ}) yields \begin{equation} \frac{1}{r}\frac{d}{dr}\left( r\ \frac{dw}{dr}\right) =-\frac{\wp }{\kappa}\,, \label{axequ1} \end{equation where $\wp $ denotes the pressure gradient along the nanotube. The cylindrical symmetry of the nanotube yields the solution of Eq. (\ref{axequ1}) in the form \begin{equation*} w=-\frac{\wp }{\kappa }\,\frac{r^{2}}{4}+b, \end{equation*} where $b$ is constant. Condition (\ref{slip}) implies \begin{equation*} -\frac{\wp }{4\kappa}R^{2}+b=L_{s}\frac{\wp }{2\kappa }R \end{equation* and consequently, \begin{equation*} w=\frac{\wp }{4\kappa }\left( -r^{2}+R(R+L_{s})\right). \end{equation* The density in the nanotube is closely equal to $\rho _{l}$ and the volume flow through the nanotube is $Q=2\pi \displaystyle\int_{0}^{R}w\,r\,dr=\frac{\pi\,\wp } 8\,\kappa }\,R^{3}(R+4L_{s})$. With $Q_{o}$ denoting the Poiseuille flow corresponding to a tube of the same radius $R$, we obtain: \begin{equation} Q=Q_{o}\left( 1+\frac{4L_{s}}{R}\right). \label{flow} \end{equation In most cases, the Navier length is of the micron order ( L_{s}=1\,\mu m=10^{3}\,$nm) \cite{Majumder}. If we consider a nanotube with $R=2\, nm, we obtain $Q=2\times 10^{3}\,Q_{o}$. For $R=50\,$nm, that we consider as the maximum radius of nanotube with respect to microfluidics, we obtain $Q=40\,Q_{o}$. Consequently, the flow of liquid in nanofluidics is dramatically more important than the Poiseuille flow in cylindrical tubes. In the case of gases, we obtain the same results for nanotube of radius $R=50\,$nm corresponding to a Knudsen number smaller than $0.5$. For nanotubes of radius smaller than $30\,$nm, when the nanotube is unusually filled up with a vapour phase, the flow is not anymore a continuous flow but is relevant to kinetic of rarified gases. Nevertheless, the magnitude of this flow is of several order more important than Poiseuille flow.\newline We have to emphasis that, when the mother bulk is vapour, in classical Poiseuille flow, the phase is vapour in the tube but, for a nanotube the phase is generally liquid (as in conditions presented in Fig. 2 and Fig. 3) and the volume flow through the nanotube is approximatively: \begin{equation*} Q=Q_{o}\frac{\rho _{l}}{\rho _{v}}\left( 1+\frac{4L_{s}}{R}\right). \label{flow2} \end{equation* In the case of water $\rho _{l}/ \rho _{v}\sim 10^3$, we get a volume flow $10^{3}$ time more important than the volume flow obtained in Eq. (\re {flow}): \begin{equation*} Q \sim 10^{6}\,Q_{o}\,. \end{equation* \section{Conclusion} A nanotube with a diameter ranging between 4 and 100 nanometres is filled up only with one liquid phase or one vapour phase independently of the external mother bulk. For nanotubes with diameters smaller than 20 nm, the fluid phase inside the nanotube is generally liquid. For nanotubes of large diameters with respect to the molecular scale, the fluid phase can be liquid or vapour according to the values of the disjoining pressure and of the physicochemical properties of the tube walls.\newline For nanotubes with small diameters, the flows can be significantly greater than usual Poiseuille flows, especially if the mother bulk consists of vapour.\newline These results, obtained by using a nonlinear model of continuum mechanics and its associated differential equations, are in good agreement with experiments and molecular dynamics calculations. \\ \noindent \textbf{Acknowledgements:} The paper has been supported by the XVI Conference on Waves and Stability in Continuous Media and Italian \emph{Gruppo Nazionale per la Fisica Matematica}. \section*{Abstract (Not appropriate in this style!)}% \else \small \begin{center}{\bf Abstract\vspace{-.5em}\vspace{\z@}}\end{center}% \quotation \fi }% }{% }% \@ifundefined{endabstract}{\def\endabstract {\if@twocolumn\else\endquotation\fi}}{}% \@ifundefined{maketitle}{\def\maketitle#1{}}{}% \@ifundefined{affiliation}{\def\affiliation#1{}}{}% \@ifundefined{proof}{\def\proof{\noindent{\bfseries Proof. }}}{}% \@ifundefined{endproof}{\def\endproof{\mbox{\ \rule{.1in}{.1in}}}}{}% \@ifundefined{newfield}{\def\newfield#1#2{}}{}% \@ifundefined{chapter}{\def\chapter#1{\par(Chapter head:)#1\par }% \newcount\c@chapter}{}% \@ifundefined{part}{\def\part#1{\par(Part head:)#1\par }}{}% \@ifundefined{section}{\def\section#1{\par(Section head:)#1\par }}{}% \@ifundefined{subsection}{\def\subsection#1% {\par(Subsection head:)#1\par }}{}% \@ifundefined{subsubsection}{\def\subsubsection#1% {\par(Subsubsection head:)#1\par }}{}% \@ifundefined{paragraph}{\def\paragraph#1% {\par(Subsubsubsection head:)#1\par }}{}% \@ifundefined{subparagraph}{\def\subparagraph#1% {\par(Subsubsubsubsection head:)#1\par }}{}% \@ifundefined{therefore}{\def\therefore{}}{}% \@ifundefined{backepsilon}{\def\backepsilon{}}{}% \@ifundefined{yen}{\def\yen{\hbox{\rm\rlap=Y}}}{}% \@ifundefined{registered}{% \def\registered{\relax\ifmmode{}\r@gistered \else$\m@th\r@gistered$\fi}% \def\r@gistered{^{\ooalign {\hfil\raise.07ex\hbox{$\scriptstyle\rm\RIfM@\expandafter\text@\else\expandafter\mbox\fi{R}$}\hfil\crcr \mathhexbox20D}}}}{}% \@ifundefined{Eth}{\def\Eth{}}{}% \@ifundefined{eth}{\def\eth{}}{}% \@ifundefined{Thorn}{\def\Thorn{}}{}% \@ifundefined{thorn}{\def\thorn{}}{}% \def\TEXTsymbol#1{\mbox{$#1$}}% \@ifundefined{degree}{\def\degree{{}^{\circ}}}{}% \newdimen\theight \def\Column{% \vadjust{\setbox\z@=\hbox{\scriptsize\quad\quad tcol}% \theight=\ht\z@\advance\theight by \dp\z@\advance\theight by \lineskip \kern -\theight \vbox to \theight{% \rightline{\rlap{\box\z@}}% \vss }% }% }% \def\qed{% \ifhmode\unskip\nobreak\fi\ifmmode\ifinner\else\hskip5\p@\fi\fi \hbox{\hskip5\p@\vrule width4\p@ height6\p@ depth1.5\p@\hskip\p@}% }% \def\cents{\hbox{\rm\rlap/c}}% \def\miss{\hbox{\vrule height2\p@ width 2\p@ depth\z@}}% \def\vvert{\Vert \def\tcol#1{{\baselineskip=6\p@ \vcenter{#1}} \Column} % \def\dB{\hbox{{}} \def\mB#1{\hbox{$#1$} \def\nB#1{\hbox{#1} \def\note{$^{\dag}}% \defLaTeX2e{LaTeX2e} \def\chkcompat{% \if@compatibility \else \usepackage{latexsym} \fi } \ifx\fmtnameLaTeX2e \DeclareOldFontCommand{\rm}{\normalfont\rmfamily}{\mathrm} \DeclareOldFontCommand{\sf}{\normalfont\sffamily}{\mathsf} \DeclareOldFontCommand{\tt}{\normalfont\ttfamily}{\mathtt} \DeclareOldFontCommand{\bf}{\normalfont\bfseries}{\mathbf} \DeclareOldFontCommand{\it}{\normalfont\itshape}{\mathit} \DeclareOldFontCommand{\sl}{\normalfont\slshape}{\@nomath\sl} \DeclareOldFontCommand{\sc}{\normalfont\scshape}{\@nomath\sc} \chkcompat \fi \def\alpha{{\Greekmath 010B}}% \def\beta{{\Greekmath 010C}}% \def\gamma{{\Greekmath 010D}}% \def\delta{{\Greekmath 010E}}% \def\epsilon{{\Greekmath 010F}}% \def\zeta{{\Greekmath 0110}}% \def\eta{{\Greekmath 0111}}% \def\theta{{\Greekmath 0112}}% \def\iota{{\Greekmath 0113}}% \def\kappa{{\Greekmath 0114}}% \def\lambda{{\Greekmath 0115}}% \def\mu{{\Greekmath 0116}}% \def\nu{{\Greekmath 0117}}% \def\xi{{\Greekmath 0118}}% \def\pi{{\Greekmath 0119}}% \def\rho{{\Greekmath 011A}}% \def\sigma{{\Greekmath 011B}}% \def\tau{{\Greekmath 011C}}% \def\upsilon{{\Greekmath 011D}}% \def\phi{{\Greekmath 011E}}% \def\chi{{\Greekmath 011F}}% \def\psi{{\Greekmath 0120}}% \def\omega{{\Greekmath 0121}}% \def\varepsilon{{\Greekmath 0122}}% \def\vartheta{{\Greekmath 0123}}% \def\varpi{{\Greekmath 0124}}% \def\varrho{{\Greekmath 0125}}% \def\varsigma{{\Greekmath 0126}}% \def\varphi{{\Greekmath 0127}}% \def{\Greekmath 0272}{{\Greekmath 0272}} \def\FindBoldGroup{% {\setbox0=\hbox{$\mathbf{x\global\edef\theboldgroup{\the\mathgroup}}$}}% } \def\Greekmath#1#2#3#4{% \if@compatibility \ifnum\mathgroup=\symbold \mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}% \else \mathchar"#1#2#3#4% \fi \else \FindBoldGroup \ifnum\mathgroup=\theboldgroup \mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}% \else \mathchar"#1#2#3#4% \fi \fi} \newif\ifGreekBold \GreekBoldfalse \let\SAVEPBF=\pbf \def\pbf{\GreekBoldtrue\SAVEPBF}% \@ifundefined{theorem}{\newtheorem{theorem}{Theorem}}{} \@ifundefined{lemma}{\newtheorem{lemma}[theorem]{Lemma}}{} \@ifundefined{corollary}{\newtheorem{corollary}[theorem]{Corollary}}{} \@ifundefined{conjecture}{\newtheorem{conjecture}[theorem]{Conjecture}}{} \@ifundefined{proposition}{\newtheorem{proposition}[theorem]{Proposition}}{} \@ifundefined{axiom}{\newtheorem{axiom}{Axiom}}{} \@ifundefined{remark}{\newtheorem{remark}{Remark}}{} \@ifundefined{example}{\newtheorem{example}{Example}}{} \@ifundefined{exercise}{\newtheorem{exercise}{Exercise}}{} \@ifundefined{definition}{\newtheorem{definition}{Definition}}{} \@ifundefined{mathletters}{% \newcounter{equationnumber} \def\mathletters{% \addtocounter{equation}{1} \edef\@currentlabel{\arabic{equation}}% \setcounter{equationnumber}{\c@equation} \setcounter{equation}{0}% \edef\arabic{equation}{\@currentlabel\noexpand\alph{equation}}% } \def\endmathletters{% \setcounter{equation}{\value{equationnumber}}% } }{} \@ifundefined{BibTeX}{% \def\BibTeX{{\rm B\kern-.05em{\sc i\kern-.025em b}\kern-.08em T\kern-.1667em\lower.7ex\hbox{E}\kern-.125emX}}}{}% \@ifundefined{AmS}% {\def\AmS{{\protect\usefont{OMS}{cmsy}{m}{n}% A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}}{}% \@ifundefined{AmSTeX}{\def\AmSTeX{\protect\AmS-\protect\TeX\@}}{}% \ifx\ds@amstex\relax \message{amstex already loaded}\makeatother\endinpu \else \@ifpackageloaded{amstex}% {\message{amstex already loaded}\makeatother\endinput} {} \@ifpackageloaded{amsgen}% {\message{amsgen already loaded}\makeatother\endinput} {} \fi \let\DOTSI\relax \def\RIfM@{\relax\ifmmode}% \def\FN@{\futurelet\next}% \newcount\intno@ \def\iint{\DOTSI\intno@\tw@\FN@\ints@}% \def\iiint{\DOTSI\intno@\thr@@\FN@\ints@}% \def\iiiint{\DOTSI\intno@4 \FN@\ints@}% \def\idotsint{\DOTSI\intno@\z@\FN@\ints@}% \def\ints@{\findlimits@\ints@@}% \newif\iflimtoken@ \newif\iflimits@ \def\findlimits@{\limtoken@true\ifx\next\limits\limits@true \else\ifx\next\nolimits\limits@false\else \limtoken@false\ifx\ilimits@\nolimits\limits@false\else \ifinner\limits@false\else\limits@true\fi\fi\fi\fi}% \def\multint@{\int\ifnum\intno@=\z@\intdots@ \else\intkern@\fi \ifnum\intno@>\tw@\int\intkern@\fi \ifnum\intno@>\thr@@\int\intkern@\fi \int \def\multintlimits@{\intop\ifnum\intno@=\z@\intdots@\else\intkern@\fi \ifnum\intno@>\tw@\intop\intkern@\fi \ifnum\intno@>\thr@@\intop\intkern@\fi\intop}% \def\intic@{% \mathchoice{\hskip.5em}{\hskip.4em}{\hskip.4em}{\hskip.4em}}% \def\negintic@{\mathchoice {\hskip-.5em}{\hskip-.4em}{\hskip-.4em}{\hskip-.4em}}% \def\ints@@{\iflimtoken@ \def\ints@@@{\iflimits@\negintic@ \mathop{\intic@\multintlimits@}\limits \else\multint@\nolimits\fi \eat@ \else \def\ints@@@{\iflimits@\negintic@ \mathop{\intic@\multintlimits@}\limits\else \multint@\nolimits\fi}\fi\ints@@@}% \def\intkern@{\mathchoice{\!\!\!}{\!\!}{\!\!}{\!\!}}% \def\plaincdots@{\mathinner{\cdotp\cdotp\cdotp}}% \def\intdots@{\mathchoice{\plaincdots@}% {{\cdotp}\mkern1.5mu{\cdotp}\mkern1.5mu{\cdotp}}% {{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}% {{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}}% \def\RIfM@{\relax\protect\ifmmode} \def\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\RIfM@\expandafter\RIfM@\expandafter\text@\else\expandafter\mbox\fi@\else\expandafter\mbox\fi} \let\nfss@text\RIfM@\expandafter\text@\else\expandafter\mbox\fi \def\RIfM@\expandafter\text@\else\expandafter\mbox\fi@#1{\mathchoice {\textdef@\displaystyle\f@size{#1}}% {\textdef@\textstyle\tf@size{\firstchoice@false #1}}% {\textdef@\textstyle\sf@size{\firstchoice@false #1}}% {\textdef@\textstyle \ssf@size{\firstchoice@false #1}}% \glb@settings} \def\textdef@#1#2#3{\hbox{{% \everymath{#1}% \let\f@size#2\selectfont #3}}} \newif\iffirstchoice@ \firstchoice@true \def\Let@{\relax\iffalse{\fi\let\\=\cr\iffalse}\fi}% \def\vspace@{\def\vspace##1{\crcr\noalign{\vskip##1\relax}}}% \def\multilimits@{\bgroup\vspace@\Let@ \baselineskip\fontdimen10 \scriptfont\tw@ \advance\baselineskip\fontdimen12 \scriptfont\tw@ \lineskip\thr@@\fontdimen8 \scriptfont\thr@@ \lineskiplimit\lineskip \vbox\bgroup\ialign\bgroup\hfil$\m@th\scriptstyle{##}$\hfil\crcr}% \def\Sb{_\multilimits@}% \def\endSb{\crcr\egroup\egroup\egroup}% \def\Sp{^\multilimits@}% \let\endSp\endSb \newdimen\ex@ \ex@.2326ex \def\rightarrowfill@#1{$#1\m@th\mathord-\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$}% \def\leftarrowfill@#1{$#1\m@th\mathord\leftarrow\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill\mkern-6mu\mathord-$}% \def\leftrightarrowfill@#1{$#1\m@th\mathord\leftarrow \mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$}% \def\overrightarrow{\mathpalette\overrightarrow@}% \def\overrightarrow@#1#2{\vbox{\ialign{##\crcr\rightarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \let\overarrow\overrightarrow \def\overleftarrow{\mathpalette\overleftarrow@}% \def\overleftarrow@#1#2{\vbox{\ialign{##\crcr\leftarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \def\overleftrightarrow{\mathpalette\overleftrightarrow@}% \def\overleftrightarrow@#1#2{\vbox{\ialign{##\crcr \leftrightarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \def\underrightarrow{\mathpalette\underrightarrow@}% \def\underrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil $\crcr\noalign{\nointerlineskip}\rightarrowfill@#1\crcr}}}% \let\underarrow\underrightarrow \def\underleftarrow{\mathpalette\underleftarrow@}% \def\underleftarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil $\crcr\noalign{\nointerlineskip}\leftarrowfill@#1\crcr}}}% \def\underleftrightarrow{\mathpalette\underleftrightarrow@}% \def\underleftrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th \hfil#1#2\hfil$\crcr \noalign{\nointerlineskip}\leftrightarrowfill@#1\crcr}}}% \def\qopnamewl@#1{\mathop{\operator@font#1}\nlimits@} \let\nlimits@\displaylimits \def\setboxz@h{\setbox\z@\hbox} \def\varlim@#1#2{\mathop{\vtop{\ialign{##\crcr \hfil$#1\m@th\operator@font lim$\hfil\crcr \noalign{\nointerlineskip}#2#1\crcr \noalign{\nointerlineskip\kern-\ex@}\crcr}}}} \def\rightarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@ $#1\copy\z@\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\box\z@\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$} \def\leftarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@ $#1\mathord\leftarrow\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\copy\z@\mkern-2mu$}\hfill \mkern-6mu\box\z@$} \def\qopnamewl@{proj\,lim}{\qopnamewl@{proj\,lim}} \def\qopnamewl@{inj\,lim}{\qopnamewl@{inj\,lim}} \def\mathpalette\varlim@\rightarrowfill@{\mathpalette\varlim@\rightarrowfill@} \def\mathpalette\varlim@\leftarrowfill@{\mathpalette\varlim@\leftarrowfill@} \def\mathpalette\varliminf@{}{\mathpalette\mathpalette\varliminf@{}@{}} \def\mathpalette\varliminf@{}@#1{\mathop{\underline{\vrule\@depth.2\ex@\@width\z@ \hbox{$#1\m@th\operator@font lim$}}}} \def\mathpalette\varlimsup@{}{\mathpalette\mathpalette\varlimsup@{}@{}} \def\mathpalette\varlimsup@{}@#1{\mathop{\overline {\hbox{$#1\m@th\operator@font lim$}}}} \def\tfrac#1#2{{\textstyle {#1 \over #2}}}% \def\dfrac#1#2{{\displaystyle {#1 \over #2}}}% \def\binom#1#2{{#1 \choose #2}}% \def\tbinom#1#2{{\textstyle {#1 \choose #2}}}% \def\dbinom#1#2{{\displaystyle {#1 \choose #2}}}% \def\QATOP#1#2{{#1 \atop #2}}% \def\QTATOP#1#2{{\textstyle {#1 \atop #2}}}% \def\QDATOP#1#2{{\displaystyle {#1 \atop #2}}}% \def\QABOVE#1#2#3{{#2 \above#1 #3}}% \def\QTABOVE#1#2#3{{\textstyle {#2 \above#1 #3}}}% \def\QDABOVE#1#2#3{{\displaystyle {#2 \above#1 #3}}}% \def\QOVERD#1#2#3#4{{#3 \overwithdelims#1#2 #4}}% \def\QTOVERD#1#2#3#4{{\textstyle {#3 \overwithdelims#1#2 #4}}}% \def\QDOVERD#1#2#3#4{{\displaystyle {#3 \overwithdelims#1#2 #4}}}% \def\QATOPD#1#2#3#4{{#3 \atopwithdelims#1#2 #4}}% \def\QTATOPD#1#2#3#4{{\textstyle {#3 \atopwithdelims#1#2 #4}}}% \def\QDATOPD#1#2#3#4{{\displaystyle {#3 \atopwithdelims#1#2 #4}}}% \def\QABOVED#1#2#3#4#5{{#4 \abovewithdelims#1#2#3 #5}}% \def\QTABOVED#1#2#3#4#5{{\textstyle {#4 \abovewithdelims#1#2#3 #5}}}% \def\QDABOVED#1#2#3#4#5{{\displaystyle {#4 \abovewithdelims#1#2#3 #5}}}% \def\tint{\mathop{\textstyle \int}}% \def\tiint{\mathop{\textstyle \iint }}% \def\tiiint{\mathop{\textstyle \iiint }}% \def\tiiiint{\mathop{\textstyle \iiiint }}% \def\tidotsint{\mathop{\textstyle \idotsint }}% \def\toint{\mathop{\textstyle \oint}}% \def\tsum{\mathop{\textstyle \sum }}% \def\tprod{\mathop{\textstyle \prod }}% \def\tbigcap{\mathop{\textstyle \bigcap }}% \def\tbigwedge{\mathop{\textstyle \bigwedge }}% \def\tbigoplus{\mathop{\textstyle \bigoplus }}% \def\tbigodot{\mathop{\textstyle \bigodot }}% \def\tbigsqcup{\mathop{\textstyle \bigsqcup }}% \def\tcoprod{\mathop{\textstyle \coprod }}% \def\tbigcup{\mathop{\textstyle \bigcup }}% \def\tbigvee{\mathop{\textstyle \bigvee }}% \def\tbigotimes{\mathop{\textstyle \bigotimes }}% \def\tbiguplus{\mathop{\textstyle \biguplus }}% \def\dint{\mathop{\displaystyle \int}}% \def\diint{\mathop{\displaystyle \iint }}% \def\diiint{\mathop{\displaystyle \iiint }}% \def\diiiint{\mathop{\displaystyle \iiiint }}% \def\didotsint{\mathop{\displaystyle \idotsint }}% \def\doint{\mathop{\displaystyle \oint}}% \def\dsum{\mathop{\displaystyle \sum }}% \def\dprod{\mathop{\displaystyle \prod }}% \def\dbigcap{\mathop{\displaystyle \bigcap }}% \def\dbigwedge{\mathop{\displaystyle \bigwedge }}% \def\dbigoplus{\mathop{\displaystyle \bigoplus }}% \def\dbigodot{\mathop{\displaystyle \bigodot }}% \def\dbigsqcup{\mathop{\displaystyle \bigsqcup }}% \def\dcoprod{\mathop{\displaystyle \coprod }}% \def\dbigcup{\mathop{\displaystyle \bigcup }}% \def\dbigvee{\mathop{\displaystyle \bigvee }}% \def\dbigotimes{\mathop{\displaystyle \bigotimes }}% \def\dbiguplus{\mathop{\displaystyle \biguplus }}% \def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}% \begingroup \catcode `|=0 \catcode `[= 1 \catcode`]=2 \catcode `\{=12 \catcode `\}=12 \catcode`\\=12 |gdef|@alignverbatim#1\end{align}[#1|end[align]] |gdef|@salignverbatim#1\end{align*}[#1|end[align*]] |gdef|@alignatverbatim#1\end{alignat}[#1|end[alignat]] |gdef|@salignatverbatim#1\end{alignat*}[#1|end[alignat*]] |gdef|@xalignatverbatim#1\end{xalignat}[#1|end[xalignat]] |gdef|@sxalignatverbatim#1\end{xalignat*}[#1|end[xalignat*]] |gdef|@gatherverbatim#1\end{gather}[#1|end[gather]] |gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]] |gdef|@gatherverbatim#1\end{gather}[#1|end[gather]] |gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]] |gdef|@multilineverbatim#1\end{multiline}[#1|end[multiline]] |gdef|@smultilineverbatim#1\end{multiline*}[#1|end[multiline*]] |gdef|@arraxverbatim#1\end{arrax}[#1|end[arrax]] |gdef|@sarraxverbatim#1\end{arrax*}[#1|end[arrax*]] |gdef|@tabulaxverbatim#1\end{tabulax}[#1|end[tabulax]] |gdef|@stabulaxverbatim#1\end{tabulax*}[#1|end[tabulax*]] |endgroup \def\align{\@verbatim \frenchspacing\@vobeyspaces \@alignverbatim You are using the "align" environment in a style in which it is not defined.} \let\endalign=\endtrivlist \@namedef{align*}{\@verbatim\@salignverbatim You are using the "align*" environment in a style in which it is not defined.} \expandafter\let\csname endalign*\endcsname =\endtrivlist \def\alignat{\@verbatim \frenchspacing\@vobeyspaces \@alignatverbatim You are using the "alignat" environment in a style in which it is not defined.} \let\endalignat=\endtrivlist \@namedef{alignat*}{\@verbatim\@salignatverbatim You are using the "alignat*" environment in a style in which it is not defined.} \expandafter\let\csname endalignat*\endcsname =\endtrivlist \def\xalignat{\@verbatim \frenchspacing\@vobeyspaces \@xalignatverbatim You are using the "xalignat" environment in a style in which it is not defined.} \let\endxalignat=\endtrivlist \@namedef{xalignat*}{\@verbatim\@sxalignatverbatim You are using the "xalignat*" environment in a style in which it is not defined.} \expandafter\let\csname endxalignat*\endcsname =\endtrivlist \def\gather{\@verbatim \frenchspacing\@vobeyspaces \@gatherverbatim You are using the "gather" environment in a style in which it is not defined.} \let\endgather=\endtrivlist \@namedef{gather*}{\@verbatim\@sgatherverbatim You are using the "gather*" environment in a style in which it is not defined.} \expandafter\let\csname endgather*\endcsname =\endtrivlist \def\multiline{\@verbatim \frenchspacing\@vobeyspaces \@multilineverbatim You are using the "multiline" environment in a style in which it is not defined.} \let\endmultiline=\endtrivlist \@namedef{multiline*}{\@verbatim\@smultilineverbatim You are using the "multiline*" environment in a style in which it is not defined.} \expandafter\let\csname endmultiline*\endcsname =\endtrivlist \def\arrax{\@verbatim \frenchspacing\@vobeyspaces \@arraxverbatim You are using a type of "array" construct that is only allowed in AmS-LaTeX.} \let\endarrax=\endtrivlist \def\tabulax{\@verbatim \frenchspacing\@vobeyspaces \@tabulaxverbatim You are using a type of "tabular" construct that is only allowed in AmS-LaTeX.} \let\endtabulax=\endtrivlist \@namedef{arrax*}{\@verbatim\@sarraxverbatim You are using a type of "array*" construct that is only allowed in AmS-LaTeX.} \expandafter\let\csname endarrax*\endcsname =\endtrivlist \@namedef{tabulax*}{\@verbatim\@stabulaxverbatim You are using a type of "tabular*" construct that is only allowed in AmS-LaTeX.} \expandafter\let\csname endtabulax*\endcsname =\endtrivlist \def\@@eqncr{\let\@tempa\relax \ifcase\@eqcnt \def\@tempa{& & &}\or \def\@tempa{& &}% \else \def\@tempa{&}\fi \@tempa \if@eqnsw \iftag@ \@taggnum \else \@eqnnum\stepcounter{equation}% \fi \fi \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@eqnswtrue \global\@eqcnt\z@\cr} \def\endequation{% \ifmmode\ifinner \iftag@ \addtocounter{equation}{-1} $\hfil \displaywidth\linewidth\@taggnum\egroup \endtrivlist \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@ignoretrue \else $\hfil \displaywidth\linewidth\@eqnnum\egroup \endtrivlist \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@ignoretrue \fi \else \iftag@ \addtocounter{equation}{-1} \eqno \hbox{\@taggnum} \global\@ifnextchar*{\@tagstar}{\@tag}@false% $$\global\@ignoretrue \else \eqno \hbox{\@eqnnum $$\global\@ignoretrue \fi \fi\fi } \newif\iftag@ \@ifnextchar*{\@tagstar}{\@tag}@false \def\@ifnextchar*{\@tagstar}{\@tag}{\@ifnextchar*{\@tagstar}{\@tag}} \def\@tag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@tagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}% } \makeatother \endinput
\section*{Help} } {} \ifdraft{}{\excludecomment{help}} \specialcomment{changes} { \section*{Changes} } {\listoftodos[TODO's etc.]} \ifdraft{}{\excludecomment{changes}} \specialcomment{detail} {\textcolor{blue!50!black}{/** }\begingroup\color{blue!50!black}} {\endgroup\textcolor{blue!50!black}{**/} } \ifdraft{}{\excludecomment{detail}} \newcommand{\old}[1]{\textcolor{red}{\sout{\ifdraft{#1}{}}}} \newcommand{\new}[1]{\ifdraft{\textcolor{green!50!black}{#1}}{#1}} \hypersetup{ pdfauthor={Sebastian Mayer, Tino Ullrich, and Jan Vyb\'iral}, pdfkeywords={ridge functions, entropy numbers, sampling, information-based complexity}, pdftitle={Approximation properties of ridge functions} } \begin{document} \title{Entropy and sampling numbers of classes of ridge functions} \author{Sebastian Mayer$^a$, Tino Ullrich$^{a,}$\footnote{Corresponding author. Email: <EMAIL>, Tel: +49 228 73 62224}, and Jan Vyb\'iral$^b$} \maketitle \vspace{-1cm} \begin{center} $^a$ Hausdorff-Center for Mathematics, Endenicher Allee 62, 53115 Bonn, Germany\\ $^b$ Department of Mathematics, Technical University Berlin, Stra{\ss}e des 17. Juni 136, 10623 Berlin, Germany\end{center} \begin{abstract} We study properties of ridge functions $f(x)=g(a\cdot x)$ in high dimensions $d$ from the viewpoint of approximation theory. The considered function classes consist of ridge functions such that the profile $g$ is a member of a univariate Lipschitz class with smoothness $\alpha > 0$ (including infinite smoothness), and the ridge direction $a$ has $p$-norm $\|a\|_p \leq 1$. First, we investigate entropy numbers in order to quantify the compactness of these ridge function classes in $L_{\infty}$. We show that they are essentially as compact as the class of univariate Lipschitz functions. Second, we examine sampling numbers and face two extreme cases. In case $p=2$, sampling ridge functions on the Euclidean unit ball faces the curse of dimensionality. It is thus as difficult as sampling general multivariate Lipschitz functions, a result in sharp contrast to the result on entropy numbers. When we additionally assume that all feasible profiles have a first derivative uniformly bounded away from zero in the origin, then the complexity of sampling ridge functions reduces drastically to the complexity of sampling univariate Lipschitz functions. In between, the sampling problem's degree of difficulty varies, depending on the values of $\alpha$ and $p$. Surprisingly, we see almost the entire hierarchy of tractability levels as introduced in the recent monographs by Novak and Wo\'zniakowski. \end{abstract} \begin{help} \begin{itemize} \item \new{Piece of text which is new in this version.} \item \old{Piece of text that should be removed for some reason.} \item \textcolor{blue!50!black}{/** more detailed descriptions; internal comments; $\dots$ **/} \item Remove the \texttt{draft} option to hide working draft annotations. \end{itemize} \end{help} \begin{changes} Changes to draft v11. \begin{itemize} \item Introduced the class of adaptive algorithms. Lower bounds in Section \ref{sec:sampling} now formulated in terms of adaptive algorithms. \item I checked the rules how to use commas in English language. Hopefully, everything is correct know. To get a better feeling, I additionally compared with text books written by native English speakers. \item At several points, I tried to make sentences more concise. \end{itemize} \end{changes} \section{Introduction} \old{Complex systems are nowadays ubiquitous in real-world applications. Quite often,}Functions depending on a large number (or even infinitely many) variables naturally appear in many real-world applications\old{when modelling these systems}. Since analytical representations are rarely available, there is a need to compute approximations to such functions or at least functionals thereof. Examples include parametric and stochastic PDEs \cite{CDMR, Schwab11}, data analysis and learning theory \cite{BvG,CZ,HTF}, quantum chemistry \cite{FHKS}, and mathematical finance \cite{PT}. It is a very well-known fact that approximation of smooth multivariate functions in many cases suffers from the so-called \emph{curse of dimensionality}. Especially, for fixed smoothness, the order of approximation decays rapidly with increasing dimension \cite{DL,lomavo96}. Actually, a recent result \cite{NW_2009_2} from the area of \emph{information-based complexity} states that on the unit cube, even uniform approximation of infinitely differentiable functions is intractable in high dimensions. These results naturally lead to the search for other assumptions than smoothness which would allow for tractable approximation, but would still be broad enough to include real-world applications. There are many different conditions of this kind. Usually, they require additional structure; for example, that the functions under consideration are tensor products or belong to some sort of weighted function space. We refer to \cite{NW1, TWW} and \cite{NW2} for a detailed discussion of (in)tractability of high-dimensional problems. In this work, we are interested in functions which take the form of a \emph{ridge}. This means that we look at functions where each $f$ is constant along lines perpendicular to some specific direction, say $a$. In other words, the function is of the form $f(x)=g(a\cdot x)$, where $g$ is a univariate function called the profile. Ridge functions provide a simple, coordinate-independent model, which describes inherently one-dimensional structures hidden in a high-dimensional ambient space. That the unknown functions take the form of a ridge is a frequent assumption in statistics, for instance, in the context of \emph{single index models}. For several of such statistical problems, minimax bounds have been studied on the basis of algorithms which exploit the ridge structure \cite{G,HJS,RWY}. Another point of view on ridge functions, which has attracted attention for more than 30 years, is to approximate \emph{by} ridge functions. An early work in this direction is \cite{LS75}, which took motivations from computerized tomography, and in which the term ``ridge function'' was actually coined. Another seminal paper is \cite{FrSt81}, which introduced \emph{projection pursuit regression} for data analysis. More recent works include the mathematical analysis of neural networks \cite{C2,Pinkus:NeuralNetworks}, and wavelet-type analysis \cite{CD}. For a survey on further approximation-theoretical results, we refer the reader to \cite{Pinkus:ApproxRidge}. For classical setups in statistics and data analysis, it is typical that we have no influence on the choice of sampling points. In contrast, problems of \emph{active learning} allow to \new{\emph{freely} choose} a limited number of samples from which to recover the function. Such a situation occurs, for instance, if sampling the unknown function at a point is realized by a (costly) PDE solver. In this context, ridge functions have appeared only recently as function models. The papers \cite{CT,CDDKP,FSV12} provide several algorithms and upper bounds for the approximation error. We continue in the direction of active learning, addressing two questions concerning the approximation of ridge functions. First, we ask how ``complex'' the classes of ridge functions are compared to uni- and multivariate Lipschitz functions. We measure complexity in terms of \emph{entropy numbers}, a classical concept in approximation theory. Second, we ask how hard it is to approximate ridge functions having only function values as information. Here, especially lower bounds are of interest to us. We formulate our results in terms of \emph{sampling numbers}. It should be pointed out, however, that we use a broader notion of sampling numbers than classical approximation theory does. As in the classical sense, we also consider a \emph{worst-case} setting with error measured in $L_{\infty}$. But sampling points may be chosen \emph{adaptively}. Both for entropy and sampling numbers, we consider classes of ridge functions defined on the $d$-dimensional Euclidean unit ball. These classes are characterized by three parameters: the profiles' order of Lipschitz smoothness $\alpha > 0$ (including infinite smoothness $\alpha=\infty$); a norm parameter $0 < p \leq 2$ indicating the $\ell_p^d$-ball in which ridge directions must be contained; and a parameter $0 \leq \kappa \leq 1$ to impose the restriction $|g'(0)| \geq \kappa$ on the first derivative of all feasible profiles $g$ (of course, this last parameter makes only sense in case of $\alpha > 1$). Regarding entropy numbers, the considered ridge function classes show a very uniform behaviour. For all possible parameter values, it turns out that they are essentially as compact as the class of univariate Lipschitz functions of the same order. For the sampling problem on the contrary, we find a much more diverse picture. On a first glance, the simple structure of ridge functions misleads one into thinking that approximating them should not be too much harder than approximating a univariate function. But this is far from true in general. Actually, in our specific setting, the sampling problem's degree of difficulty crucially depends on the constraint $|g'(0)| \geq \kappa$. If $\kappa > 0$, then it becomes possible to first recover the ridge direction efficiently. What remains then is only the one-dimensional problem of sampling the profile. In this scenario, the ridge structure indeed has a sweeping impact and the sampling problem is \emph{polynomially tractable}. But without the constraint on first derivatives and when all vectors in the domain may occur as ridge direction ($p=2$), sampling of ridge functions is essentially as hard as sampling of general Lipschitz functions over the same domain. It even suffers from the \emph{curse of dimensionality}, as long as we have only finite smoothness of profiles. For other configurations of the parameters $\alpha$ and $p$, the sampling problem's level of difficulty varies in between the extreme cases of polynomial tractability and curse of dimensionality. Surprisingly, we obtain almost the entire spectrum of degrees of tractability as introduced in the recent monographs by Novak and Wo\'zniakowski. The work is organized as follows. In Section \ref{sec:prelim}, we define the setting in a precise way and introduce central concepts. Section \ref{sec:entropy} then is dedicated to the study of entropy numbers for the considered function classes. Lower and upper bounds on sampling numbers are found in Section \ref{sec:sampling}. Finally, in Section \ref{sec:tractability}, we interpret our findings on sampling numbers in the language of information based-complexity. \section{Preliminaries} \label{sec:prelim} When $X$ denotes a (quasi-)Banach space of functions, equipped with the (quasi-)norm $\|\cdot\|_X$, we write $B_X = \{ f \in X: \; \|f\|_X < 1\}$ for the open unit ball and $\bar B_X$ for its closure. In case that $X=\ell_p^d(\field R) = (\field R,\|\cdot\|_p)$ we additionally use the notation $B_p^d$ for the open unit ball and $\ensuremath{\mathds S^{d-1}}_p$ for the unit sphere in $\ell_p^d$. \subsection{Ridge function classes} The specific form of ridge functions suggests to describe a class of such functions in terms of two parameters: one to determine the smoothness of profiles, the other to restrict the norm of ridge directions. Regarding smoothness, we require that ridge profiles are Lipschitz of some order. For the reader's convenience, let us briefly recall this notion. Let $\Omega\subset \field R^d$ be a bounded domain and $s$ be a natural number. The function space $C^{s}(\Omega)$ consists of those functions over the domain $\Omega$ which have partial derivatives up to order $s$ in the interior $\mathring{\Omega}$ of $\Omega$, and these derivatives are moreover bounded and continuous in $\Omega$. Formally, \[ C^s(\Omega) = \big\{ f: \Omega \to \field R: \quad \|f\|_{C^s} := \max_{ |\gamma| \leq s} \|D^{\gamma} f\|_{\infty} < \infty \big\}, \] where, for any multi-index $\gamma = (\gamma_1,\dots,\gamma_d) \in \N_0^d$, the partial differential operator $D^{\gamma}$ is given by \[ D^{\gamma}f := \frac{\partial^{|\gamma|} f}{\partial x_1^{\gamma_1} \cdots \partial x_d^{\gamma_d}}\,. \] Here we have written $|\gamma| = \sum_{i=1}^d \gamma_i$ for the order of $D^{\gamma}$. For the vector of first derivatives we use the usual notation $\nabla f = (\partial f/\partial x_1,\dots, \partial f/\partial x_d)$. Beside $C^s(\Omega)$ we further need the space of infinitely differentiable functions $C^{\infty}(\Omega)$ defined by \begin{equation}\label{f62} C^{\infty}(\Omega) = \big\{ f: \Omega \to \field R: \quad \|f\|_{C^{\infty}} := \sup_{ \gamma \in \N_0^d} \|D^{\gamma} f\|_{\infty} < \infty \big\}\,. \end{equation} For a function $f:\Omega\to\field R$ and any positive number $0 < \beta \leq 1$, the \emph{H\"older constant} of order $\beta$ is given by \begin{equation}\label{beta} |f|_{\beta}:=\sup_{\substack{x,y\in\Omega\\ x\not=y}}\frac{|f(x)-f(y)|}{2\min\{1, \|x-y\|_1\}^{\beta}}\;. \end{equation} This definition immediately implies the relation \begin{equation}\label{f64} |f|_{\beta} \leq |f|_{\beta'}\mbox{ if } 0<\beta < \beta' \leq 1. \end{equation} Now, for any $\alpha >0$, we can define the \emph{Lipschitz space} $\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$. If we let $s=\llfloor \alpha \rrfloor$ be the largest integer \emph{strictly less} than $\alpha$, it contains those functions in $C^s(\Omega)$ which have partial derivatives of order $s$ which are moreover H\"older-continuous of order $\beta = \alpha - s>0$. Formally, \[ \ensuremath{\mathrm{Lip}}_{\alpha}(\Omega) = \big\{ f \in C^s(\Omega): \quad \|f\|_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)} := \max \{ \|f\|_{C^s}, \ \max_{|\gamma| = s} |D^{\gamma}f|_{\beta} \} < \infty \big\}. \] For $s\in \N_0$ and $1 \geq \beta_2>\beta_1 > 0$ the following embeddings hold true \begin{equation}\label{f60} C^{\infty}(\Omega) \subset \ensuremath{\mathrm{Lip}}_{s+\beta_2}(\Omega) \subset \ensuremath{\mathrm{Lip}}_{s+\beta_1}(\Omega) \subset C^s(\Omega) \subset \ensuremath{\mathrm{Lip}}_s(\Omega)\,, \end{equation} where the respective identity operators are of norm one. In other words, the respective unit balls satisfy the same relation. Note that the fourth inclusion only makes sense if $s\geq 1$. The third embedding is a trivial consequence of the definition. The second embedding follows from the third, \old{the fourth,} and \eqref{f64}. The fourth embedding and the second imply the first. So it remains to establish the fourth embedding. We have to show that for every $\gamma \in \N_0^d$ with $|\gamma| = s-1$ it holds $|D^{\gamma}f|_1 \leq \|f\|_{C^s}$\,. On the one hand, Taylor's formula in $\field R^d$ gives for some $0<\theta<1$ \begin{equation}\nonumber \begin{split} |D^{\gamma}f(x)-D^{\gamma}f(y)| &= |\nabla (D^{\gamma}f)(x+\theta(y-x))\cdot (x-y)| \\ &\leq \max_{|\beta| = s}\|D^{\beta}f\|_{\infty}\cdot \|x-y\|_1\\ &\leq \|f\|_{C^s}\|x-y\|_1\,. \end{split} \end{equation} On the other hand, we have $|D^{\gamma}f(x)-D^{\gamma}f(y)| \leq 2\|f\|_{C^s}$\,. Both estimates together yield $|D^{\gamma}f|_1 \leq \|f\|_{C^s}$\,. Having introduced Lipschitz spaces, we can give a formal definition of our ridge functions classes. For the rest of the paper, we fix as function domain the closed unit ball $$ \Omega = \bar B_2^d = \{x\in\field R^d~:~\|x\|_2 \leq 1\}. $$ As before, let $\alpha > 0$ denote the order of Lipschitz smoothness. Further, let $0 < p \leq 2$. We define the class of ridge functions with Lipschitz profiles as \begin{align}\label{def_ridge} \Rs{\alpha,p} = \left\lbrace f: \Omega \to \field R \; : \; f(x) = g(a\cdot x), \; \|g\|_{\ensuremath{\mathrm{Lip}}_{\alpha}[-1,1]} \leq 1 ,\; \|a\|_p \leq 1 \right\rbrace. \end{align} In addition, we define the class of ridge functions with infinitely differentiable profiles by $$ \Rs{\infty,p} = \left\lbrace f: \Omega \to \field R \; : \; f(x) = g(a\cdot x), \; \|g\|_{C^{\infty}[-1,1]} \leq 1 ,\; \|a\|_p \leq 1 \right\rbrace. $$ \noindent Let us collect basic properties of these classes. \begin{lemma}\label{emb} For any $\alpha > 0$ and $0 < p \leq 2$ the class $\Rs{\alpha,p}$ is contained in $\bar B_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)}$ and $\Rs{\infty,p}$ is contained in $\bar B_{C^{\infty}(\Omega)}$. \end{lemma} \begin{proof} Let $f\in \Rs{\alpha,p}$ and $s=\llfloor\alpha\rrfloor$. Furthermore, let $\gamma \in \N_0^d$ be such that $|\gamma| \leq s$. Then, there exists $g\in \ensuremath{\mathrm{Lip}}_{\alpha}([-1,1])$ with $$ D^{\gamma}f(x) = D^{|\gamma|}g(a\cdot x)a^{\gamma}\,,\quad x\in \Omega\,, $$ where we used the convention $a^\gamma = \prod_{i=1}^d a_i^{\gamma_i}$\,. Therefore, we have $$ \|D^{\gamma} f\|_{\infty} \leq \|D^{|\gamma|} g\|_{\infty}\|a\|_{\infty}^{|\gamma|} \leq \|a\|_p^{|\gamma|} \leq 1\,. $$ If we let $s \to \infty$ this immediately implies $\Rs{\infty,p} \subset \bar B_{C^{\infty}(\Omega)}$. Moreover, if $|\gamma| = s$ and $\beta = \alpha-s$ we obtain by H\"older's inequality for $x,y \in \Omega$ \begin{equation}\nonumber \begin{split} |D^{\gamma}f(x)-D^{\gamma}f(y)| &= |a^{\gamma}|\cdot|D^sg(a\cdot x)-D^sg(a\cdot y)|\\ &\leq \|a\|_p^s\cdot |D^sg|_{\beta}\cdot 2\min\{1,\|a\|_p\cdot \|x-y\|_1\}^{\beta}\\ &\leq 2\min\{1,\|x-y\|_1\}^{\beta}\,. \end{split} \end{equation} Consequently, we have $\|f\|_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)} \leq 1$ and hence $\Rs{\alpha,p} \subset \bar B_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)}$. \end{proof} \noindent Note that in the special case $\alpha=1$, we have Lipschitz-continuous profiles. Whenever $0<\alpha_1 < \alpha_2 \leq \infty$, we have $\Rs{\alpha_2, p} \subset \Rs{\alpha_1,p}$, which is an immediate consequence of \eqref{f60}. Likewise, for $p < q$ we have the relation $\Rs{\alpha, p} \subset \Rs{\alpha, q}$. Finally, for Lipschitz smoothness $\alpha > 1$, we want to introduce a restricted version of $\Rs{\alpha,p}$, where profiles obey the additional constraint $|g'(0)| \geq \kappa > 0$. We define \begin{align} \Rs{\alpha,p,\kappa} = \{ g(a\cdot) \in \Rs{\alpha,p}: \quad |g'(0)| \geq \kappa\}. \end{align} Whenever we say in the sequel that we consider ridge functions with first derivatives bounded away from zero in the origin, we mean that they are contained in the class $\Rs{\alpha,p,\kappa}$ for some $0 < \kappa \le 1$. \paragraph{Taylor expansion.} We introduce a straight-forward, multivariate extension of Taylor's expansion on intervals to ridge functions in $\Rs{\alpha,p}$ and functions in $\ensuremath{\mathrm{Lip}}_\alpha(\Omega)$. For $x,x^0 \in \mathring{\Omega}$ we define the function $\Phi_x(\cdot)$ by $$ \Phi_x(t) := f(x^0+t(x-x^0))\,,\quad t\in [0,1]\,. $$ \begin{lemma} Let $\alpha>1$ and $\alpha = s+\beta$, $s\in \N$, $0<\beta\leq 1$. Let further $f\in \ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$ and $x,x^0 \in \mathring{\Omega}$. Then there is a real number $\theta \in (0,1)$ such that $$ f(x) = T_{s,x^0}f(x) + R_{s,x^0}f(x)\,, $$ where the Taylor polynomial $T_{s,x^0}f(x)$ is given by \begin{equation} T_{s,x^0}f(x) = \sum\limits_{j=0}^s \frac{\Phi_x^{(j)}(0)}{j!} = \sum\limits_{|\gamma| \leq s} \frac{D^{\gamma}f(x^0)}{\gamma!}(x-x^0)^{\gamma} \end{equation} and the remainder \begin{align}\label{remainder1} R_{s,x^0}f(x) &= \frac{1}{s!}\Big(\Phi_x^{(s)}(\theta)-\Phi_x^{(s)}(0)\Big)\\ \label{remainder2}&= \sum\limits_{|\gamma| = s} \frac{D^{\gamma}f(x^0+\theta(x-x^0))-D^{\gamma}f(x^0)}{\gamma!}(x-x^0)^{\gamma}\,. \end{align} \end{lemma} \noindent The previous lemma has a nice consequence for the approximation of functions from $\Rs{\alpha,p}$ in case $\alpha>1$ and $0<p\le 2$\,. Let $p'$ denote the dual index of $p$ given by $1/\max\{p,1\} + 1/p' = 1$. \begin{lemma}\label{lem:taylorapprox} Let $\alpha = s+\beta > 1$ and $\Omega = \bar B^d_2$.\\ {\em (i)} For $f\in \ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$ and $x, x^0 \in \mathring{\Omega}$ we have $$ |f(x) - T_{s,x^0}f(x)| \leq 2\|f\|_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)}\frac{\|x-x^0\|_{1}^{\alpha}}{s!}\,. $$ {\em (ii)} Let $0<p\leq 2$. Then for $f\in \Rs{\alpha,p}$ we have the slightly better estimate $$ |f(x) - T_{s,x^0}f(x)| \leq \frac{2}{s!}\|x-x^0\|_{p'}^{\alpha}\,. $$ \end{lemma} \begin{proof} To prove (i) we use \eqref{remainder2} and the definition of $\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$ and estimate as follows \begin{align}\nonumber |f(x) - T_{s,x^0}f(x)| &\leq \sum\limits_{|\gamma| = s} \frac{|D^{\gamma}f(x^0+\theta(x-x^0))-D^{\gamma}f(x^0)|}{\gamma!}|(x-x^0)^{\gamma}|\\ \nonumber&\leq 2\|f\|_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)}\min\{1,\|x-x^0\|_1\}^{\beta}\cdot \sum\limits_{|\gamma|=s} \frac{\prod\limits_{i=1}^d |x_i-x^0_i|^{\gamma_i}}{\gamma!}\,. \end{align} Using mathematical induction it is straight-forward to verify the multinomial identity $$ (a_1+\dots+a_d)^s = \sum\limits_{|\gamma| = s}\frac{s!}{\gamma!}a_1^{\gamma_1}\cdot\dots\cdot a_d^{\gamma_d}\,. $$ Hence, choosing $a_i = |x_i-x^0_i|$ we can continue estimating $$ |f(x) - T_{s,x^0}f(x)| \leq 2\|f\|_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)}\min\{1,\|x-x^0\|_1\}^{\beta} \frac{\|x-x^0\|_{1}^{s}}{s!} $$ and obtain the assertion in (i).\\ For showing the improved version (ii) for functions of type $f(x) = g(a\cdot x)$ we use formula \eqref{remainder1} of the Taylor remainder. We easily see that for $t\in (0,1)$ it holds $$ \Phi^{(s)}_{x}(t) = g^{(s)}\Big(a\cdot (x^0+t(x-x^0))\Big)\cdot [a\cdot (x-x^0)]^s\,. $$ Using H\"older continuity of $g^{(s)}$ of order $\beta$ and H\"older's inequality we see that \begin{align}\nonumber |f(x) - T_{s,x^0}f(x)| &\leq \frac{1}{s!}\Big|[a\cdot (x-x^0)]^s \cdot \Bigl\{g^{(s)}\Big(a\cdot (x^0+\theta(x-x^0))\Big)-g^{(s)}(a\cdot x^0)\Bigr\}\Big| \\ &\nonumber\leq \frac{1}{s!}\|a\|_p^s\cdot \|x-x^0\|^s_{p'}\cdot 2\min\{1, |\theta a\cdot (x-x^0)|^{\beta}\} \\ &\nonumber\leq \frac{2}{s!}\|x-x^0\|_{p'}^{\alpha}\,. \end{align} The proof is complete. \end{proof} \subsection{Information complexity and tractability} \label{sec:ibc} In this work, we want to approximate ridge functions from $\mathcal F = \Rs{\alpha,p}$ or $\mathcal F = \Rs{\alpha,p,\kappa}$ by means of deterministic sampling algorithms, using a limited amount of function values. Any allowed algorithm $S$ consists of an \emph{information map} $N_S^{\text{ada}}:{\mathcal F}\to \field R^n$, and a \emph{reconstruction map} $\varphi_S:\field R^n\to L_{\infty}(\Omega)$. The former provides, for $f \in \mathcal F$, function values $f(x_1),\dots,f(x_n)$ at points $x_1,\dots,x_n \in\Omega$, which are allowed to be chosen \emph{adaptively}. Adaptivity here means that $x_i$ may depend on the preceding values $f(x_1),\dots, f(x_{i-1})$. According to \cite{NW1}, we speak of \emph{standard information}. The reconstruction map then builds an approximation to $f$ based on those function values provided by the information map. Formally, we consider the class of deterministic, adaptive sampling algorithms $\ensuremath{\mathcal S^{\text{ada}}} = \bigcup_{n \in \N} \ensuremath{\mathcal S^{\text{ada}}}_n$, where \begin{equation*} \begin{split} \ensuremath{\mathcal S^{\text{ada}}}_n = \Big\{&S: \mathcal F \rightarrow L_{\infty}(\bar B_2^d)~:~\\ &S = \varphi_S \circ N^{\text{ada}}_S, \varphi: \field R^m \rightarrow L_{\infty}, \varphi(0)=0, \; N_S^{\text{ada}}: \mathcal F \to \field R^m, \; m \leq n \Big\}\,. \end{split} \end{equation*} Let us shortly comment on the restriction $\varphi(0) = 0$. Clearly, if $N_S(f) = 0$ for some $f$ then $\|f-S(f)\| = \|f\|$. Hence, such a function $f$ can never be well approximated by $S(f)$ since the error can not be smaller than $\|f\|$. Without the restriction $\varphi(0) = 0$ either the function $f$ or $-f$ is a bad function in this respect. Indeed, assume $N_S(f) = 0$ then $S(f) = S(-f) = \varphi_S(0) = b \in \field R$. Then \begin{align*} \| f\| &= 1/2 \| 2 f\| = 1/2 \| f - S(f) + f + S(f) \|\\ &\leq 1/2 \big\{ \| f - S(f)\| + \| -f - S(-f) \| \big\}\\ &\leq \max \big\{ \| f - S(f)\|, \| -f - S(-f) \| \big\}. \end{align*} For the given class of adaptive algorithms, the \emph{$n$-th minimal worst-case error} \[ g^{\text{ada}}_{n,d}(\mathcal F, L_{\infty}) := \err_{n,d}(\mathcal F, \ensuremath{\mathcal S^{\text{ada}}},L_{\infty}) = \inf \big\{\sup_{f \in \mathcal F} \|f-S(f)\|_{\infty} : S \in \ensuremath{\mathcal S^{\text{ada}}}_n \big\}, \] describes the approximation error which the best possible adaptive algorithm at most makes for a given budget of sampling points and any function from $\mathcal F$. Stressing that function values are the only available information, we refer to $g^{\text{ada}}_{n,d}(\mathcal F, L_{\infty})$ as the $n$-th \emph{ (adaptive) sampling number}. To reveal the effect of adaption, it is useful to compare adaptive algorithms with the subclass $\ensuremath{\mathcal S} \subset \ensuremath{\mathcal S^{\text{ada}}}$ of \emph{non-adaptive}, deterministic algorithms; that is, for each algorithm $S \in \ensuremath{\mathcal S}$ the information map is now of the form $N_S = (\delta_{x_1},\dots, \delta_{x_n})$, with $n \in \N$ and $x_1,\dots,x_n \in \bar B_2^d$. This corresponds to \emph{non-adaptive standard information} in \cite{NW1}. The associated $n$-th worst-case error \[ g_{n,d}(\mathcal F,L_{\infty}) := \inf_{S \in \mathcal S_n} \sup_{f \in \mathcal F} \norm{f - S(f)}_{\infty} = \err_{n,d}(\mathcal F, \mathcal S_n, L_{\infty}). \] coincides with the standard $n$-th \emph{sampling number} as known from classical approximation theory. As a third restriction, let us introduce the $n$-th \emph{linear} sampling number $g_{n,d}^{\text{lin}}(\mathcal F, L_{\infty})$; here, only algorithms from $\ensuremath{\mathcal S}$ with linear reconstruction map are allowed. Clearly, \[ g_{n,d}^{\text{ada}}(\mathcal F, L_{\infty}) \leq g_{n,d}(\mathcal F, L_{\infty}) \leq g_{n,d}^{\text{lin}}(\mathcal F, L_{\infty}). \] \begin{remark} There are results, see \cite[Section 4.2]{NW1}, which show that neither adaptivity nor non-linearity of algorithms does help under rather general conditions. These include two conditions which are certainly not met in our setting: (1) we only have function values as information, not general linear functionals; (2) the considered ridge function classes $\Rs{\alpha,p}$ and $\Rs{\alpha,p,\kappa}$ are not convex (however, they are at least symmetric). Nevertheless, the analysis in Section \ref{sec:sampling} reveals that in our setting, both adaptivity and non-linearity cannot lead to any substantial improvement in the approximation of ridge functions. \end{remark} Whenever we speak of sampling of ridge functions in the following, we refer to the problem of approximating ridge functions in $\mathcal F$ by sampling algorithms from $\ensuremath{\mathcal S^{\text{ada}}}$, the $L_{\infty}$-approximation error measured in the worst-case. Its \emph{information complexity} $n(\varepsilon,d)$ is given for $0<\varepsilon\le 1$ and $d\in\N$ by \[ n(\varepsilon,d) := \min \{n \in \N: g^{\text{ada}}_{n,d}(\mathcal F,L_{\infty}) \leq \varepsilon\}. \] \subsection{Entropy numbers} The concept of entropy numbers is central to this work. They can be understood as a measure to quantify the compactness of a set w.r.t.\ some reference space. For a detailed exposure and historical remarks, we refer to the monographs \cite{CaSt90,EdTr96}. The $k$-th entropy number $e_k(K,X)$ of a subset $K$ of a (quasi-)Banach space $X$ is defined as \begin{equation}\label{defi:entropy} e_k(K,X)=\inf\Big\{\varepsilon>0:K \subset \bigcup\limits_{j=1}^{2^{k-1}} (x_j+\varepsilon \bar B_X) \text{ for some } x_1,\dots,x_{2^{k-1}}\in X\Big\}\,. \end{equation} Note that $e_k(K,X) = \inf \{ \varepsilon >0: N_{\varepsilon}(K, X) \leq 2^{k-1}\}$ holds true, where \begin{align}\label{defi:covering_number} N_{\varepsilon}(K,X) := \min \Big\{n \in \N: \quad \exists x_1,\dots, x_n \in X: \; K \subset \bigcup_{j=1}^n (x_j + \varepsilon \bar B_X)\Big\} \end{align} denotes the \emph{covering number} of the set $K$ in the space $X$, which is the minimal natural number $n$ such that there is an $\varepsilon$-net of $K$ in $X$ of $n$ elements. We can introduce entropy numbers for operators, as well. The $k$-th entropy number $e_k(T)$ of an operator $T:X\to Y$ between two quasi-Banach spaces $X$ and $Y$ is defined by \begin{equation}\label{f1} e_k(T)=e_k(T(\bar B_X),Y). \end{equation} The results in Section \ref{sec:entropy} and \ref{sec:sampling} rely to a great degree on entropy numbers of the identity operator between the two finite dimensional spaces $X = \ell^d_p(\field R)$, and $Y = \ell^d_q(\field R)$. Thanks to \cite{Sch84, EdTr96, Tr97, Ku01}, their behavior is completely understood. For the reader's convenience, we restate the result. \begin{lemma}\label{lem:schuett} Let $0<p\leq q \leq \infty$ and let $k$ and $d$ be natural numbers. Then, $$ e_k(\bar B_p^d,\ell_q^d) \asymp \left\{ \begin{array}{rcl}1&:&1\leq k \leq \log(d),\\ \Big(\frac{\log(1+d/k)}{k}\Big)^{1/p-1/q}&:&\log(d)\leq k\leq d,\\ 2^{-\frac{k}{d}}d^{1/q-1/p}&:&k\geq d\,. \end{array} \right. $$ The constants behind ``$\asymp$'' do neither depend on $k$ nor on $d$. They only depend on the parameters $p$ and $q$. \end{lemma} If we consider entropy numbers of $\ell_p^d$-spheres instead of $\ell_p^d$-balls in $\ell_q^d$, the situation is quite similar. We are not aware of a reference where this has already been formulated thoroughly. \begin{lemma}\label{res:entropy_sphere} Let $d\in \N$, $d\geq 2$, $0<p\leq q \leq \infty$, and $\bar p = \min\{1,p\}$. Then, \begin{enumerate}[label=(\roman*)] \item \[ 2^{-k/(d-1)} d^{1/q-1/p} \; \lesssim \; e_k(\ensuremath{\mathds S^{d-1}}_p, \ell_q^d) \; \lesssim \; 2^{-k/(d-\bar p)} d^{1/q-1/p}, \quad k \geq d. \] \item $$ e_k(\ensuremath{\mathds S^{d-1}}_p,\ell_q^d) \asymp \left\{ \begin{array}{rcl}1&:&1\leq k \leq \log(d),\\ \Big(\frac{\log(1+d/k)}{k}\Big)^{1/p-1/q}&:&\log(d)\leq k\leq d. \end{array} \right. $$ \end{enumerate} The constants behind ``$\asymp$'' only depend on $p$ and $q$. \end{lemma} \begin{proof} For given $\varepsilon > 0$, an $\varepsilon$-covering $\{y_1,\dots,y_N\}$ of $\ensuremath{\mathds S^{d-1}}_p$ in $\ell_p^d$ fulfils \begin{align}\label{eq:entropy_sphere_1} (1+\varepsilon) \bar B_p^d \setminus (1-\varepsilon) \bar B_p^d \subseteq & \bigcup_{i=1}^N (y_i + 2^{1/\bar p} \varepsilon \bar B_p^d). \end{align} Let $\bar q = \min\{1,q\}$. For given $\varepsilon > 0$, a maximal set $\{x_1,\dots,x_M\} \subset \ensuremath{\mathds S^{d-1}}_p$ of vectors with mutual distance greater $\varepsilon$ obeys \begin{align}\label{eq:entropy_sphere_2} & \bigcup_{i=1}^M (x_i + 2^{-1/\bar q} \, \varepsilon \bar B_q^d) \subseteq (1 + \varepsilon_d^{\bar p})^{1/\bar p} \bar B_p^d \setminus (1 - \varepsilon_d^{\bar p})^{1/\bar p} \bar B_p^d, \end{align} where $\varepsilon_d = 2^{-1/\bar q} \, \varepsilon \, d^{1/p-1/q}$. \noindent\emph{(i).} A standard volume argument applied to \eqref{eq:entropy_sphere_1} yields $h(\varepsilon) \leq N \varepsilon^d 2^{d/\bar p}$, where $h(\varepsilon) = (1+\varepsilon)^d - (1-\varepsilon)^d$. First-order Taylor expansion in $\varepsilon$ allows to estimate $h(\varepsilon) \geq d\varepsilon$. Solving for $N$ yields a lower bound for covering numbers in case $p=q$. The lower bound in case $p\neq q$ follows from the trivial estimate $e_k(\ensuremath{\mathds S^{d-1}}_p, \ell_q^d) \geq d^{1/q-1/p} \; e_k(\ensuremath{\mathds S^{d-1}}_p,\ell_p^d)$. For the upper bound in case $p=q$ a standard volume argument applied to \eqref{eq:entropy_sphere_2} yields $M \varepsilon^d 2^{-d/\bar p} \leq h_{p}(\varepsilon^{\bar p}/2)$ with $h_{p}(x) = (1+x)^{d/\bar p} - (1-x)^{d/\bar p}$. The mean value theorem gives $h(x) \leq d/\bar p\, 2^{d/\bar p}x$ if $0<x\leq 1$. Hence, we get $h_{p}(\varepsilon^{\bar p}/2) \leq d/\bar p\, 2^{d/\bar p} \varepsilon^{\bar p}/2$. Solving for $M$ gives an upper bound for packing numbers and hence also for covering numbers. In case $p\neq q$ we again use \eqref{eq:entropy_sphere_2} and pass to volumes. This time the quotient $\mbox{vol}(B_p^d)/\mbox{vol}(B_q^d)$ remains in the upper bound for $M$. The given bounds now easily translate to the stated bounds on entropy numbers. In case $p\neq q$ one has to take $$ \Big[\frac{\mbox{vol}(B_p^d)}{\mbox{vol}(B_q^d)}\Big]^{1/(d-\bar p)} \asymp d^{1/q-1/p} $$ into account to get the additional factor in $d$. \noindent\emph{(ii).} The proof by K\"uhn \cite{Ku01} immediately gives the lower bound. The upper bound follows trivially from $\ensuremath{\mathds S^{d-1}}_p \subset \bar B_p^d$. \end{proof} \begin{remark} Note, that in case $p \geq 1$ we have the sharp bounds $$ e_k(\ensuremath{\mathds S^{d-1}}_p,\ell_q^d) \asymp \left\{ \begin{array}{rcl}1&:&1\leq k \leq \log(d),\\ \Big(\frac{\log(1+d/k)}{k}\Big)^{1/p-1/q}&:&\log(d)\leq k\leq d,\\ 2^{-\frac{k}{d-1}}d^{1/q-1/p}&:& k \geq d\,. \end{array} \right. $$ In case $p<1$ there remains a gap between the upper and lower estimate for $e_k(\ensuremath{\mathds S^{d-1}}_p,\ell_q^d)$ if $k\geq d$. However, this gap can be closed by using a different proof technique, see \cite{HiMa13}. \end{remark} \section{Entropy numbers of ridge functions} \label{sec:entropy} This section is devoted to the study of entropy numbers of the classes $\Rs{\alpha,p}$ and $\Rs{\alpha,p,\kappa}$. Especially, we want to relate their behavior to that of entropy numbers of uni- and multivariate Lipschitz functions. This will give us an understanding how ``large'' the ridge function classes are. Let us stress that we are interested in the dependence of the entropy numbers on the underlying dimension $d$, as it is usually done in the area of information-based complexity. To begin with, let us examine uni- and multivariate Lipschitz functions from $\ensuremath{\mathrm{Lip}}_{\alpha}[-1,1]$ and $\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$. Recall the notation $B_{\alpha}:=B_{\ensuremath{\mathrm{Lip}}_{\alpha}[-1,1]}$ and $B_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)}$ for the respective unit balls. The behavior of entropy numbers of univariate Lipschitz functions is well-known, see for instance \cite[Chap. 15, \S 2, Thm. 2.6]{lomavo96}. \begin{lemma}\label{lem:en_univariate_Lip} For $\alpha>0$ there exist two constants $0<c_{\alpha}<C_{\alpha}$ such that \[ c_{\alpha} k^{-\alpha} \leq e_k(\bar B_{\alpha},L_{\infty}([-1,1])) \leq C_{\alpha}k^{-\alpha}\,,\quad k\in \N\,. \] \end{lemma} \noindent This behavior does not change if we consider only functions with first derivative in the origin bounded away from zero, as we do with the profiles in the class $\Rs{\alpha,p,\kappa}$. \begin{proposition}\label{res:en_univariate_Lip_derivative} Let $\alpha > 1$ and $0 < \kappa \le 1$. Consider the class \[ \ensuremath{\mathrm{Lip}}_{\alpha}^{\kappa}(\intcc{-1,1}) = \{ f \in \ensuremath{\mathrm{Lip}}_{\alpha}(\intcc{-1,1}): \|f\|_{\ensuremath{\mathrm{Lip}}_{\alpha}[-1,1]}\leq 1~,~ \abs{f'(0)} \geq \kappa\}. \] For the entropy numbers of this class we have two constants $0<c_{\alpha}<C_{\alpha}$, such that \[ c_{\alpha} k^{-\alpha} \leq e_k(\ensuremath{\mathrm{Lip}}_{\alpha}^{\kappa}([-1,1]), L_{\infty}([-1,1])) \leq C_{\alpha} k^{-\alpha}\,,\quad k\in \N\,. \] \end{proposition} \begin{proof} The upper bound is immediate by Lemma \ref{lem:en_univariate_Lip}. The lower bound is proven in the same way as for general univariate Lipschitz functions of order $\alpha$ except that we have to adapt the ``bad'' functions such that they meet the constraint on the first derivative in the origin. Put again $s=\llfloor\alpha\rrfloor$ and $\beta = \alpha -s>0 $. Consider the standard smooth bump function \begin{equation}\label{eq:bump'} \varphi(x) = \begin{cases} e^{-\frac{1}{1-x^2}}\quad &: \ |x|< 1,\\ 0\quad &:\ |x|\ge 1. \end{cases} \end{equation} Let \[ \psi_{k,b}(x) = \frac{c_{\alpha} \cdot \varphi(5k(x-b))}{k^{\alpha}},\quad k\in \N,\ b\in \field R, \] where $c_{\alpha} = 1/(5^\alpha\norm{\varphi}_{\ensuremath{\mathrm{Lip}}_\alpha})$. The scaling factor $c_{\alpha} k^{-\alpha}$ assures $\psi_{k,b} \in \ensuremath{\mathrm{Lip}}_{\alpha}(\intcc{-1,1})$. Let $a=\pi/4-1/5$ and $I=[a, a+2/5]\subset(0,1)$. We put $h(x)=\sin(x)$ and \begin{align}\label{eq:maxcosin} \gamma=\sup_{j\in{\N}_0}\max_{x\in I}|h^{(j)}(x)|=\max_{x\in I}\max \{\cos(x),\sin(x) \}<1. \end{align} For any multi-index $\theta = (\theta_1, \dots, \theta_k)\in \{0,1\}^k$ let \[ g_{\theta} = (1-\gamma)\sum_{j=1} ^k \theta_j \psi_{k,b_j}, \qquad b_j = a + \frac{2j-1}{5k}. \] Observe, that ${\rm supp \, } g_{\theta}\subset I.$ There are $2^k$ such multi-indices and for two different multi-indices $\hat \theta$ and $\tilde \theta$ we have \[ \norm{g_{\hat \theta} - g_{\tilde \theta}}_{\infty} = (1-\gamma) \norm{\psi_{k,0}}_{\infty} = c_{\alpha} (1-\gamma)e^{-1}k^{-\alpha}. \] Put $f_{\theta} = h + g_{\theta}$. Because of the scaling factors, it is assured that $f_{\theta} \in \ensuremath{\mathrm{Lip}}_{\alpha}^{\kappa}(\intcc{-1,1})$. On the other hand, $f'_\theta(0)=\cos(0)=1.$ Obviously, $\norm{f_{\tilde \theta} - f_{\hat \theta}}_{\infty} = \norm{g_{\tilde \theta} - g_{\hat \theta}}_{\infty}$. We conclude \[ e_k(\ensuremath{\mathrm{Lip}}_{\alpha}^{\kappa}(\intcc{-1,1}), L_{\infty}) \geq c'_{\alpha} k^{-\alpha} \] for $c'_\alpha=(1-\gamma)e^{-1}c_\alpha.$ \end{proof} Considering multivariate Lipschitz functions, decay rates of entropy numbers change dramatically compared to those of univariate Lipschitz functions; they depend exponentially on $1/d$. This is known if the domain is a cube $\Omega=I^d$, see \cite[Chap. 15, \S 2]{lomavo96}. We provide an extension to our situation where the domain is $\Omega = \bar B_2^d$. \begin{proposition} Let $\alpha > 0$. For natural numbers $n$ and $k$ such that $2^{k-1} < n \leq 2^k$ we have \[ e_n(id: \ensuremath{\mathrm{Lip}}_{\alpha}(\bar B_2^d) \to L_{\infty}(\bar B_2^d)) \geq c_{\alpha} e_{k+1}(id: \ell_2^d \to \ell_2^d)^{\alpha}. \] In particular, \old{for $n \geq 2^d$,} we have $e_n(id: \ensuremath{\mathrm{Lip}}_{\alpha}(\bar B_2^d) \to L_{\infty}(\bar B_2^d)) \gtrsim n^{-\alpha/d}$. \end{proposition} \begin{proof} Consider the radial bump function $\varphi(x)$ given by \begin{equation}\label{eq:bump} \varphi(x) = \begin{cases} e^{-\frac{1}{1-\|x\|_2^2}}\quad &: \ \|x\|_2< 1,\\ 0\quad &:\ \|x\|_2\ge 1. \end{cases} \end{equation} Let $s=\llfloor\alpha\rrfloor$. With $c_{\alpha} := (\norm{\varphi}_{\ensuremath{\mathrm{Lip}}_\alpha})^{-1}$ the rescaling \[ \varphi_{\varepsilon}^{\alpha}(x) := c_{\alpha} \varepsilon^{\alpha} \varphi(x/\varepsilon) \] is contained in the closed unit ball of $\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$. For $\varepsilon > 0$ let $\{x_1, \dots, x_n\}$ be a maximal set of $2\varepsilon$-separated points in the Euclidean ball $\bar B_2^d$, the distance measured in $\ell_2^d$. For every multi-index $\theta \in \{0,1\}^n$, we define \[ f_{\theta}(x) := \sum_{j=1}^n \theta_j \varphi^{\alpha}_{\varepsilon}(x-x_j). \] By construction of $\varphi_{\varepsilon}^{\alpha}$, it is assured that $f_{\theta} \in \ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$ and $\norm{f_{\theta}}_{\ensuremath{\mathrm{Lip}}_{\alpha}} \leq 1$. Moreover, we see immediately that $\norm{f_{\theta}}_{\infty}=c_{\alpha} e^{-1}\varepsilon^{\alpha}$, and \[ \norm{f_{\theta} - f_{\theta'}}_{\infty} \ge c_{\alpha} e^{-1}\varepsilon^{\alpha} =: \varepsilon_1 \] for $\theta \neq \theta'$. Therefore, the set $\{f_{\theta}: \theta \in \{0,1\}^n\}$ consists of $2^n$ functions with mutual distances greater than or equal to $\varepsilon_1$. Now choose $\varepsilon$ such that \[ e_{k+1}(\bar B_2^d, \ell_2^d) < 2\varepsilon < e_k(\bar B_2^d, \ell_2^d). \] Then, for $n$ as above, we have $2^k \geq n > 2^{k-1}$, and \[ 2^{n-1} < N_{\varepsilon_1/2}(\bar B_{\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)},L_{\infty}). \] We conclude \[ e_n(id: \ensuremath{\mathrm{Lip}}_{\alpha}(\Omega) \to L_{\infty}(\Omega)) > \varepsilon_1/2 > c'_{\alpha} e_{k+1}(id: \ell_2^d \to \ell_2^d)^{\alpha} \] for $c_\alpha'=c_\alpha/(4e).$ Now it follows immediately from the estimate above and Lemma \ref{lem:schuett} that \[ e_n(id: \ensuremath{\mathrm{Lip}}_{\alpha}(\Omega) \to L_{\infty}(\Omega)) \gtrsim 2^{-\alpha(k+1)/d} \gtrsim n^{-\alpha/d}. \] \end{proof} Now consider ridge functions with Lipschitz profile as given by the class $\Rs{\alpha,p}$. \begin{theorem}\label{res:ridge_entropy} Let $d$ be a natural number, $\alpha > 0$, and $0 < p \leq 2$. Then, for any $k\in \N$, \[ \frac 12 \max \{ e_{2k}(\bar B_p^d,\ell_2^d), e_{2k}(\bar B_{\alpha},L_\infty) \} \leq e_{2k}(\Rs{\alpha,p},L_\infty) \leq e_{k}(\bar B_p^d,\ell_2^d)^{\min\{\alpha,1\}} + e_{k}(\bar B_{\alpha},L_\infty) \;. \] \end{theorem} \begin{proof} \emph{Lower bounds:} For $\varepsilon > 0$ let $g_1,\dots,g_n$ be a maximal set of univariate Lipschitz functions in $\bar B_{\alpha}$ with mutual distances $\norm{g_i - g_j}_{\infty} > \varepsilon$ for $i\not=j$. Now, let $a = (1,0,\dots,0)$ and put $f_i(x) = g_i(a\cdot x)$ for $i=1,\dots,n$. Then, of course, we have $f_i \in \Rs{\alpha,p}$, and $$ \norm{f_i - f_j}_{\infty} = \|g_i-g_j\|_{\infty} > \varepsilon\,. $$ Consequently, the functions $f_1, \dots, f_n$ are $\varepsilon$-separated, as well. This implies $$ e_{2k}(\Rs{\alpha,p},L_\infty) \ge\, \frac12\, e_{2k}(\bar B_{\alpha},L_{\infty})\,. $$ \begin{detail} For arbitrary $\delta > 0$, choose $\delta > 0$ such that \[ e_{2k}(\bar B_{\alpha},L_{\infty}) -\delta < 2 \varepsilon < e_{2k}(\bar B_{\alpha},L_{\infty}). \] Then, \[ N_{\varepsilon}(\Rs{\alpha,p}) \geq M_{2\varepsilon}(\Rs{\alpha,p}) \geq M_{2\varepsilon}(\bar B_{\alpha}) \geq N_{2\varepsilon}(\bar B_{\alpha}) > 2^{2k-1}. \] By definition of entropy numbers, this implies $\varepsilon < e_{2k}(\Rs{\alpha,p})$. Letting $\delta \to 0$, this further implies \[ \frac 12 e_{2k}(\bar B_{\alpha}) \leq e_{2k}(\Rs{\alpha,p}). \] \end{detail} On the other hand, for $\varepsilon > 0$, let $a_1, \dots, a_n$ be a maximal set of vectors in $\bar B_p^d$ with pairwise distances $\norm{a_i-a_j}_2 > \varepsilon$. Furthermore, let $g(t)=t$ and put $ \tilde f_i(x) = g(a_i\cdot x)$ for $i=1,\dots,n$. Then $\tilde f_i \in \Rs{\alpha,p}$ and \begin{align*} \|\tilde f_i- \tilde f_j\|_\infty&=\sup_{x\in \bar B_2^d}|\tilde f_i(x)-\tilde f_j(x)|=\sup_{x\in \bar B_2^d}|g(a_i\cdot x)-g(a_j\cdot x)|\\ &=\sup_{x\in \bar B_2^d}|(a_i-a_j)\cdot x|=\|a_i-a_j\|_2 > \varepsilon. \end{align*} Thus, the functions $\tilde f_1, \dots, \tilde f_n$ are $\varepsilon$-separated w.r.t. the $L_{\infty}$-norm. This implies $$ e_{2k}(\Rs{\alpha,p},L_\infty) \ge\,\frac{1}{2}\, e_{2k}(\bar B_p^d,\ell_2^d)\,. $$ \emph{Upper bound:} Let $1/2>\varepsilon_1, \varepsilon_2>0$ be fixed and put $\varepsilon := \varepsilon_1^{\bar \alpha} + \varepsilon_2$. Let $\mathcal N=\{g_1,\dots,g_n\}$ be a minimal $\varepsilon_1$-net of $\bar B_{\alpha}$ in the $L_\infty$-norm. Further, let $\mathcal M=\{a_1,\dots,a_m\}$ be a minimal $\varepsilon_2$-net of $\bar B_p^d$ in the $\ell_2^d$-norm. Now, fix some ridge function $f:x \mapsto g(a\cdot x)$ in $\Rs{\alpha,p}$, i.e. $\|g\|_{\ensuremath{\mathrm{Lip}}_\alpha}\le 1$ and $\|a\|_p\le 1$. Then there is a function $g_i\in{\mathcal N}$ with $\|g-g_i\|_\infty\le \varepsilon_1$ and a vector $a_j\in{\mathcal M}$ with $\|a-a_j\|_2\le\varepsilon_2.$ Putting this together and writing $\bar\alpha=\min\{\alpha,1\}$, we obtain \begin{align*} \|g(a\cdot x)-g_i(a_j\cdot x)\|_\infty&\le \sup_{x\in \bar B_2^d}|g(a\cdot x)-g(a_j\cdot x)|+|g(a_j\cdot x)-g_i(a_j\cdot x)|\\ &\le \sup_{x\in \bar B_2^d} |g|_{\bar\alpha}\cdot|a\cdot x-a_j\cdot x|^{\bar\alpha}+\|g-g_i\|_\infty\\ &\le \|a-a_j\|^{\bar\alpha}_2 +\|g-g_i\|_\infty\le \varepsilon_1^{\bar\alpha} + \varepsilon_2 = \varepsilon. \end{align*} Hence, the set $ \{x\to g(a\cdot x):g\in {\mathcal N}, a\in{\mathcal M}\} $ is an $\varepsilon$-net of $\Rs{\alpha,p}$ in $L_\infty(\Omega)$ with cardinality \[ \#{\mathcal N}\cdot\#{\mathcal M} = N_{\varepsilon_1}(\bar B_{\alpha},L_{\infty}) \cdot N_{\varepsilon_2}(\bar B_p^d, \ell_2^d) \;. \] Consequently, $N_{\varepsilon}(\Rs{\alpha,p},L_{\infty}) \leq \#{\mathcal N}\cdot\#{\mathcal M}$ and we conclude that \[ e_{2k}(\Rs{\alpha,p},L_{\infty}) \leq e_{k}(\bar B_p^d, \ell_2^d)^{\bar\alpha} + e_{k}(\bar B_{\alpha}, L_{\infty}) \;. \] \begin{detail} For arbitrary $\delta > 0$, choose $\varepsilon_1, \varepsilon_2$ such that \begin{align*} e_{k}(id: \ell_p^d \to \ell_2^d) &< \varepsilon_1 < e_{k}(id: \ell_p^d \to \ell_2^d) + \delta,\\ e_{k}(\bar B_{\alpha},L_{\infty}) &< \varepsilon_2 < e_{k}(\bar B_{\alpha},L_{\infty}) + \delta. \end{align*} Because of $N_{\varepsilon_1+ \varepsilon_2}(\Rs{\alpha,p}, L_{\infty} \leq 2^{2k-2} < 2^{2k-1}$, we obtain in the limit $\delta \to 0$ \[ e_{2k}(\Rs{\alpha,p}) \leq \varepsilon_1 + \varepsilon_2 \leq e_{k}(id: \ell_p^d \to \ell_2^d) + e_{k}(\bar B_{\alpha}, L_{\infty}). \] \end{detail} \end{proof} \begin{remark} In view of Proposition \ref{res:en_univariate_Lip_derivative}, it is easy to see that Theorem \ref{res:ridge_entropy} keeps valid if we replace the class $\Rs{\alpha,p}$ by $\Rs{\alpha,p,\kappa}$. \end{remark} We exemplify the consequences of Theorem \ref{res:ridge_entropy} by considering the case $p=2$; for $0 < p < 2$ estimates would be similar. As the corollary below shows, entropy numbers of ridge functions asymptotically decay as fast as those of their profiles. In contrast to multivariate Lipschitz functions on $\Omega$, the dimension $d$ does not appear in the decay rate's exponent. It only affects how long we have to wait until the asymptotic decay becomes visible. \begin{corollary}\label{thm33} Let $d$ be a natural number and $\alpha > 0$. For the entropy numbers of $\Rs{\alpha,2}$ in $L_{\infty}(\Omega)$ we have \begin{equation}\label{eq:entr} \max(k^{-\alpha},2^{-k/d})\lesssim e_k(\Rs{\alpha,2},L_\infty)\lesssim \begin{cases} 1 & : k \le c_{\alpha} d\log d,\\ k^{-\alpha} & : k \ge c_{\alpha} d\log d\,, \end{cases} \end{equation} for some universal constant $c_{\alpha}>0$ which does not depend on $d$. \end{corollary} Before we turn to the proof, let us note that \eqref{eq:entr} implies that $$ e_k(\Rs{\alpha,2},L_\infty) \asymp 1\quad \text{if}\quad k\le d, $$ and $$ e_k(\Rs{\alpha,2},L_\infty) \asymp k^{-\alpha} \quad \text{if}\quad k\ge c_{\alpha} d\ln d. $$ Hence, entropy numbers of ridge functions are guaranteed to decay like those of their profiles for $k\ge c_{\alpha}d\log d$---and surely behave differently for $k\le d$. \begin{proof}[Proof of Corollary \ref{thm33}] The lower bound in \eqref{eq:entr} follows from Theorem \ref{res:ridge_entropy} combined with Lemma \ref{lem:schuett}, and Lemma \ref{lem:en_univariate_Lip}. The upper bounds are proven in the same manner, using the simple fact that for every $\alpha>0$ there are two constants $c_\alpha,c'_\alpha>0$, such that $k\ge c_\alpha d\log d$ implies that $2^{-\min\{\alpha,1\}k/d}\le c_\alpha'k^{-\alpha}.$ \begin{detail} To see the upper bound, let $\mathcal M$, $\mathcal N$ as in the proof of Theorem \ref{res:ridge_entropy}. It follows from \cite[Chapter 15]{lomavo96} that one may assume that $\#{\mathcal M}\le (3/\varepsilon)^d$. Furthermore, by Lemma \ref{lem:en_univariate_Lip}, we observe that there is a constant $c>0$, such that \[ \#{\mathcal N}\le 2^{c\varepsilon^{-1/\alpha}}. \] In detail, the argument is as follows. There is a constant $c' > 0$ such that $e_k < c'k^{-\alpha}$. By definition of entropy numbers, for all $\delta > e_k$ we have $N_{\delta}(B_{\text{Lip}\del{\intcc{-1,1}}}) \leq 2^k$. Now put $\varepsilon = c'k^{-\alpha}$, $c = c'^{1/\alpha}$ and choose $\mathcal N$ an optimal $\varepsilon$-net. Hence, for every $1/2>\varepsilon>0$, there is a $2\varepsilon$-net of $\Rs{\alpha,2}$ in $L_\infty(\Omega)$ with cardinality at most $(3/\varepsilon)^d\cdot 2^{c\varepsilon^{-1/\alpha}}.$ Let us convert this into an upper bound for the entropy numbers of $\Rs{\alpha,2}$. Let $\varepsilon_0 >0$ such that $(6/\varepsilon_0)^{1/\alpha} = d\log d$. Taking the logarithm gives the estimate $\log d \geq (2\alpha)^{-1}\log(6/\varepsilon_0)$. If we plug this back into the equation and exponentiate we obtain \[ (6/\varepsilon_0)^d \leq 2^{2\alpha (6/\varepsilon_0)^{1/\alpha}} \] and, by monotonicity, this estimate remains valid for all $0<\varepsilon \leq \varepsilon_0$. Consequently, we find $C$ such that $N_{\varepsilon}(\Rs{\alpha,2},L_{\infty}) \leq 2^{k-1}$ for $\varepsilon = Ck^{-\alpha}$, provided $k \geq c_{\alpha} d\log d$ with the constant $c_{\alpha} = (C/6)^{1/\alpha}$ independent of $k$ and $d$. We conclude $e_k(\Rs{\alpha,2},L_{\infty}) \lesssim k^{-\alpha}$. Observe that for $\varepsilon_0^{1/\alpha} = 6^{1/\alpha}(d\log d)^{-1}$ we have \[ \log\del{(6/\varepsilon_0)^{1/\alpha}} = \log d + \log\log d \leq 2 \log d \] which implies $\log d \geq \frac {1}{2\alpha} \log(6/\varepsilon_0)$. Therefore, $(6/\varepsilon_0)^{1/\alpha} \geq (2\alpha)^{-1} d \log(6/\varepsilon_0)$. The estimate holds also true for all $\varepsilon \leq \varepsilon_0$. To see this put $h(x) = \alpha x^{1/\alpha} (\log x)^{-1}$. The function $h(6/\varepsilon)$ is decreasing for all $0 < \varepsilon < 3\cdot 2^{1-\alpha}$, as \[ \od{}{\varepsilon}\del{h(6/\varepsilon)} = -6\frac{ (6/\varepsilon)^{1/\alpha -1} \del{\log(6/\varepsilon)-\alpha}}{\del{\varepsilon \log(6/\varepsilon)}^2} \] is negative for $\log \frac{6}{\varepsilon} > \alpha$, i.e. $\varepsilon < 3\cdot 2^{1-\alpha}$. Hence $\alpha (6 /\varepsilon)^{1/\alpha} \del{\log 6 / \varepsilon}^{-1} \geq d/2$ for $\varepsilon < \varepsilon_0$, provided $d > 1$. This yields \[ \log N_{\varepsilon}\del{\Rs{\alpha,2},L_{\infty}} \leq d \log(6/\varepsilon) + \frac{2c}{\varepsilon} \leq \frac{2 \cdot 6^{1/\alpha} + 2c}{\varepsilon^{1/\alpha}}. \] Consequently, we find $C > 0$ such that $N_{\varepsilon}\del{\Rs{\alpha,2},L_{\infty}} \leq 2^{k-1}$ holds true for $\varepsilon = C k^{-\alpha}$, provided $k \geq \del{\frac C6}^{1/\alpha} d\log(d)$. We conclude \[ e_k(\Rs{\alpha,2},L_{\infty}) \lesssim k^{-\alpha}, \] provided $k \geq c_{\alpha} \del{d\log d}$. Note that the constant $c_{\alpha} =(C/6)^{1/\alpha}$ is independent of $d$. \end{detail} \end{proof} Summarizing this section, the classes of ridge functions with Lipschitz profiles of order $\alpha$ are essentially as compact as the class of univariate Lipschitz functions of order $\alpha$. Consequently, when speaking in terms of metric entropy, these classes of functions must be much smaller than the class of multivariate Lipschitz functions of order $\alpha$. \section{Sampling numbers of ridge functions} \label{sec:sampling} In light of Section \ref{sec:entropy}, one is led to think that efficient sampling of ridge functions should be feasible. Moreover, their simple, two-component structure naturally suggests a two-step procedure: first, use a portion of the available function samples to identify either the profile or the direction; then, use the remaining samples to unveil the other component. However, in Subsection \ref{sec:sampling1}, we learn that for ridge functions in the class $\Rs{\alpha,p}$, sampling is almost as hard as sampling of general multivariate Lipschitz functions on the Euclidean unit ball. In particular, such two-step procedures as sketched above cannot work in an efficient manner. It needs additional assumptions on the ridge profiles or directions. We discuss this in Subsection \ref{sec:sampling2}. \subsection{Sampling of functions in $\Rs{\alpha,p}$} \label{sec:sampling1} As usual, throughout the section let $\alpha > 0$ be the Lipschitz smoothness of profiles, $s = \llfloor \alpha \rrfloor$ the order up to which derivatives exist, and let $0 < p \leq 2$ indicate the $p$-norm such that ridge directions are contained in the closed $\ell_p^d$-ball. The algorithms we use to derive upper bounds are essentially the same as those which are known to be optimal for general multivariate Lipschitz functions. Albeit, the ridge structure allows a slightly improved analysis at least in case $p < 2$. \begin{proposition}\label{res:sampling_upper_bound} Let $\alpha>0$ and $0<p\leq 2$. For $n \geq {d+s \choose s}$ sampling points the $n$-th sampling number is bounded from above by \begin{equation}\label{eq:sampl} g_{n,d}^{\text{lin}}(\Rs{\alpha, p},L_{\infty}) \leq e_{k-\Delta}(\bar B_2^d, \ell_{p'}^d)^{\alpha} \,, \end{equation} where $k = \lfloor \log n \rfloor +2$, $\Delta = 1+\lceil \log {d+s \choose s} \rceil$, and $p'$ is the dual index of $p$. \end{proposition} \begin{proof} \emph{Case $\alpha \leq 1$:} In this case, $s=0$ and $\Delta = 1.$ We choose sampling points $x_1, \dots, x_{2^{k-2}}$ such that they form an $\varepsilon$-covering of $\bar B_2^d$ in $\ell_{p'}^d$. Given this covering, we construct (measurable) sets $U_1, \dots, U_{2^{k-2}}$ such that $U_i \subseteq x_i+\varepsilon \bar B_{p'}^d$ for $i=1,\dots,2^{k-2}$ and \[ \bigcup_{i=1}^{2^{k-2}} \Bigl(x_i+\varepsilon \bar B_{p'}^d\Bigr) = \bigcup_{i=1}^{2^{k-2}} U_i, \qquad U_i \cap U_j = \emptyset \; \text{for } i \neq j. \] Now we use piecewise constant interpolation: we approximate $f = g(a\cdot ) \in \Rs{\alpha,p}$ by $S f := \sum_{i=1}^{2^{k-2}} f(x_i) \mathds{1}_{U_i}$. Then, \begin{align}\label{f7} \norm{f - Sf}_{\infty} &= \sup_{i=1,\dots,2^{k-1}} \sup_{x \in U_i} \abs{f(x) - f(x_i)}\\ &\leq \sup_{i=1,\dots,2^{k-2}} \sup_{x \in U_i} \norm{g}_{\ensuremath{\mathrm{Lip}}_\alpha} \norm{a}^\alpha_p \norm{x-x_i}_{p'}^{\alpha} \leq \varepsilon^{\alpha}. \end{align} The smallest $\varepsilon$ is determined by the $(k-1)$-th entropy number $e_{k-1}(\bar B_2^d, \ell_{p'}^d)$. Consequently, \begin{equation}\label{f8} g^{\text{lin}}_{n,d}(\Rs{\alpha,p},L_{\infty}) \leq g^{ \text{lin}}_{2^{k-2},d}(\Rs{\alpha,p},L_{\infty})\leq e_{k-1}(\bar B_2^d,\ell_{p'}^d)^{\alpha}. \end{equation} \emph{Case $\alpha > 1$:} We choose the sampling points $x_1, \dots, x_{2^{k-\Delta-1}}$ and the sets $U_1,\dots,U_{2^{k-\Delta-1}}$ as above. However, instead of piecewise constant interpolation we apply on each of the sets $U_i \subseteq x_i+\varepsilon \bar B_{p'}^d$ a Taylor formula of order $s$ around the center $x_i$. That is, to approximate a given $f=g(a\cdot ) \in \Rs{\alpha,p}$ we set $Sf := \sum_{i=1}^{2^{k-\Delta-1}} T_{x_i,s} f \mathds{1}_{U_i}$. Then, by Lemma \ref{lem:taylorapprox} (ii), we have \begin{align*} \norm{f - Sf}_{\infty} &= \sup_{i=1,\dots,2^{k-\Delta-1}} \sup_{x \in U_i} \abs{f(x) - T_{x_i,s}f(x)} \leq \frac{1}{s!} \norm{x - x_i}_{p'}^{\alpha} \leq \varepsilon^{\alpha}. \end{align*} It takes $2^{k-\Delta-1} {d+s \choose s} \leq n$ function values to approximate all the $T_{x_i,s}$ above up to arbitrary precision by finite-order differences, cf.\ \cite{V2013}. The smallest $\varepsilon$ is now determined by the $(k-\Delta)$-th entropy number $e_{k-\Delta}(\bar B_2^d,\ell_{p'}^d)$. We conclude \begin{align} g^{\mathrm{lin}}_{n,d}(\Rs{\alpha,p},L_{\infty}) \leq g^{\mathrm{lin}}_{2^{k-\Delta-1},d}(\Rs{\alpha,p},L_{\infty}) \leq e_{k-\Delta}(\bar B_2^d, \ell_{p'}^d)^{\alpha}. \end{align} \end{proof} We turn to an analysis of lower bounds for the classes $\Rs{\alpha,p}$. Our strategy is to find ``bad'' directions which map, for a given budget $n \in \N$, all possible choices of $n$ sampling points to a small range of $[-1,1]$. There, we let the ``fooling'' profiles be zero; outside of that range, we let the profiles climb as steep as possible. Proposition \ref{res:sampling_lower_bound} below states the lower bound that results from this strategy, provided that the ``bad'' directions are given by some $\mathcal M \subseteq \bar B_p^d \setminus \{0\}$. We discuss appropriate choices of $\mathcal M$ later. In the sequel, we use the mapping $\Psi:\field R^d\setminus\{0\}\to \ensuremath{\mathds S^{d-1}}_2$ defined by $x\mapsto x/\|x\|_2$\,. \begin{proposition}\label{res:sampling_lower_bound} Let $\alpha>0$, $0<p\leq 2$, and $\mathcal M \subseteq \bar B_p^d \setminus \{0\}$. Then, for all natural numbers $k$ and $n$ with $n\leq 2^{k-1}$, we have \[ g^{\text{ada}}_{n,d}(\Rs{\alpha,p},L_{\infty}) \geq c_{\alpha} \inf_{a \in \mathcal M} \|a\|_2^{\alpha} \cdot e_k(\Psi(\mathcal M),\ell_2^d)^{2\alpha}\,. \] The constant $c_{\alpha}$ depends only on $\alpha$. \end{proposition} \begin{proof} Let us first describe the ``fooling'' profiles in detail. For each $a \in \mathcal M$ and $\varepsilon<1$, we define a function \begin{align}\label{eq:bad_function} g_{a,\varepsilon}(t)=\vartheta_\alpha \big[(t-\|a\|_2(1-\varepsilon^2/2))_+\big]^{\alpha} \end{align} on the interval $[-1,1]$. The factor $\vartheta_\alpha$ assures that $\|g_{a,\varepsilon}\|_{\ensuremath{\mathrm{Lip}}_\alpha[-1,1]} = 1$. Put $f_{a,\varepsilon}(x) = g_{a,\varepsilon}(a\cdot x)$. By construction, we have that $f_{a,\varepsilon} \in \Rs{\alpha,p}$. Moreover, whenever $x \in \bar B_2^d$ and $a\in \mathcal{M}$ is such that \begin{align}\label{f31} \varepsilon^2 < \norm{x-\Psi(a)}^2_2 \end{align} then $\varepsilon^2 \leq 2-2(x \cdot \Psi(a))$ and hence \begin{equation}\label{f3} x \cdot a = \|a\|_2 (x \cdot \Psi(a)) < \|a\|_2(1-\varepsilon^2/2)\,. \end{equation} Therefore, \eqref{f31} implies $f_{a,\varepsilon}(x) = 0$. Now, let $n \leq 2^{k-1}$ and $S \in \ensuremath{\mathcal S^{\text{ada}}}_n$ be an adaptive algorithm with a budget of $n$ sampling points. Clearly, the first sampling point $x_1$ must have been fixed by $S$ in advance. Then, let $x_2, \dots, x_n$ be the sampling points which $S$ would choose when applied to the zero function. Furthermore, let $F(x_1,\dots,x_n) \subseteq \Rs{\alpha,p}$ denote the set of functions that make $S$ choose the very points $x_1,\dots, x_n$. Obviously, we have $f_{a,\varepsilon} \in F(x_1,\dots,x_n)$ if \eqref{f31} holds for every $x_i$, $i=1,...,n$. This is true for some $a\in \mathcal{M}$ if we choose $\varepsilon<e_k(\Psi({\mathcal M}),\ell_2^d)$. For the respective function $f_{a,\varepsilon}$, we have in particular $N^{\text{ada}}_S(f_{a,\varepsilon}) = 0$ and hence $S[f_{a,\varepsilon}] = S[-f_{a,\varepsilon}]$. Consequently, \begin{equation}\label{f5} \max \big\{ \|f_{a,\varepsilon} - S[f_{a,\varepsilon}]\|_{\infty}, \|-f_{a,\varepsilon} - S[-f_{a,\varepsilon}]\|_{\infty} \big\} \geq \|f_{a,\varepsilon}\|_{\infty} = g_{a,\varepsilon}(\|a\|_2) = c_{\alpha} \|a\|_2^{\alpha} \varepsilon^{2\alpha}\,, \end{equation} where $c_{\alpha} := 2^{-\alpha} \vartheta_{\alpha}$. Since $\varepsilon$ has been chosen arbitrarily but less than $e_{k}(\Psi(\mathcal M),\ell_2^d)$, we are allowed to replace $\varepsilon$ by $e_{k}(\Psi(\mathcal M),\ell_2^d)$ in \eqref{f5} and get \[ \sup_{f \in \Rs{\alpha,p}} \|f - S(f)\|_{\infty} \geq c_{\alpha} \inf_{a \in \mathcal M} \|a\|_2^{\alpha}\cdot e_{k}(\Psi(\mathcal M),\ell_2^d)^{2\alpha}. \] Taking the infimum over all algorithms $S \in \ensuremath{\mathcal S^{\text{ada}}}_n$ yields \[ g^{\text{ada}}_{n,d}(\Rs{\alpha, p},L_{\infty}) \geq c_{\alpha} \, \inf_{a \in \mathcal M} \|a\|_2^{\alpha} \, e_{k}(\Psi(\mathcal M), \ell_2^d)^{2\alpha}. \] \end{proof} \begin{theorem}\label{res:ridge_sampling} Let $\alpha>0$, $s = \llfloor \alpha \rrfloor$, and $0<p\leq 2$. For the classes $\Rs{\alpha,p}$, we have the following bounds: \\ {\em (i)} The $n$-th (linear) sampling number is bounded from above by \begin{align*} g^{\mathrm{lin}}_{n,d}(\Rs{\alpha,p},L_{\infty}) \leq C_{p,\alpha} \begin{cases} 1 &: n \leq 2d{d+s \choose s},\\ \left[ \frac{ \log(1+d/\log n_1)}{\log n_1} \right]^{\alpha(1/\max\{1,p\} - 1/2)} &: 2d{d+s \choose s} < n \leq 2^{d+1}{d+s \choose s},\\ n^{-\alpha/d} \; d^{-\alpha(1/\max\{p,1\}-1/2)} &: n > 2^{d+1}{d+s \choose s}\,, \end{cases} \end{align*} where $n_1 = n/{[2{d+s \choose s}]}$, and the constant $C_{p,\alpha}$ depends only on $\alpha$ and $p$.\\\\ {\em (ii)} The $n$-th (adaptive) sampling number is bounded from below by \begin{align*} g^{\mathrm{ada}}_{n,d}(\Rs{\alpha,p},L_{\infty}) \geq c_{p,\alpha} \begin{cases} 1 &: n <d,\\ \left[ \frac{\log_2 \del{1+ d/(2+\log_2 n)}}{2+\log_2 n}\right]^{\alpha(1/p - 1/2)} &: d \leq n <2^{d-1},\\ n^{-2\alpha/(d-1)} \; d^{-\alpha(1/p - 1/2)} &: n \geq 2^{d-1}\,. \end{cases} \end{align*} The constant $c_{p,\alpha}$ depends only on $\alpha$ and $p$. \end{theorem} \begin{proof} \emph{(i)} The upper bound is a direct consequence of Proposition \ref{res:sampling_upper_bound} and Lemma \ref{lem:schuett}. Note that, for $k$ and $\Delta$ as in Proposition \ref{res:sampling_upper_bound}, it holds true that $k - \Delta - 2 \leq \log n_1 \leq k - \Delta$. Note also that \[ {d+s \choose s}^{\alpha/d} \leq (1+s)^{s\alpha/d} d^{s\alpha/d} \leq ((1+s)e)^{s\alpha} \] ensures that the constant $C_{p,\alpha}$ can be chosen independently of $d$ and $n$. \emph{(ii) Case $n < d$}. Let $\mathcal M = \{\pm e_1,\dots,\pm e_d\}$ be the set of positive and negative canonical unit vectors. Clearly, we have $\sharp \mathcal{M} = 2d$ and every two distinct vectors in $\mathcal{M}$ have mutual $\ell_2^d$-distance equal to or greater than $\sqrt{2}$. Let $k$ be the smallest integer such that $n \leq 2^{k-1}$; this implies $2^{k-1}<2d$. Hence, whenever $2^{k-1}$ balls of radius $\varepsilon$ cover the set $\mathcal M$, there is at least one $\varepsilon$-ball which contains two elements from $\mathcal{M}$. In consequence, we have $2\varepsilon\geq\sqrt{2}$ and hence $e_k(\mathcal{M},\ell_2^d) \geq \sqrt{2}/2$. By Proposition \ref{res:sampling_lower_bound} and the fact that $\mathcal{M} = \Psi(\mathcal{M})$, we obtain \[ g^{\text{ada}}_{n,d}(\Rs{\alpha,p},L_{\infty}) \geq c_{\alpha} e_k(\mathcal{M},\ell_2^d)^{2\alpha} \geq c_{\alpha} 2^{-\alpha}\,. \] \noindent{\em Case $d \leq n < 2^{d-1}$}. For $m\leq d$, consider the subset of $m$-sparse vectors of the $p$-sphere, \[ \ensuremath{\mathfrak S^{d-1}}_{m,p} = \big\{ x \in \ensuremath{\mathds S^{d-1}}_p: \;\sharp\;{\rm supp \, }(x) = m \big\}. \] Using the combinatorial construction of \cite{GS}, cf.\ also \cite{FPRU10}, we know that there exist at least $(d/(4m))^{m/2}$ vectors in $\Psi(\ensuremath{\mathfrak S^{d-1}}_{m,p})=\ensuremath{\mathfrak S^{d-1}}_{m,2}$ having mutual $\ell_2^d$-distance greater than $1/\sqrt{2}$. Therefore, we have \begin{equation}\label{f32} \ell \leq m/2 \log(d/(4m))\quad \implies \quad e_{\ell}(\Psi(\ensuremath{\mathfrak S^{d-1}}_{m,p}),\ell_2^d) \geq \sqrt{2}/4. \end{equation} Let $k$ again be the smallest integer such that $n \leq 2^{k-1}$. Hence, $k\leq d$. Choose \[ m^* := \big\lfloor \min \{ 4k/\log(d/(4k)), k \} \big\rfloor\leq k\,. \] Because of $k>\log d$, we have $\min\{\log d,4\} \leq m^{*}\leq d$. Put $\mathcal M = \ensuremath{\mathfrak S^{d-1}}_{m^*,p}$. If $k \leq d/64$, then $\log(d/(4k)) \geq 4$ and $k \leq m^*\log(d/(4k))/2 \leq m^*\log(d/(4m^*))/2$. Hence, by \eqref{f32}, one has $e_k(\Psi(\ensuremath{\mathfrak S^{d-1}}_{m^*,p}),\ell_2^d) \geq \sqrt{2}/4.$ Consequently, by Proposition \ref{res:sampling_lower_bound}, it follows that \begin{align*} g^{\text{ada}}_{n,d}(\Rs{\alpha,d},L_{\infty}) &\geq c_{\alpha} (m^*)^{\alpha(1/2-1/p)} e_k(\Psi(\ensuremath{\mathfrak S^{d-1}}_{m^*,p}),\ell_2^d)^{2\alpha} \\ &\geq c_{\alpha}8^{-\alpha}4^{-\alpha(1/p-1/2)}\Big[\frac{\log(d/(4k))}{k}\Big]^{\alpha(1/p-1/2)}\\ &\geq c_{\alpha} 8^{-\alpha} 8^{-\alpha(1/p-1/2)} \left( \frac{\log(1+d/k)}{k}\right)^{\alpha(1/p-1/2)}\\ &\geq c_{p,\alpha}\left( \frac{\log(1+d/k)}{k}\right)^{\alpha(1/p-1/2)}\,. \end{align*} On the other hand, if $d/64 < k \leq d$, then $m^* = k$. By $\mathds S_2^{k-1} \subset \Psi(\ensuremath{\mathfrak S^{d-1}}_{m^*,p}) \subset \ensuremath{\mathds S^{d-1}}_2$ and Lemma \ref{res:entropy_sphere}, we have $e_k(\Psi(\ensuremath{\mathfrak S^{d-1}}_{m^*,p}), \ell_2^d) \asymp 1$. Proposition \ref{res:sampling_lower_bound}, together with $\log(1+d/k) < 8$ for $k>d/64$, implies \begin{align*} g^{\text{ada}}_{n,d}(\Rs{\alpha,p},L_\infty) \geq c'_{\alpha} k^{-\alpha(1/p-1/2)} &\geq c'_{\alpha}8^{-\alpha(1/p-1/2)} \left( \frac{\log(1+d/k)}{k} \right)^{\alpha(1/p-1/2)}\\ &= c'_{p,\alpha}\left( \frac{\log(1+d/k)}{k} \right)^{\alpha(1/p-1/2)}\,. \end{align*} \noindent{\em Case $n \geq 2^{d-1}$}. Again, $k$ is chosen such that $2^{k-2}<n\leq 2^{k-1}$, which implies $k \geq d$. In this case, we choose $\mathcal M = \ensuremath{\mathds S^{d-1}}_p$. By Lemma \ref{res:entropy_sphere} and Proposition \ref{res:sampling_lower_bound}, we obtain \begin{align*} g^{\text{ada}}_{n,d}(\Rs{\alpha,p},L_{\infty}) &\geq c_{\alpha} \; d^{-\alpha(1/p-1/2)}e_{k}(\ensuremath{\mathds S^{d-1}}_2,\ell_2^d)^{2\alpha}\\ &\geq c_{\alpha}d^{-\alpha(1/p-1/2)} \; (4n)^{-2\alpha/(d-1)}\\ &\geq c_{\alpha}4^{-2\alpha}d^{-\alpha(1/p-1/2)}n^{-2\alpha/(d-1)}\,. \end{align*} This completes the proof. \end{proof} \begin{remark}\label{prop:p=2} Consider the situation $p = 2$. For sampling numbers with $n \leq 2^{d-1}$, we have \[ g^{\text{ada}}_{n,d}(\Rs{\alpha,2},L_{\infty}) \asymp 1. \] For sampling numbers with $n \geq 2^{d+1} {d+s \choose s}$, we have \begin{equation}\label{eq:gap} n^{-2\alpha/(d-1)}\lesssim g^{\text{ada}}_{n,d}(\Rs{\alpha, 2},L_{\infty}) \lesssim n^{-\alpha/d}. \end{equation} The upper estimate on sampling numbers is exactly the same as for multivariate Lipschitz functions from $\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$. Although there is a gap between lower and upper bound in \eqref{eq:gap}, the factor $1/(d-1)$ in the exponent of the lower bound allows us to conclude that sampling of ridge functions in $\Rs{\alpha,2}$ is nearly as hard as sampling of general Lipschitz functions from $\ensuremath{\mathrm{Lip}}_{\alpha}(\Omega)$. Hence, we have the opposite situation to Section \ref{sec:entropy}, where ridge functions in $\Rs{\alpha,2}$ behave similar to univariate Lipschitz functions. \end{remark} \begin{remark}\label{rem:CDDKP} Let us consider the modified ridge function classes $\tilde{\mathcal R}_d^{\alpha,p}$ and $\bar{\mathcal R}_{d}^{\alpha,p}$ defined by \begin{equation}\label{mod_ridge} \tilde{\mathcal R}_d^{\alpha,p} := \big\{ f: [0,1]^d \to \field R\,:\, f(x) = g(a \cdot x), \; \|g\|_{\ensuremath{\mathrm{Lip}}_{\alpha}[0,1]} \leq 1 ,\; \|a\|_p \leq 1, \; a \geq 0 \big\}, \end{equation} for $0<p\leq 1$, and \begin{equation}\label{mod_ridge2} {\bar{\mathcal{R}}}_{d}^{\alpha,p} := \big\{ f: \bar B^d_2 \cap [0,1]^d\, \to \field R\,:\,f(x) = g(a \cdot x), \; \|g\|_{\ensuremath{\mathrm{Lip}}_{\alpha}[0,1]} \leq 1 ,\; \|a\|_p \leq 1, \; a \geq 0 \big\}. \end{equation} for $0<p\leq 2$. Here, $a\ge 0$ means, that all coordinates of $a$ are non-negative.\\ {\em (i)} In the recent paper \cite{CDDKP} it has been shown that there is an adaptive algorithm which attains a decay rate of $n^{-\alpha}$ for the worst-case $L_{\infty}$-approximation error with respect to the class $\tilde{\mathcal R}_d^{\alpha,1}$, provided that $n \geq d$. In terms of adaptive sampling numbers (such that the feasible algorithms are adjusted to the domain $[0,1]^d$), this reads as \begin{align}\label{eq:ada_helps} g^{\text{ada}}_{n,d}(\tilde{\mathcal R}_d^{\alpha,1}, L_{\infty}) \leq C_{\alpha}n^{-\alpha}, \quad n \geq d. \end{align} At the same time, a careful inspection of the proofs of Propositions \ref{res:sampling_upper_bound}, \ref{res:sampling_lower_bound}, and Theorem \ref{res:ridge_sampling} shows that the results can be carried over to the classes $\bar{\mathcal{ R}}_d^{\alpha,p}$ for all $0<p\leq 2$. In particular, for $0<p\leq 1$, we have the lower bound \begin{align}\label{eq:ada_doesnt_help} g^{\text{ada}}_{n,d}(\bar{\mathcal R}_d^{\alpha,p}, L_{\infty}) \geq c_{p,\alpha} n^{-2\alpha/(d-1)}d^{\alpha(1/2-1/p)} , \quad n \in \N\,. \end{align} The estimates \eqref{eq:ada_helps} and \eqref{eq:ada_doesnt_help} look conflicting at first glance. We encounter the rather surprising phenomenon, that enlarging the domain of the class of functions under consideration leads to better approximation rates. To understand this, let us briefly sketch the adaptive algorithm of \cite{CDDKP}. For $f = g(a\cdot) \in \tilde{\mathcal R}_d^{\alpha,p}$ not the zero function, the idea is to first sample along the diagonal of the first orthant, that is, at points $x = t (1,\dots,1)$ with $t \in [0,1]$. Importantly, it is guaranteed that we can take samples from the whole relevant range $[0,\|a\|_1]$ of the profile $g$ of $f$. This in turn assures that, by sampling adaptively along the diagonal, we find a small range in $[0,\|a\|_1]$ where the absolute value of $g'$ is strictly larger than 0. Then, the ridge direction $a$ can be recovered in a similar way as we do in Subsection \ref{sec:sampling2}. On the other hand, for the classes $\bar{\mathcal R}_d^{\alpha,p}$, this adaptive algorithm will not work. Assume we sample again along the (rescaled) diagonal. This time, we can be sure that we are able to reach every point in the intervall $[0,\|a\|_1/\sqrt{d}]$. But this interval is in most cases strictly included in the relevant interval $[0,\|a\|_2]$ for $g$. Hence, it is not guaranteed anymore that we sample the whole relevant range of $g$ and find an interval on which $g'$ is not zero. (ii) Admittedly, the domain $\Omega = [0,1]^d \cap B^d_2$ in \eqref{mod_ridge2} is a somewhat artificial choice in case of $p\leq1$, whereas the cube $\Omega = [0,1]^d$ seems natural. Conversely, the definition in \eqref{mod_ridge} is not reasonable in case $p>1$, since then $a\cdot x$ might exceed the domain interval for $g$. However, $\Omega = [0,1]^d \cap B^d_2$ is the natural choice for $p=2$ in \eqref{mod_ridge2}. In this situation, we suffer from the curse of dimensionality for adaptive algorithms using standard information, see Remark \ref{prop:p=2} and Theorem \ref{res:tractability},(1) below. This shows that the condition $p\leq1$ is essential in the setting of \cite{CDDKP} and that \eqref{eq:ada_helps} can not be true for the class $\bar{\mathcal{R}}^{\alpha,2}_d$. \end{remark} \subsection{Recovery of ridge directions} \label{sec:sampling2} We return to the question under which conditions the two-step procedure sketched at the beginning of Section \ref{sec:sampling} is successful. The adaptive algorithm of \cite{CDDKP}, which we have already discussed in Remark \ref{rem:CDDKP}, first approximates the profile $g$. Unfortunately, we could already argue that this algorithm cannot work in our setting. There is an opposite approach in Fornasier et al. \cite{FSV12}, which first tries to recover the ridge direction and conforms to our setting. Following the ideas of \cite{BP}, the authors developed an efficient scheme using Taylor's formula to approximate ridge functions with $C^s$ profile obeying certain integral condition on the modulus of its derivative. This condition was satisfied for example if $\abs{g'(0)} \ge \kappa > 0$. In their approach, the smoothness parameter $s$ had to be at least $2$. Using a slightly different analysis, this scheme turns out to work for Lipschitz profiles of order $\alpha>1$. Before we turn to the analysis, let us sketch the Taylor-based scheme in more detail. As transposes of matrices and vectors appear frequently, for reasons of convenience, we write $a \cdot x = a^Tx$ for the remainder of this subsection. Now, Taylor's formula in direction $e_i$ yields \begin{align*} f(h e_i) &= f(0) + h \nabla f(\xi^{(i)}_h e_i)^T e_i\\ &= g(0) + h g'(\xi^{(i)}_h a_i) a_i\,. \end{align*} Hence, we can expose the vector $a$, distorted by a diagonal matrix with components \[ \xi_h = (g'(\xi^{(1)}_h a_1), \dots, g'(\xi^{(d)}_h a_d)) \] on the diagonal. In total, we have to spend only $d+1$ function evaluations for that. Moreover, each of $\xi_h$'s components can be pushed arbitrarily close to $g'(0)$. This gives an estimate $\hat{a}$ of $a/\norm{a}_2$, along which we can now conduct classical univariate approximation. Effectively, one samples a distorted version of $g$ given by \[ \tilde g: \intcc{-1,1} \rightarrow \field R, \; t \mapsto f(t\hat a) = g(t a^T \hat a). \] The approximation $\hat g$ obtained in this way, together with $\hat a$, forms the sampling approximation to $f$, \[ \hat f(x) = \hat g(\hat a^Tx). \] Observe that $\tilde g(\hat a^T x) = g(a^T \hat a \hat a ^T x)$, so it is crucial that $\hat a \hat a ^T$ spans a subspace which is close to the one-dimensional subspace spanned by $aa^T$, in the sense that \[ \big\| a^T ( I_d - \hat a \hat a^T) \big\|_2 \] has to be small. Importantly, this gives the freedom to approximate $a$ only up to a sign. Finally, let us note that if the factor $g'(0)$ can become arbitrary small, the information we get through Taylor's scheme about $a$ becomes also arbitrarily bad. Hence, for this approach to work, it is necessary to require $\abs{g'(0)} \ge \kappa$. \begin{lemma}\label{lem:reconstruct_a} Let $0 < \beta \leq 1$, $0 < \kappa \leq 1$, and $\varepsilon > 0$. Further let $\delta = \frac{\varepsilon \cdot \kappa}{2+\varepsilon}$ and $h = (\delta/2)^{1/\beta}$. For any $g \in \ensuremath{\mathrm{Lip}}^{\kappa}_{1+\beta}(\intcc{-1,1})$ and $a \in \bar B_2^d$ with $a\not=0$ let $f = g(a \cdot )$. Put \begin{equation}\label{eq:rec_a} \tilde{a}_i = \frac{f(he_i) - f(0)}{h}, \quad i=1, \dots, d \end{equation} and $\hat a = \tilde a / \|\tilde a\|_2$. Then \[ \big\| { \ensuremath{\mbox{\rm sign} \,} }(g'(0)) \hat a - a/\|a\|_2 \big\|_2 \leq \varepsilon. \] \end{lemma} \begin{proof} By the mean value theorem of calculus there exist $\xi_h^{(i)} \in \intcc{0,h}$ such that \[ \tilde a_i = g'(\xi_h^{(i)}a_i)a_i. \] By H\"older continuity we get \[ |g'(\xi_h^{(i)}a_i)-g'(0)| < 2|g'|_{\beta} |a_i|^{\beta} |h|^{\beta} \leq \delta \] for all $i=1,\dots,d$. Let us observe that $\delta<\kappa$ and, therefore, $\tilde a\not=0$ and $\hat a$ is well defined. Put $\xi = (g'(\xi_h^{(i)}a_i))_{i=1}^d$. Then we can write $\tilde a = \mathrm{diag}(\xi)a$. For the norm of $\tilde a$ we get \begin{align*} \|\tilde a\|_2 &\leq \| \mathrm{diag}(\xi)a - g'(0)a\|_2 + |g'(0)| \|a\|_2\\ &\leq \max_{i=1,\dots,d} |g'(\xi^{(i)}_h a_i) - g'(0)| \|a\|_2 + |g'(0)| \|a\|_2\\ &\leq (\delta + |g'(0)|) \|a\|_2. \end{align*} Analogously, by the inverse triangle inequality $\|\tilde a\|_2 \geq (|g'(0)|-\delta) \|a\|_2$. In particular, \[ \big|\|\tilde a\|_2 / \|a\|_2 - |g'(0)| \big| \leq \delta. \] Now, writing $\gamma = { \ensuremath{\mbox{\rm sign} \,} }(g'(0))$, we observe \begin{align*} \big\|\gamma \hat a - a/\|a\|_2 \big\|_2 &\leq \big\| \gamma \hat a - |g'(0)| a / \|\tilde a\|_2 \big\|_2 + \big\| |g'(0)| a / \|\tilde a\|_2 - a/\|a\|_2 \big\|_2\\ &=\|\tilde a\|_2^{-1} \big( \| (\mathrm{diag}(\xi) - g'(0) I_d) \ a\|_2 + \big| |g'(0)| - \|\tilde a\|_2/\|a\|_2\big| \ \|a\|_2 \big)\\ &\leq 2\delta \|a\|_2 / \| \tilde a\|_2 \leq 2\delta/(|g'(0)|-\delta) \leq 2\delta/(\kappa - \delta) = \varepsilon. \end{align*} \end{proof} Having recovered the ridge direction, we manage to unveil the one-dimensional structure from the high-dimensional ambient space. In other words, recovery of the ridge direction is a \emph{dimensionality reduction} step. What remains is the problem of sampling the profile, which can be done using standard techniques. In combination, this leads to the following result: \begin{theorem}\label{res:ridge_sampling_derivative} Let $\alpha > 1$ and $0 < \kappa \le 1$. \begin{enumerate} \item[(i)] Let $n\le d-1$. Then $g_{n,d}(\Rs{\alpha,2,\kappa},L_{\infty})=g_{n,d}^{\mathrm{lin}}(\Rs{\alpha,2,\kappa},L_{\infty})=1$. \item[(ii)] Let $n\ge d+1$. Then \begin{align*} c_{\alpha} \cdot n^{-\alpha} \leq g_{n,d}(\Rs{\alpha,2,\kappa},L_{\infty}) \le g_{n,d}^{\mathrm{lin}}(\Rs{\alpha,2,\kappa},L_{\infty}) \leq C_\alpha (n-d)^{-\alpha} \end{align*} with constant $c_{\alpha}$ and $C_\alpha$, which depend on $\alpha$ only. \end{enumerate} \end{theorem} \begin{proof} \emph{(i)} It is enough to show that $g_{n,d}(\Rs{\alpha,2,\kappa},L_{\infty})\ge 1$ for $n\le d-1$. Let us assume that a given (adaptive) approximation method samples at $x_1,\dots,x_n$ and let us denote by $L$ their linear span. Then ${\rm dim}\,L\le n<d$ and we may find $a\in{\field R}^d$ with $\|a\|_2=1$ orthogonal to all $x_1,\dots,x_n.$ Finally, if we define $g(t)=t$, we obtain \begin{align*} 1&=\|g(a^T\cdot)\|_\infty\le\frac{1}{2}\cdot\Bigl\{\|g(a^T\cdot)-S_n(g(a^T\cdot))\|_\infty+\|-g(a^T\cdot)-S_n(-g(a^T\cdot))\|_\infty\Bigr\}\\ &\le g_{n,d}(\Rs{\alpha,2,\kappa},L_{\infty}). \end{align*} \emph{(ii)} Fix some $0 < \varepsilon < 1$. Let $\hat{a}$ denote the reconstruction of $a$ obtained by Lemma \ref{lem:reconstruct_a}, which uses $d+1$ sampling points of $f$. We estimate $g$ by sampling the distorted version \[ \tilde g: \intcc{-1,1} \rightarrow \field R, \; t \mapsto f(t\hat a) = g(t a^T \hat a). \] Re-using the value $g(0)$ which we have already employed for the recovery of $a$, we spend $k=n-d\ge 1$ sampling points and obtain a function $\hat g$ with $\|\hat g-\tilde g\|_\infty\le \varepsilon:=C'_\alpha k^{-\alpha}\|\tilde g\|_{\ensuremath{\mathrm{Lip}}_\alpha}$. Now put $\hat f(x) = \hat g(\hat a^T x)$ as our approximation to $f$. To control the total approximation error, observe that \[ | \hat f(x) - f(x) | \leq | \hat g(\hat a^T x) - \tilde g (\hat a^T x) | + | \tilde g(\hat a^T x) - g(a^T x) |=:E_1+E_2. \] For the first error term $E_1$, we immediately get \[ E_1 \leq \| \hat g-\tilde g \|_\infty \leq \varepsilon=C'_\alpha \| \tilde g\|_{\ensuremath{\mathrm{Lip}}_\alpha} k^{-\alpha}\le C_\alpha' k^{-\alpha} \] as $\| \tilde g \|_{\ensuremath{\mathrm{Lip}}_\alpha} \leq \|a\|_2 \, \|g\|_{\ensuremath{\mathrm{Lip}}_\alpha}\le 1$. For the second error term, note that \begin{align*} E_2 &= | g(a^T \hat a \hat a^T x) - g(a^T x) | \leq \| g \|_{\ensuremath{\mathrm{Lip}}_\alpha} \, \| a^T (I_d - \hat a \hat a^T) \|_2 \, \|x\|_2\\ &\leq \|g \|_{\ensuremath{\mathrm{Lip}}_\alpha} \, \|x\|_2 \, \|a\|_2 \, \big\| a^T/\|a\|_2 \ (I_d - \hat a \hat a^T) \big\|_2. \end{align*} We do not know the exact value of the subspace stability term $\| a^T / \|a\|_2 \ (I_d - \hat a \hat a^T) \|_2$. But because $ \hat a \hat a^T$ is the identity in direction of $\hat a$, we have the estimate \begin{align*} \big\| a^T / \|a\|_2 \ (I_d - \hat a \hat a^T) \big\|_2 & = \big\| \big(a / \|a\|_2 - { \ensuremath{\mbox{\rm sign} \,} }(g'(0)) \hat a \big)^T \ (I_d - \hat a \hat a^T) \big\|_2\\ & \leq \| I_d - \hat a \hat a^T \|_{2 \rightarrow 2} \ \big\| a/ \|a\|_2 - { \ensuremath{\mbox{\rm sign} \,} }(g'(0)) \hat a \big\|_2 \\ &\leq \varepsilon. \end{align*} For the last inequality, we have used Lemma \ref{lem:reconstruct_a} and the fact that $\| I_d - \hat a \hat a^T\|_{2 \rightarrow 2} \leq 1$. As a consequence, \[ E_2 \leq \|x\|_2 \, \|a\|_2 \, \|g\|_{\ensuremath{\mathrm{Lip}}_\alpha} \, \varepsilon \leq \varepsilon. \] Putting everything together, we conclude \[ \| \hat f - f \|_{\infty} \leq 2 \varepsilon\le 2C_\alpha'k^{-\alpha}. \] Let us turn to the lower bound. Assume we are given a feasible approximation method $S_n$ that samples at points $\{x_1,\dots, x_n\}\subset \Omega$. Let $\psi_{k,b}$ be as in the proof of Proposition \ref{res:en_univariate_Lip_derivative}. There is an interval $I'\subset I=[\pi/4-1/5,\pi/4+1/5]$ of length $|I'|=1/(5n)$ such that $I'$ does not contain any of the first coordinates of $x_1,\dots,x_n$; in other words, it is disjoint with $\{x_1\cdot e_1,\dots,x_n \cdot e_1\}$, where $e_1=(1,0,\dots,0)$ is the first canonical unit vector. Furthermore, let $b$ be the center of $I'$, put $\psi = \psi_{2n,b}$, and $a=e_1.$ Finally, with $\gamma$ as in \eqref{eq:maxcosin}, we write \begin{align*} f(x)&=\sin(x\cdot e_1),\\ f_+(x)&=\sin(x\cdot e_1)+(1-\gamma)\psi(x \cdot e_1),\\ f_-(x)&=\sin(x \cdot e_1)-(1-\gamma)\psi(x \cdot e_1). \end{align*} As $S_n(f)=S_n(f_+)=S_n(f_-)$ and all the three functions are in $\Rs{\alpha,2,\kappa}$, we may use the triangle inequality \begin{align*} \|(1-\gamma)\psi\|_\infty&=\|(1-\gamma)\psi(e_1\cdot)\|_\infty\\ &\le \frac{1}{2}\Bigl\{ \|(1-\gamma)\psi(e_1\cdot)+f-S_n(f)\|_\infty+\|(1-\gamma)\psi(e_1\cdot)-[f-S_n(f)]\|_\infty\Bigr\}\\ &=\frac{1}{2}\Bigl\{\|f_+-S_n(f_+)\|_\infty+\|f_--S_n(f_-)\|_\infty\Bigr\}, \end{align*} to conclude that \[ g_{n,d}(\Rs{\alpha,2,\kappa},L_{\infty}) \gtrsim n^{-\alpha}, \] with a constant depending only on $\alpha.$ \end{proof} \begin{remark} Once we have control on the derivative in the origin, cf.\ Section \ref{sec:sampling2}, recovery of the ridge direction and approximation of the ridge profile can be addressed independently. Formula \eqref{eq:rec_a} is based on the simple observation that \[ \frac{\partial f}{\partial x_i}(0)=g'(0)a_i=g'(0)\<a,e_i\> \] might be well approximated by first order differences. Furthermore, this holds also for every other direction $\varphi\in\ensuremath{\mathds S^{d-1}}_2$, i.e., \[ \frac{\partial f}{\partial \varphi}(0)=g'(0)\<a,\varphi\> \] can be approximated by differences \[ \frac{f(h\varphi)-f(0)}{h}. \] Taking the directions $\varphi_1,\dots,\varphi_{m_{\Phi}}$ at random (and appropriately normalized), one can approximate the scalar products $\{\<a,\varphi_i\>\}_{i=1}^{m_{\Phi}}$. Finally, if one assumes that $a\in \bar B_p^d$ for $0<p\le 1$, one can recover a good approximation to $a$ by the \emph{sparse recovery} methods of the modern area of \emph{compressed sensing}. This approach has been investigated in \cite{FSV12}. Although the algorithms of compressed sensing involve random matrices, once a random matrix with good sensing properties (typically with small constants of their Restricted Isometry Property) is fixed, the algorithms become fully deterministic. This allows to transfer the estimates of \cite{FSV12} into the language of information based complexity. It follows from the results of \cite{FSV12} that if $0<p\le 1$ and $$ c\kappa^{-\frac{2p}{2-p}}\log d\le m_{\Phi} \le Cd, $$ for two universal positive constants $c,C$, then a function $f\in\Rs{2,p,\kappa}$ might be recovered with high probability up to the precision \[ \Bigl[\frac{m_{\Phi}}{\log(d/m_{\Phi})}\Bigr]^{1/2-1/p}+(n-m_{\Phi})^{-2} \] using $n>m_{\Phi}$ sampling points. If $1/p\le 5/2$ and $c'\kappa^{-\frac{2p}{2-p}}\log d\le n \le C'd,$ this implies that \[ g^{\text{lin}}_{n,d}(\Rs{2,p,\kappa},L_\infty)\lesssim \Bigl[\frac{n}{\log(d/n)}\Bigr]^{1/2-1/p} \] and the same estimate holds if $1/p>5/2$ and $c'\kappa^{-\frac{2p}{2-p}}\log d\le n \le c''(\log d)^{\frac{1/p-1/2}{1/p-5/2}}$. Finally, if $c''(\log d)^{\frac{1/p-1/2}{1/p-5/2}}\le n\le C'd$, we obtain \[ g^{\text{lin}}_{n,d}(\Rs{2,p,\kappa},L_\infty)\lesssim n^{-2}. \] \end{remark} \section{Tractability results} \label{sec:tractability} For the classification of ridge function sampling by degrees of difficulty, the field of information-based complexity \cite{NW1} provides a family of notions of so-called \emph{tractability}. Despite of their simple structure, ridge functions lead to a surprisingly rich class of sampling problems in regard of these notions: we run across almost the whole hierarchy of degrees of tractability if we vary the problem parameters $\alpha$ and $p$, or add the constraint on the profiles' first derivative in the origin. Let us briefly introduce the standard notions of tractability. We say that a problem is \emph{polynomially tractable} if its information complexity $n(\varepsilon,d)$ is bounded polynomially in $\varepsilon^{-1}$ and $d$, i.e. there exist numbers $c,p,q > 0$ such that \[ n(\varepsilon,d) \leq c \, \varepsilon^{-p} \, d^q \mbox{ for all $0<\varepsilon<1$ and all $d\in{\N}$.} \] A problem is called \emph{quasi-polynomially tractable} if there exist two constants $C,t>0$ such that \begin{equation}\label{eq:qptrac} n(\varepsilon,d) \leq C\exp(t(1+\ln(1/\varepsilon))(1+\ln d))\,. \end{equation} It is called \emph{weakly tractable} if \begin{equation}\label{eq:wtrac} \lim\limits_{1/\varepsilon+d\to\infty} \frac{\log n(\varepsilon,d)}{1/\varepsilon + d} = 0\,, \end{equation} i.e., the information complexity $n(\varepsilon,d)$ neither depends exponentially on $1/\varepsilon$ nor on $d$. We say that a problem is \emph{intractable}, if \eqref{eq:wtrac} does not hold. If for some fixed $0<\varepsilon<1$ the number $n(\varepsilon,d)$ is an exponential function in $d$ then a problem is, of course, intractable. In that case, we say that the problem suffers from {\em the curse of dimensionality}. To make it precise, we face the curse if there exist positive numbers $c, \varepsilon_0, \gamma$ such that \begin{equation}\label{eq:curse} n(\varepsilon,d) \geq c(1+\gamma)^d\,,\quad \mbox{for all } 0<\varepsilon\leq \varepsilon_0 \mbox{ and infinitely many }d\in \N\,. \end{equation} In the language of IBC, Theorems \ref{res:ridge_sampling} and \ref{res:ridge_sampling_derivative} now read as follows: \begin{theorem}\label{res:tractability} Consider the problem of ridge function sampling as defined in Subsection \ref{sec:ibc}. Assume that ridge profiles have at least Lipschitz smoothness $\alpha > 0$; further, assume that ridge directions are contained in the closed $\ell_p^d$-unit ball for $p \in \intoc{0,2}$. Then sampling of ridge functions in the class $\Rs{\alpha,p}$ \begin{enumerate}[label=(\arabic*)] \item suffers from the curse of dimensionality if $p=2$ and $\alpha<\infty$, \item never suffers from the curse of dimensionality if $p < 2$, \item is intractable if $p<2$ and $\alpha \le \frac{1}{1/p-1/2}$, \item is weakly tractable if $p<2$ and $\alpha > \frac{1}{1/\max\{1,p\}-1/2}$, \item is quasi-polynomially tractable if $\alpha = \infty$, \item and with positive first derivatives of the profiles in the origin it is polynomially tractable, no matter what the values of $\alpha$ and $p$ are. \end{enumerate} \end{theorem} To prove Theorem \ref{res:tractability}, we translate Theorem \ref{res:ridge_sampling} into bounds on the information complexity \[ n(\varepsilon,d) = \min \{n \in \N: \; g_{n,d}(\Rs{\alpha,p},L_{\infty}) \leq \varepsilon\}. \] \begin{lemma}\label{logn_upper_bounds} Let $p < 2$ and $\alpha > 0$. Set $\eta = \alpha(1/2 - 1/p') = \alpha(1/\max\{1,p\} - 1/2)$ and define \begin{align*} \varepsilon_1^U := C_{p,\alpha} \left[ \frac{\log(1 + d/\log d)}{\log d}\right]^{\eta}, & \qquad \varepsilon_2^U := C_{p,\alpha} \del{\frac 1d}^{\eta}. \end{align*} Then there are positive constants $C_0$ and $C_1$ such that \begin{align*} \log n(\varepsilon,d) \leq C_0 + C_1 \begin{cases} \log d &: \varepsilon_1^U \leq \varepsilon \leq 1,\\ \log d \cdot (1/\varepsilon)^{1/\eta}&: \varepsilon_2^U \leq \varepsilon < \varepsilon_1^U,\\ \log(1/\varepsilon) \cdot (1/\varepsilon)^{1/\eta} &: \varepsilon < \varepsilon_2^U. \end{cases} \end{align*} The constants depend only on $p$ and $\alpha$. \end{lemma} \begin{detail} \begin{proof} For auxiliary reasons, let us introduce \[ \varepsilon_3^U := 2^{-\alpha(d+1)/d} {d+s \choose s}^{-\alpha/d} \varepsilon_2^U, \] as well as the shorthand $\eta = \alpha(1/2 - 1/p') = \alpha(1/\max\{1,p\} - 1/2)$. \begin{itemize} \item For $N_1 := 2d{d+s \choose s}$ we have $g_{N_1,d} \leq \varepsilon_1^U$. By monotonicity of information complexity, for all $\varepsilon \geq \varepsilon_1^U$ \[ \log_2 n(\varepsilon,d) \leq \log_2 n(\varepsilon_1^U,d) \leq \log_2 N_1 \leq 1 + s \log(1+s) + (1+s)\log d. \] \item Consider $\varepsilon_2^U \leq \varepsilon < \varepsilon_1^U$. Write \begin{align*} \varepsilon(n) := C_{p,\alpha} \left[ \frac{ \log_2(1+d/\log_2 n_1) }{\log_2 n_1} \right]^\eta \leq C_{p,\alpha} \left[ \frac{ \log_2(1+d)}{\log_2 n_1} \right]^\eta\\ \Rightarrow \log n \leq 1 + \log {d+s \choose s} + C_{p,\alpha}^{1/\eta} \log(1+d) \varepsilon(n)^{-1/\eta}. \end{align*} There is a natural $n$ with $N_1 < n,n+1 \leq 2^{d+1} {d+s \choose s} =: N_2$ such that \[ \varepsilon(n+1) \leq \varepsilon \leq \varepsilon(n). \] By monotonicity of information complexity, \begin{align*} \log_2 n(\varepsilon,d) & \leq \log_2 n(\varepsilon(n+1),d) \leq \log_2(n+1) \leq 2 \log_2 n\\ & \leq 2 \big(1 +s\log(s+1) + s\log(d) + C_{p,\alpha}^{1/\eta} \varepsilon(n+1)^{-1/\eta} \log(1+d) \big)\\ & \leq 2 + 2s \log(s+1)+ 2(s+2C_{p,\alpha}^{1/\eta}) \varepsilon^{-1/\eta} \log d. \end{align*} \item Assume $\varepsilon < \varepsilon_3^U$. Find $n,n+1 > 2^{d+1} {d+s \choose s}$ such that $\varepsilon(n+1) \leq \varepsilon \leq \varepsilon(n)$. Let $\varepsilon(n) := C_{p,\alpha} n^{-\alpha/d} d^{-\eta}$. Since $\varepsilon(n) < \varepsilon^U_3 < \varepsilon_2^U$ we have $d < C_{p,\alpha}^{1/\eta} \varepsilon(n)^{-1/\eta}$, which yields \begin{align*} \log_2 n \leq \alpha^{-1} (\log_2 C_ {p,\alpha} + 1) \log_2(1/\varepsilon(n)) (1/\varepsilon(n))^{1/\eta}. \end{align*} By monotonicity of information complexity, \begin{align*} \log_2 n(\varepsilon,d) &\leq \log_2 n(\varepsilon(n+1),d) \leq \log_2 (n+1) < 2\log_2 n\\ &\leq 2\alpha^{-1} (\log_2 C_{p,\alpha} + 1) \log_2(1/\varepsilon(n)) (1/\varepsilon(n))^{1/\eta}\\ &\leq 2\alpha^{-1} (\log_2 C_{p,\alpha} + 1) \log_2(1/\varepsilon) (1/\varepsilon)^{1/\eta}. \end{align*} \item Assume $\varepsilon_3^U \leq \varepsilon < \varepsilon_2^U$. Put $B = 2^{\alpha(d+1)/d} {d+s \choose s}^{\alpha/d}$. Then \[ \varepsilon_3^U/B \leq \varepsilon/B < \varepsilon_3; \] thus \begin{align*} \log n(\varepsilon,d) \leq \log n(\varepsilon/B,d) \leq \alpha^{-1} (\log_2 C_{p,\alpha} + 1) \log_2(B/\varepsilon) (B/\varepsilon)^{1/\eta}. \end{align*} The constant $B$ can be bounded from above by a universal constant independent of $d$ and $n$. Hence we get \[ \log_2 n(\varepsilon,d) \leq \tilde C_{p,\alpha} \alpha^{-1} (\log_2 C_{p,\alpha} + 1) \log_2(1/\varepsilon) (1/\varepsilon)^{1/\eta}. \] \item Set \begin{align*} C_0 &:= 2 + 2s\log(1+s),\\ C_1 &:= 2\max \big\{ s+1, s+2C_{p,\alpha}^{1/\eta}, \tilde C_{p,\alpha} \alpha^{-1} (\log_2 C_{p,\alpha} + 1) \big\}. \end{align*} \end{itemize} \end{proof} \end{detail} \begin{lemma}\label{logn_lower_bounds} Let $p < 2$ and $\alpha > 0$. Put \begin{align*} \varepsilon_1^L := c_{p,\alpha} \left[ \frac{\log(1 + d/\log d)}{\log d}\right]^{\alpha(1/p-1/2)}, & \quad \varepsilon_2^L := c_{p,\alpha} \del{\frac 1d}^{\alpha(1/p-1/2)}, & \varepsilon_3^L := 4^{-\alpha} \varepsilon_2^L. \end{align*} Then there are universal constants $c_0$, $c_1$, which depend only on $p$ and $\alpha$, such that \begin{align*} \log n(\varepsilon,d) \geq c_0 + c_1 (1/\varepsilon)^{\alpha^{-1} (1/p - 1/2)^{-1}} \end{align*} for $\varepsilon_3^L \leq \varepsilon < \varepsilon_1^L$. \end{lemma} \begin{detail} \begin{proof}[Proof of Lemma \ref{logn_lower_bounds}] Put $\gamma= \alpha (1/p - 1/2)$. \begin{itemize} \item Consider $\varepsilon_2^L \leq \varepsilon < \varepsilon_1^L$. Put \[ \varepsilon(n) = c_{p,\alpha} \left[\frac{\log_2(1 +d/(\log_2 n +2)}{\log_2 n + 2}\right]^{\gamma} \] for $d < n \leq 2^{d-1}$. With $1/(\log_2 n + 2) \geq 1/(d+1) \geq 1/(2d)$ we have \[ \varepsilon(n) \geq c_{p,\alpha} \left[\frac{\log_2(3/2)}{\log n + 2}\right]^{\gamma}. \] Now choose $d < n,n+1 \leq 2^{d-1}$ such that $\varepsilon(n+1) \leq \varepsilon \leq \varepsilon(n)$. By monotonicity of information complexity, we have \begin{align*} \log_2 n(\varepsilon,d) &\geq \log_2 n( \varepsilon(n),d) \geq \log_2 n \geq 1/2 \log_2(n+1)\\ &\geq -1 + 1/2c_{p,\alpha}^{1/\gamma} \frac{\log(3/2)}{\varepsilon(n+1)^{1/\gamma}} \geq -1 + 1/2c_{p,\alpha}^{1/\gamma} \frac{\log(3/2)}{\varepsilon^{1/\gamma}}. \end{align*} \item Assume $\varepsilon_3^L \leq \varepsilon < \varepsilon_2^L$. Then $\varepsilon_2^L \leq 4^{\alpha}\varepsilon$. \[ \log_2 n(\varepsilon,d) \geq \log_2 n(4^{\alpha}\varepsilon,d) \geq -1 + 1/2 c_{p,\alpha}^{1/\gamma} 4^{-\alpha/\gamma} \frac{\log_2(3/2)} {\varepsilon^{1/\gamma}}. \] \item Put $c_0 := -1$, $c_1 := 1/2 c_{p,\alpha}^{1/\gamma_{p,\alpha}} 4^{-\alpha/\gamma}$. \end{itemize} \end{proof} \end{detail} \begin{proof}[Proof of Theorem \ref{res:tractability}] (1). For $n \leq 2^{d-2}$, the lower bound in Theorem \ref{res:ridge_sampling} gives \[ g_{n,d}(\Rs{\alpha,2},L_{\infty}) \geq c_{p,\alpha} =: \varepsilon_0.\] Hence, $n(\varepsilon,d) \geq 2^{d-2}$ for all $\varepsilon < \varepsilon_0$ and we have the curse of dimensionality. (2). Since $\alpha_1 > \alpha_2$ implies $\Rs{\alpha_1,p} \subseteq \Rs{\alpha_2,p}$, we can w.l.o.g. assume $\alpha \leq 1$. We choose an arbitrary $\varepsilon_2^U \leq \varepsilon \leq 1$. By Lemma \ref{logn_upper_bounds}, \[ n(\varepsilon,d) \leq 2^{C_0} d^{ C_1 \varepsilon^{\alpha^{-1} (1/\max\{1,p\} - 1/2)^{-1}}}. \] By our assumption $\varepsilon \geq \varepsilon_2^U$, this is true for all natural $d > (C_{p,\alpha} /\varepsilon)^{\alpha^{-1} (1/\max\{1,p\} - 1/2)^{-1}}$. Hence, the curse of dimensionality does not occur. (3). Put $\gamma = \alpha (1/p - 1/2)$. Assume $d \to \infty$ and $\varepsilon_3^L \leq \varepsilon< \varepsilon_2^L$. The latter implies \[ \left(\frac{c_{p,\alpha}}{4^\alpha}\right)^{1/\gamma} (1/\varepsilon)^{1/\gamma} \leq d < c_{p,\alpha}^{1/\gamma} (1/\varepsilon)^{1/\gamma}. \] This yields \[ \frac{ \log_2 n(\varepsilon,d) }{ d+1/\varepsilon } \geq \frac{c_0}{d+1/\varepsilon} + c_1 \frac{ (1/\varepsilon)^{1/\gamma}}{ c_{p,\alpha}^{1/\gamma} (1/\varepsilon)^{1/\gamma} + 1/\varepsilon}. \] Assuming that $\alpha \leq 1/(1/p - 1/2)$, we have $\gamma \leq 1$ and thus $1/\varepsilon \leq (1/\varepsilon)^{1/\gamma}$. We conclude that \[ \frac{ \log n(\varepsilon,d) }{ d+1/\varepsilon } \geq \frac{c_1}{c_{p,\alpha}^{1/\gamma} + 1 } > 0. \] Consequently, the problem is not weakly tractable; and thus intractable. (4). Put $x= 1/\varepsilon + d$. By Lemma \ref{logn_upper_bounds} and $1/\varepsilon\leq x$, $d \leq x$, we have \begin{align*} \log n(\varepsilon,d) \leq C_0 + C_1 \log(x) x^{\alpha^{-1} (1/\max\{1,p\} - 1/2)^{-1}}. \end{align*} Now, if $\alpha > \frac{1}{1/\max\{1,p\}-1/2}$, then $\lim_{x\rightarrow \infty}\ x^{-1} \log n(\varepsilon,d) = 0$. (5). By embedding arguments it is enough to consider the class $\Rs{\infty,2}$. We approximate the function $f \in \Rs{\infty,2}$ via the Taylor polynomial $T_{s,0}f(x)$ in $x^0 = 0$. Lemma \ref{lem:taylorapprox}, (ii) gives for every $s\in \N$ the bound $$ \|f-T_{s,0}f\|_{\infty} \leq \frac{2}{s!}\,. $$ Let $\varepsilon>0$ be given and let $s\in \N$ be the smallest integer such that $2/s! \leq \varepsilon$. Then $(s-1)! \leq 2/\varepsilon$ and therefore $[(s-1)/e]^{s-1} \leq (s-1)! \leq 2/\varepsilon$. This gives \begin{equation}\label{f61} (s-1)\ln((s-1)/e) \leq \ln(2/\varepsilon)\,. \end{equation} We know from \cite{V2013} that it requires $\binom{s+d}{s}$ function values to approximate the Taylor polynomial up to arbitrary (but fixed) precision. Hence, using \eqref{f61}, we see that there is a constant $t>0$ such that $$ \ln n(\varepsilon,d) \leq s\ln(e(d+1)) \leq t(1+\ln(1/\varepsilon))(1+\ln d), $$ which is \eqref{eq:qptrac}. (6). From Theorem \ref{res:ridge_sampling_derivative} we can immediately conclude $\varepsilon^{-1/\alpha} \lesssim n(\varepsilon,d) \lesssim \varepsilon^{-1/\alpha}$, where the constants behind ``$\lesssim$'' behave polynomially in $d$. Consequently, sampling of ridge functions in $\Rs{\alpha,2,\kappa}$ is polynomially tractable. \end{proof} By Lemma \ref{emb}, we know that $\Rs{\infty,2}$ is a subclass of the unit ball in $C^{\infty}(\Omega)$. Besides, we know that approximation using function values is quasi-polynomially tractable in $\Rs{\infty,2}$, see Theorem \ref{res:tractability}. What is the respective tractability level in $C^{\infty}(\Omega)\,$? Or, to put it differently: how much do we gain by imposing a ridge structure in $C^{\infty}(\Omega)\,$? The seminal paper \cite{NW_2009_2} tells us that approximation in $C^{\infty}([0,1]^d)$ suffers from the curse of dimensionality when norming the space in the way as we did in \eqref{f62}. In contrast, we will show that sampling in $C^{\infty}(\Omega)$ is still weakly tractable. This is not too much of a surprise: due to the concentration of measure phenomenon, the Euclidean unit ball's volume is getting ``very small'' in high dimensions $d$; its measure scales like $(2\pi e/d)^{d/2}$. Anyhow, the result suggests that one still benefits from supposing a ridge structure; infinitely differentiable ridge functions from $\Rs{\infty,2}$ probably can be approximated easier than general functions from the unit ball of $C^{\infty}(\Omega)$. This is not guaranteed, however, because we do not show that one cannot get anything better than weak tractability for the sampling of functions in the unit ball of $C_{\infty}(\Omega)$. \begin{theorem}\label{res:infty} \label{wtrinf} The sampling problem for $C^{\infty}(\Omega)$, where the error is measured in $L_{\infty}(\Omega)$, is weakly tractable. \end{theorem} \begin{proof} Applying Lemma \ref{lem:taylorapprox}, (i) together with \eqref{f60} we obtain for any $f\in C^{\infty}(\Omega)$ with $\|f\|_{C^{\infty}(\Omega)}\leq 1$ and every $s\in \N$ the relation \begin{align}\nonumber |f(x)-T_{s,0}f(x)| &\leq \frac{2}{(s-1)!}\|x\|_1^s\quad,\quad x\in \Omega\,,\\\nonumber &\leq \frac{2d^{s/2}}{(s-1)!} \,. \end{align} Let $s\in \N$ be the smallest integer such that $2d^{s/2}/(s-1)! \leq \varepsilon$. This leads to $$ \frac{1}{\sqrt{d}}\Big(\frac{s-2}{e\sqrt{d}}\Big)^{s-2} \leq \frac{(s-2)!}{d^{\frac{s-1}{2}}} \leq \frac{2}{\varepsilon}\, $$ which implies \begin{equation}\label{f65} (s-2)\ln\Big(\frac{s-2}{e\sqrt{d}}\Big) \leq \ln(2/\varepsilon)+\frac{1}{2}\ln(d)\,. \end{equation} To approximate the Taylor polynomial $T_{s,0}f$ with arbitrary precision (uniformly in $f$) we need $\binom{d+s}{s}$ function values, see \cite[p.\ 4]{V2013}. Let us distinguish two cases. If $(s-2) \leq e^2\sqrt{d}$ we obtain $$ \ln n(\varepsilon,d) \leq s\ln(e(d+1)) \leq (e^2\sqrt{d}+2)\cdot\ln(e(d+1)) $$ and hence \eqref{eq:wtrac}\,. If $s-2 > e^2\sqrt{d}$ then \eqref{f65} yields $s-2 \leq \ln(2/\varepsilon)+\ln(d)$. Thus, $$ \ln n(\varepsilon,d) \leq s\ln(e(d+1)) \leq (\ln(2/\varepsilon)+\ln(d)+2)\cdot \ln(e(d+1)) $$ and again \eqref{eq:wtrac} holds true. This establishes weak tractability. \end{proof} \begin{remark} {(i)} The result in Theorem \ref{wtrinf} is also a consequence of the arguments in \cite[Sections 5.2, 5.3, and Section 6]{HiNoUlWo13} by putting $L_{j,d} = d^{j/2}$. \\ {(ii)} Recently, Vyb\'iral \cite{V2013} showed that there is quasi-polynomial tractability if one replaces the classical norm $\sup_{ \gamma \in \N_0^d} \|D^{\gamma} f\|_{\infty}$ by $\sup_{k\in \N_0}\sum_{|\gamma| = k}\|D^{\gamma}f\|_{\infty}/\gamma !$ in $C^{\infty}([0,1]^d)$. In contrast to that, Theorem \ref{res:infty} shows weak tractability for the classical norm on the unit ball. \end{remark} \textbf{Acknowledgments} The authors would like to thank Aicke Hinrichs, Erich Novak, and Mario Ullrich for pointing out relations to the paper \cite{HiNoUlWo13}, as well as Sjoerd Dirksen, Thomas K\"uhn, and Winfried Sickel for useful comments and discussions. The last author acknowledges the support by the DFG Research Center {\sc Matheon} ``Mathematics for key technologies'' in Berlin.
\section{Introduction} It is believed that the star-formation process should be dominated by gravitational infall motions that accrete material from the ambient cloud into a central protostellar object. Establishing the nature of infall motions and distinguishing them from other systematic motions in the cloud are not easy tasks, and they have been the subject of many papers in the last decades. However, it is not only important to detect signatures of infall motions, but also to obtain spatially-resolved information that allows us to further investigate the kinematics and physical parameters of the molecular core around the protostar. Characterizing the infall signatures acquires special interest in the case of the formation of massive stars, where different scenarios have been proposed (merging of less massive stars, competitive accretion, or monolithic collapse; e.g., \citealt{zin07}) and, therefore, where obtaining spatially-resolved kinematic information can help to discriminate between these scenarios. \citet{sne77} proposed as evidence of collapse the asymmetric self-reversed CO line profiles seen towards some embedded stars in dense interstellar clouds. Their modelling of a contracting cloud predicts a double-peaked CO line profile where the dip between the two peaks is red-shifted with respect to the systemic velocity of the cloud. Soon afterwards, this result was criticized by \citet{leu77}, who argued that the Sobolev approximation used in the \citet{sne77} calculations was invalid for the velocity fields considered, and concluded that the signature was not unequivocal for collapse. \citet{ang87} studied the infall signatures of an angularly unresolved spherically symmetric protostellar core under the assumptions of optically thick line emission and gravitational infall motions dominating the kinematics over turbulent and thermal motions. Assuming that in the inside-out gravitational collapse towards a central object the velocity increases as the radius decreases, the points with the same line-of-sight (l.o.s.) velocity will form closed surfaces (isovelocity surfaces) \citep[][see Fig. 1a]{kui78, ang87}. All the isovelocity surfaces have the same shape and decrease in size with increasing magnitude of the l.o.s. velocity. The observed flux density as a function of l.o.s. velocity (i.e., the line profile) is determined by the integrated intensity of the corresponding isovelocity surfaces. In general, a given line-of-sight intersects the same isovelocity surface twice. If the opacity is high enough, the observed emission is dominated by the emission coming from the side of the isovelocity surface facing the observer (thick line in Fig. 1a), while the emission from the rear side (thin line in Fig. 1a) remains hidden. Since in protostellar collapse the temperature increases towards the centre of the core, it turns out that the blue-shifted emission comes from points that, on the average, are closest to the protostar (and, therefore, hotter) than the corresponding red-shifted ones (see Fig. 1a). This results in asymmetric line profiles, with the blue-shifted side stronger than the red-shifted one. The assumption of a large opacity gradient allows to adopt a simple approach similar to the Sobolev approximation (see Appendix A.2 in \citealt{ang87}), as only a very thin edge of the isovelocity surface, where physical properties are unlikely to vary significantly, would be observable. Optically thin lines would have symmetric profiles. \begin{figure*} \includegraphics[width=\linewidth]{fig1.eps} \caption{{\em (a)} Surfaces of equal line-of-sight velocity for a collapsing protostellar envelope. The (0,0) coordinates correspond to the position of the central protostar. Note that a given line-of-sight intersects an isovelocity surface at two points. If the emission at each velocity is optically thick, only the part of the surface facing the observer (thick line) is observable, while the rear part (thin line) is not observable. {\em (b)} Intensity as a function of angular offset from the source centre for channel maps of different velocities. Channel maps are, in fact, images of the temperature distribution in the corresponding isovelocity surfaces. For red-shifted channels, the centre of the image corresponds to a region slightly farther away from the centre of the core (and, thus, slightly colder) than the edges. Conversely, for blue-shifted channels, the centre of the image arises from a region very close to the centre of the core (and, therefore, very hot). This produces very different images in red-shifted and blue-shifted channels. In a red-shifted channel, the intensity is almost constant over the emitting region (decreasing slightly at the centre) while in the corresponding blue-shifted channel the intensity increases sharply towards the centre of the image. Additionally, the emission becomes more compact as the absolute value of the velocity increases (from \citealt{ang91}). Calculations correspond to a simple model for an 1 M$_\odot$ protostar at 160 pc observed with a beam of $0\farcs2$ at the time when a central mass of $\sim0.5$ M$_\odot$ has been accreted.} \label{fig1} \end{figure*} \citet{zho93} reported observations of molecular lines towards B335 showing asymmetric line profiles with the blue side stronger than the red side. This result was considered as evidence of gravitational protostellar infall since the strengths and profiles of the observed lines were in agreement with the predictions of the inside-out collapse model of \citet{shu77}. Since then, numerous surveys have been carried out and ``blue asymmetries'' in line profiles have been extensively interpreted as infall signatures (e.g., \citealt{gre97}; \citealt{mar97}; \citealt{lee01}; \citealt{full05}; \citealt{chen10}), although in most cases a detailed comparison with the predictions of a self-consistent collapse model has not been performed. In several cases, inverse P-Cygni profiles have been observed in molecular lines against a background bright HII region (e.g., \citealt{ket87,ket88}; \citealt{ho96}; \citealt{zha97}; \citealt{zha98}) or against the bright dust continuum emission of the hot protostellar core itself (e.g., \citealt{dif01}; \citealt{gir09}). These red-shifted absorption features are indicative of red-shifted motions of foreground gas and, along with additional evidence, have been interpreted as infall motions of the surrounding envelope. However, the ``spectral signatures'' of infall are mostly based on the line profile of the overall observed emission, and do not provide enough information on the spatial structure of the source. So, other scenarios such as rotation, outflow, intrinsic asymmetries in the source, or the relative motion of unbound clouds are capable of causing similar features. A more complete information can be obtained from ``mapping signatures'' in images that angularly resolve the source. The use of centroid velocity maps, that give some spatial information, has been proposed to identify rotation signatures (e.g., the so-called ``blue bulge'' signature) in infalling cores (e.g., \citealt{ade88}; \citealt{nay98}) but in this case the shape of the line profile is not taken into account, resulting in an incomplete information on the kinematic and physical structure of the core. A more simple and complete mapping signature of infall can be obtained from images that angularly resolve the emission as a function of velocity. \citet{ang91} (hereafter A91) investigated the signatures of infall in the images obtained from molecular line observations. These authors calculated the predicted intensity maps for different l.o.s. velocities (channel maps) following the same assumptions considered in \citet{ang87}. The intensity map at a given l.o.s. velocity results from the emission of the corresponding isovelocity surface and, if the opacity is high enough, it is in fact an image of the excitation temperature distribution in the side of this isovelocity surface facing the observer. For red-shifted channels, the centre of the image corresponds to a region that is slightly farther away from the centre of the core (and, thus, colder) than the edges (Fig. 1a). Conversely, for blue-shifted channels, the centre of the image corresponds to the emission from a region very close to the centre of the core (and, therefore, very hot). This produces very different images in red-shifted and blue-shifted channels. In the image of a red-shifted channel, the intensity is almost constant over the emitting region, decreasing slightly towards the centre, while in the corresponding blue-shifted channel the intensity increases sharply towards the centre of the image (see Fig. 1b). Additionally, since the size of the isovelocity surfaces decreases with increasing (relative) velocity, the emission becomes more compact at higher (relative) velocities for both, red- and blue-shifted channels. The detectability of these infall signatures depends strongly on the beam to source size ratio since one needs to resolve angularly the infalling gas for this purpose. Here we present evidence of the signatures proposed by A91, as well as additional infall and rotation mapping signatures in the hot molecular core (HMC) close to the ultra-compact HII (UCHII) region G31.41+0.31 (hereafter G31 HMC). HMCs are small, dense, hot molecular clumps usually found in the proximity of UCHII regions, and are believed to trace one of the earliest observable stages in the life of massive stars. G31 HMC, located at a distance of 7.9 kpc (\citealt{cesa98}), is one of the hottest molecular cores discovered so far, and it probably harbours an O type protostar. It exhibits strong ammonia emission of high excitation transitions (\citealt{chur90}; \citealt{cesa92,cesa98}), making G31 HMC one of the few sources where a high signal-to-noise ratio analysis of the spatially resolved molecular emission can be attempted. \citet{cesa10} reported a double radio continuum source towards the centre of the ammonia core, that could trace the two jet lobes from a single central protostar or a binary system. Infall motions in G31 HMC have already been inferred by \citet{gir09} through the detection of an inverse P-Cygni profile against the bright dust emission peak. \citet{oso09} carried out a detailed modelling of the source. This model, which consists of a spherical envelope collapsing onto an O type star, is able to reproduce the observed spectral energy distribution (SED) of the source as well as the spectra of the ammonia (4,4) transition obtained in the subarcsecond angular resolution VLA observations carried out by \citet{cesa98}. In this paper we present new, high angular resolution VLA observations of the (2,2) to (6,6) ammonia transitions that reveal clear differences between the red- and blue-shifted channel maps. In Section 2 we describe our observations. In Section 3 we compare the observational results with those of the simple model of protostellar infall by A91 and with those of the more detailed modelling of G31 HMC by \citet{oso09}; we introduce an additional infall signature (the ``central blue spot''), easily identifiable in the first-order moment maps; and we study the rotation signatures in the spatial intensity profiles, and their presence in G31 HMC. In Section 4 we give our conclusions. \section{Observations} Observations of the ammonia (2,2), (3,3), (5,5) and (6,6) inversion transitions were carried out on 2009, March 1 and 3, with the VLA of the NRAO\footnote{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} in the B-configuration (VLA Project Code: AM981) providing an angular resolution of $\sim0\farcs3$. The phase centre was set to RA(J2000) = $18^{h} 47^{m} 34.506^{s}$; DEC(J2000) = $-01\degr 12\arcmin 42.97\arcsec$. We used the 4IF correlator mode with a bandwidth of 6.25 MHz ($\sim$80 km s$^{-1}$) and 31 channels of 195 kHz ($\sim$2.4 km s$^{-1}$) width each, plus a continuum channel that contains the central 75 per cent of the total bandwidth. This configuration allowed us to observe two inversion transitions simultaneously. To make sure that at least the main line and one pair of satellites fall within the observational bandwidth we centred the observation at a velocity in between the main line and the inner red-shifted satellite line. Phase, flux and bandpass calibrators were 1851+005, 3C286, and 1773-130 respectively. Pointing corrections were derived from X-band observations for all the calibrators and applied on-line. At the time of the observations, the VLA interferometer was going through a transition period to turn into the EVLA and 20 out of 27 antennas at that time were already EVLA antennas while the rest remained VLA antennas. The frequencies for the NH$_{3}$(5,5) and NH$_{3}$(6,6) inversion lines fell very close to the edge of the bandwidth of the VLA receivers and only the EVLA antennas were used for these two transitions. For the NH$_{3}$(2,2) and NH$_{3}$(3,3) inversion lines both EVLA and VLA antennas were used. Due to the narrowness of the bandwidth used, and because lines in this source are extremely broad, line-free channels were not enough to subtract the continuum properly and additional continuum observations (VLA Project Code: AM994) were carried out on 2009, May 16 for this purpose. In this continuum observation we used the same flux and phase calibrators as in the line observations. We also analysed VLA archival data of the NH$_{3}$(4,4) transition taken in the A configuration (VLA Project Code: AC748; \citealt{cesa10}). These observations were carried out on 2004, October 15, 23, and November 5, using the Fast Switching technique with a 80/40 sec source-calibrator cycle. The bandwidth was 12.5 MHz ($\sim$160 km s$^{-1}$), with a channel width of 195 kHz ($\sim$2.4 km s$^{-1}$). For all the transitions, data editing and calibration were carried out using the Astronomical Image Processing System (AIPS) package of NRAO following the standard high-frequency VLA procedures. We also followed the advice provided by the NRAO for observations with mixed EVLA and VLA antennas. Continuum was subtracted using the task UVSUB of AIPS. This correction was important only towards the position of the UCHII region and was negligible towards the HMC. Maps were obtained with natural weighting and were restored with a circular beam. For the (4,4) transition, that was observed in the A configuration, maps were obtained using a $uv$ tapering of 1300 k$\lambda$ to improve the signal-to-noise ratio. The observational set-up for all the line observations is summarized in Table~1. \begin{table*} \begin{center} \caption{Observational parameters of the ammonia observations\label{tab1}} \begin{tabular}[\textwidth]{cccccccc} \\ \hline \hline & & &{Rest} & &{Spectral} &{Synthesized} &\\ {Transition} &{VLA} &{Antenna$^{a}$} &{Frequency} &{Bandwidth} &{Resolution} &{Beam$^{b}$} &{rms}\\ {($J$,$K$)} &{Configuration} &{Type} &{(GHz)} &{(MHz)} &{(km~s$^{-1}$)} &{($''$)} &{(mJy~beam$^{-1}$)} \\ \hline (2,2) & B & VLA(7)/EVLA(20) &23.7226333 &6.25 & 2.468 &0.33 &0.7 \\ (3,3) & B &VLA(7)/EVLA(20) &23.8701292 &6.25 & 2.453 &0.33 &0.7 \\ (4,4) & A &VLA(27) &24.1394163 &12.5 & 2.426 &0.16$^{c}$ &0.5 \\ (5,5) & B &EVLA(20) &24.5329887 &6.25 & 2.386 &0.37 &0.9 \\ (6,6) & B &EVLA(20) &25.0560250 &6.25 & 2.337 &0.34 &1.0 \\ \hline \end{tabular} \end{center} $^{a}${Numbers in parenthesis indicate the number of antennas of either type.} \\ $^{b}${HPBW of the circular restoring beam.} \\ $^{c}${Interferometric visibilities have been tapered to increase the signal-to-noise ratio.} \end{table*} In Fig. 2 we show an overlay of the maps of the zero-order moment (integrated intensity) and first-order moment (intensity weighted mean velocity) of the observed ammonia (2,2), (3,3), (4,4), (5,5), and (6,6) main line emission. \begin{figure*} \includegraphics[height=0.93\textheight]{fig2.eps} \caption{Overlay of the integrated intensity (zero-order moment; contours) and the intensity weighted mean velocity (first-order moment; color scale) maps of the ammonia (2,2), (3,3), (4,4), (5,5), and (6,6) main line towards G31 HMC. Contour levels are 25, 30, 45, 60, 75, 90, 105, 120, 135, 150, 165, 180, 195, 210 mJy beam$^{-1}$ km s$^{-1}$. The synthesized beam is shown in the upper left corner of each map. Color scale ranges from 93.2 to 102 km s$^{-1}$. Note the spot of blue-shifted emission towards the integrated intensity peak in all the maps. } \label{fig2} \end{figure*} \section{Results and discussion} We detected ammonia emission in all the five observed transitions. This is the first interferometric detection of the ammonia (3,3), (5,5), and (6,6) transitions in this source. In the maps (Fig. 2), the ammonia emission appears as a compact condensation, clearly associated with the double radio continuum source (angular separation = $0\farcs2$) reported by \citet{cesa10}. The peak brightness temperatures are $\sim$100-150 K for all the observed ammonia transitions, similar to the values obtained previously for the (4,4) transition \citep{cesa98,cesa10}. The emission of the main line is optically thick in the five transitions studied, with satellite to main line intensity ratios close to unity. Lines are extremely broad ($\sim$ 13 km s$^{-1}$), so that the inner and outer satellite lines blend together even for the (2,2) transition, where the separation in velocity between satellite lines is $\sim$ 9 km s$^{-1}$. In this paper we present an analysis of the intensity as a function of the projected distance to the core centre (which we call as ``spatial intensity profile''), in order to identify possible differences between red-shifted and blue-shifted channels. A more extensive analysis of the observations will be presented in a future paper (Mayen-Gijon et al. in preparation). \subsection{Spatially resolved infall signatures} \subsubsection{Spatial intensity profiles} To search for the spatially resolved infall signatures (see section 1) we need to compare the observed intensity as a function of distance to the centre of the core for pairs of velocity channels symmetrically located at both sides of the systemic velocity. In order to improve the signal-to-noise ratio, for every channel map we averaged the emission of the main line over circular annuli of different radii and 0$\farcs$15 width, and then we plotted the average intensity as a function of the radius. The centre was chosen to be at RA(J2000) = $18^{h} 47^{m} 34.311^{s}$; DEC(J2000) = $-01\degr 12\arcmin 45.90\arcsec$, which corresponds to the position of the peak of the integrated emission of the ammonia (6,6) satellite lines. We have chosen the position of the satellite lines of the (6,6) transition because they have the lowest opacity and highest excitation among the observed lines, so they can penetrate further inside the core, and may trace better the warmer gas, close to the central heating object. Our chosen central position lies roughly in between the positions of the two embedded radio continuum sources detected by \citet{cesa10}, which are separated $\sim 0\farcs2$. The systemic velocity of the cloud is not accurately determined. In the literature we find velocities ranging from 96.24 km s$^{-1}$ (\citealt{bel05}; from ground state CH$_3$CN (12$-$11) observations) to 98.8 km s$^{-1}$ (\citealt{cesa94}, from low resolution NH$_3$(4,4) observations). We have adopted the value of V$_{LSR}$=97.4 km s$^{-1}$ from the interferometric data of \citet{bel05}, obtained by fitting simultaneously several K components of the $v_8=1$ CH$_3$CN (5-4) transition. As the original purpose of the observations presented here was not to search for infall signatures, they were not designed adequately for this study. The spectral resolution is poor ($\sim$ 2.4 km s$^{-1}$) and the observations of the different transitions were not centred exactly at the same velocity. Thus, we cannot obtain pairs of velocity channels that are exactly symmetric in velocity with respect to the rest velocity of the cloud, which hinders an accurate comparison of red- and blue-shifted channel pairs. As our best approximation, we have compared channel pairs with the most similar velocity shifts with respect to the assumed systemic velocity of the cloud. Given the uncertainty in this central velocity and the large channel widths ($\sim2.4$ km s$^{-1}$), in order to ensure that we are comparing emission with different velocity signs (i.e., red vs. blue) we have excluded the channel pair closest ($\sim \pm 2$ km s$^{-1}$) to the systemic velocity. Fig. 3 shows the observed intensity as a function of the projected distance to the core centre for blue-shifted (blue continuous lines) and red-shifted (red dotted lines) velocity channels of the ammonia (2,2), (3,3), (4,4), (5,5) and (6,6) main lines. The main lines of all these transitions are certainly optically thick (as indicated by the main to satellite line ratio that is close to unity), and a pattern is clearly identifiable in the images of all the observed transitions. For each pair of blue/red channels, the intensity of the blue-shifted channel is stronger, and increases more sharply towards the centre, than in the corresponding red-shifted channel, where the intensity distribution is flatter. Moreover, as the velocity increases with respect to the systemic velocity, the emission becomes more compact (a narrower spatial intensity profile). This behaviour is in agreement with the expected mapping signatures of infall for angularly resolved sources described by A91 (see above, and Fig. 1), suggesting that gravitational infall motions strongly contribute to the kinematics of the G31 HMC. Since the satellite lines appear to be also optically thick (at least for the lowest excitation transitions), similar signatures would be expected in the satellite line channel maps. Unfortunately, because of blending between satellite lines due to the large line widths, a similar analysis cannot be carried out for the satellite emission. \begin{figure*} \includegraphics[width=\linewidth]{fig3.eps} \caption{Circularly-averaged observed intensity as a function of angular distance to the centre of G31.41 HMC, for different pairs of blue-shifted (solid blue line) and red-shifted (dotted red line) channel maps. Results for the ammonia (2,2), (3,3), (4,4), (5,5) and (6,6) transitions (from left to right) are shown. The sizes of the synthesized beams are given in Table 1. Labels indicate the channel velocity (in km~s$^{-1}$) relative to the assumed systemic velocity of the cloud ($V_{\rm LSR}=97.4$ km~s$^{-1}$).} \label{fig3} \end{figure*} To confirm the presence of these infall signatures more quantitatively, we compare our observational results with the predictions of the model for G31 HMC developed by \citet{oso09}. These authors modelled G31 HMC as a spherically symmetric envelope of dust and gas that is collapsing onto a recently formed massive star ($M_*\simeq$ 25 M$_\odot$) that is undergoing an intense accretion phase ($\dot M\simeq 3\times10^{-3}$ M$_\odot$ yr$^{-1}$). The envelope consists of an inner, infalling region ($r_{\rm ew}\simeq 2.3\times10^4$ au) that is surrounded by a static part ($R_{\rm ext}\simeq 3\times10^4$ au). The density and velocity structure of the envelope are obtained from the solution of the dynamical collapse of a singular logatropic sphere. The stellar radiation ($L_* \simeq 8\times10^4$ L$_\odot$) and the accretion luminosity ($L_{\rm acc} \simeq 1.5\times10^5$ L$_\odot$) provide the source of heating of the envelope, whose temperature as a function of the distance to the centre is self-consistently calculated from the total luminosity. The excitation and absorption coefficient of the ammonia transitions are calculated according to the physical conditions along the envelope, and the radiative transfer is performed in order to obtain the emerging ammonia spectra. The parameters of the model are determined by fitting the observed spectral energy distribution (SED) and the VLA B-configuration NH$_3$(4,4) line spectra obtained in the observations of \citet{cesa98} that angularly resolve the source. Fig. 4 shows the intensity predicted by the G31 HMC model as a function of the projected distance to the core centre, for different pairs of blue/red-shifted channels of the ammonia (2,2) to (6,6) main lines. The angular and the spectral resolutions have been set equal to those of the observations. \begin{figure*} \includegraphics[width=\linewidth]{fig4.eps} \caption{Intensity as a function of angular distance to the centre, obtained with the model of G31 HMC by \citet{oso09}, for different pairs of blue-shifted (solid blue line) and red-shifted (dotted red line) channel maps. Results for the ammonia (2,2), (3,3), (4,4), (5,5) and (6,6) transitions (from left to right) are shown. Model results have been convolved with a Gaussian with a FWHM equal to that of the observational beam (see Table 1). Labels indicate the velocity (in km~s$^{-1}$) relative to the systemic velocity of the cloud.} \label{fig4} \end{figure*} As can be seen from Fig. 4, the spatial intensity profiles predicted by the model of G31 HMC developed by \citet{oso09} show the basic signatures of spatially resolved infall proposed by A91 from a simplified, general treatment of protostellar infall. The spatial intensity profiles of the blue-shifted channels are centrally peaked, while those of the red-shifted channels are flatter; and the emission in both red- and blue-shifted channels becomes more compact as the relative velocity increases. Furthermore, by comparing Figs. 3 and 4 we note that there is a remarkable quantitative agreement between the observations of G31 HMC and the model of this source, as the shape and range of values are similar in the observed and modelled spatial intensity profiles. Thus, we conclude that the general infall signatures described by A91 are present in G31 HMC, suggesting that infall motions have a significant contribution to the kinematics of this source. The quantitative agreement with the predictions of a specific model for G31 HMC indicates that these signatures are sensitive to the physical properties (in particular, the kinematics and temperature profile) of the source and, therefore, that in principle they can be used to verify and/or improve source models. Although there is a general agreement, the precise shape of the predicted and observed intensity profiles are not fully coincident. For example, at high velocities (relative to rest) the model emission is in general weaker and more compact than observed (bottom panels in Figs. 3 and 4). Some of the discrepancies could be due to the fact that the pairs of velocity channels extracted from the observations are not exactly symmetric in velocity because of the large channel widths and the uncertainty in the central velocity. Observations with higher spectral resolution of optically thick lines, combined with additional observations of optically thin lines (to accurately determine the central velocity) are required to better study the trends presented here. Presence of rotation and outflow motions (e.g., \citealt{cesa11}) and, in general, departures from spherical symmetry, that are not considered in the modelling of Osorio et al. (2009) (e.g., the presence of a central binary), could affect the infall signatures, and account for some of the observed differences. In Section 3.2 we will examine the effects of rotation on the spatial intensity profiles. We want to stress that G31 HMC is an exceptional source, with very strong molecular ammonia emission, making it possible to identify for the very first time spatially resolved infall signatures in molecular emission. The advent of large interferometers with improved sensitivity, such as ALMA, should make it possible to extend this kind of studies to a large number of sources in the near future. \subsubsection{The central blue spot} The infall signatures in the spatial intensity profiles produce an additional signature in the first-order moment (intensity weighted mean velocity, equivalent to a centroid velocity map). This signature is easily identifiable, and consists in a compact spot of blue-shifted emission towards the peak of integrated intensity (zero-order moment) of the source. This spot of blue-shifted emission is a consequence of the differences in intensity and spatial distribution of the emission in the red- and blue-shifted channels. As explained in Section 1 and illustrated in Fig. 1, the emission near the edges of the emitting region has a similar intensity in the pairs of red- and blue-shifted velocity channels that are symmetrically located with respect to the systemic velocity of the cloud. However, while the intensity distribution remains almost flat in the red-shifted channels, it rises sharply towards the centre in the blue-shifted channels. As a consequence, the intensity-weighted mean velocity towards the central region will appear blue-shifted because of the higher weight of the strong blue-shifted emission. Additionally, the integrated intensity (zero-order moment) will peak towards the central position. As one moves away from the centre the integrated intensity decreases, the blue and red-shifted intensities become similar, and the intensity-weighted mean velocity approaches the systemic velocity of the cloud. Therefore, the first-order moment of an infalling envelope is expected to be characterized by a compact spot of blue-shifted emission towards the position of the zero-order moment peak. We will designate this infall signature the ``central blue spot''. The ``central blue spot'' will appear overlapped with other possible velocity gradients (e.g., due to rotation) that may be present in the first-order moment of the source. This is the case of G31 HMC where the first-order moment (Fig. 2) shows both a velocity gradient from NE to SW, and a spot of blue-shifted emission towards the peak of the zero-order moment. The velocity gradient suggests rotation with a PA $\simeq 150^\circ$. The clear detection of the ``central blue spot'' signature in G31 HMC indicates that infall motions play a fundamental role in the gas kinematics of this source. As can be seen in Fig. 2, the blue spot signature is present in all the observed ammonia transitions, confirming that its detection is a robust result. It is less evident in the (2,2) transition, probably because this transition has a larger contribution from outer, colder gas. We note that the opacity is expected to decrease in the central part of the blue-shifted emitting region (see Appendices A and C of Anglada et al. 1987). This can reduce the degree of asymmetry and partially erase the blue spot signature in molecular transitions of moderately high optical depth. The observations and modelling suggest that this is not the case in G31 HMC. We also note that the ``central blue spot'' infall signature is related to the ``blue bulge'' signature introduced by Walker, Narayanan, \& Boss (1994), who realized that in the centroid velocity map of a source with a rotation velocity gradient, the blue-shifted contours cross the rotational axis into the side that would otherwise correspond to red-shifted emission. The ``blue bulge'' signature is interpreted as a consequence of the overall asymmetry between blue- and red-shifted emission produced by infall motions. The ``central blue spot'' signature introduced here traces specifically the sharp increase in the degree of asymmetry between the blue- and red-shifted intensities as one moves towards the centre, which produces a central compact blue spot. The ``blue bulge'' signature is also visible in our data. Fig. 2 shows that, for all the observed transitions, part of the blue-shifted emission (the emission shown in green color in the figure) extends to the NE and penetrates into the side that would otherwise correspond to red-shifted emission. \subsection{On the effects of rotation} As can be seen in Fig. 2, the first-order moment (intensity weighted mean velocity) of the ammonia emission in G31 HMC shows a clear velocity gradient in the SW-NE direction, with average velocities preferentially blue-shifted in the SW side and preferentially red-shifted in the NE side of the source. This velocity gradient has been previously reported by other authors using several molecular transitions (e.g., Beltr\'an et al. 2004; Gibb et al. 2004; Cesaroni et al. 2011), and has been attributed either to rotation or to a bipolar outflow. In order to properly interpret the G31 HMC results, in the following we will discuss which are the expected rotation signatures in the spatial intensity profiles, and how they affect the infall signatures discussed in the previous sections. To analyse the case of an infalling rotating envelope, and to compare it with the non-rotating case, we have considered a system with properties similar to those of the example described in A91 (Fig. 1), but with the velocity field given by the TSC (Terebey, Shu, \& Cassen 1984) prescription. The TSC formalism provides a generalised solution of the problem of the spherical collapse of a singular isothermal sphere (Shu 1977) that includes the effects of an initially uniform and slow rotation. In the TSC velocity field (equations 88-90 of Terebey, Shu, \& Cassen 1984), not only radial but also azimuthal and polar velocity components are present. As a consequence, matter does not fall radially onto the central protostar, but settles into a centrifugally supported disk around it (Terebey, Shu, \& Cassen 1984; Shu, Adams, \& Lizano 1987). In the inner region of the infalling envelope (near the centrifugal radius), the magnitude of rotation and infall velocities tend to equal, and the density adopts a flattened configuration towards the equatorial plane. At larger radii (much larger than the centrifugal radius), the density law tends to the free-fall form solution with nearly radial streamlines. Fig. 5a shows the contours of equal l.o.s. velocity of the TSC velocity field close to the equatorial $z$-$p$ plane, for the example considered. The TSC isovelocity contours are closed curves with sizes similar to those of radial infall shown in Fig. 1a. However, in radial infall all the isovelocity contours are aligned and axially symmetric with respect to the line-of-sight towards the centre, while the TSC isovelocity contours are asymmetric with respect to this line-of-sight, and they appear somewhat distorted. The axis going from the centre of the core to the vertex (the point more distant from the centre) of the TSC isovelocity contours is rotated with respect to the line-of-sight. The angle of rotation of this axis increases from the outer (lower) isovelocity contours to the inner (higher) isovelocity contours where it reaches $\sim45\degr$. This is because in the TSC formalism the rotation velocity is small in the outer parts of the infalling region, that behave almost as free-fall, while in the proximity of the centrifugal radius, in the equatorial plane, it becomes comparable to the radial component. \begin{figure*} \includegraphics[width=\linewidth]{fig5.eps} \caption{{\em (a)} Contours of equal l.o.s. velocity for a collapsing and rotating protostellar envelope in a $z$-$p$ plane close to the equator, as described by the TSC formalism (Terebey, Shu, \& Cassen 1984). The (0,0) offset corresponds to the position of the central protostar. The angular momentum is assumed to be perpendicular to the $z$-$p$ plane, and pointing inside the page (clockwise rotation, as seen from the reader). Note that a given line-of-sight intersects an isovelocity surface at two points. If the emission at each velocity is optically thick, only the part of the isovelocity surface facing the observer (thick line) is observable, while the rear part (thin line) is not observable. In comparison with the radial collapse shown in Fig. 1, where all the isovelocity surfaces are symmetric with respect to the line-of-sight to the centre, the main effect of the rotation of the cloud is to rotate the axes of the isovelocity surfaces in the sense of the cloud rotation. {\em (b)} Intensity as a function of angular offset from the source centre, in the equatorial direction, for channel maps of different l.o.s. velocity. As channel maps are, in fact, images of the temperature distribution in the corresponding isovelocity surfaces, the misalignment of the axes of the isovelocity surfaces with respect to the central line-of-sight results in asymmetries in the observed spatial intensity profiles. Calculations correspond to a central protostar of $\sim 0.5$ M$_\odot$ at 160 pc observed with a beam of $0\farcs2$. A centrifugal radius of 50 au has been assumed. } \label{fig5} \end{figure*} Following the simplified analysis of A91, that assumes a high enough optical depth, the intensity map at a given l.o.s. velocity is, in fact, an image of the excitation temperature distribution in the side facing the observer of the corresponding isovelocity surface. Since the axes of the TSC isovelocity surfaces are misaligned with respect to the line-of-sight, the observer will see an asymmetric source in each velocity channel. In the example considered (Fig. 5), the emission in the red-shifted channels will be more extended in the eastern side (negative offsets) and more compact in the western side (positive offsets), while in the blue-shifted channels the emission will appear more extended in the western side and more compact in the eastern side. Thus, the spatial intensity profiles will be asymmetric as illustrated in Fig. 5b, where we show the spatial intensity profile along the equatorial direction, for different velocity channels. As in the non-rotating case, the images in blue-shifted channels present centrally peaked intensity distributions, while the red-shifted channels present flatter intensity distributions; and the emission becomes more compact with increasing l.o.s. velocity. Then, the main difference with respect to the non-rotating case is the opposite asymmetry, with respect to the central position, of the red- and blue-shifted pairs of velocity channels. It can be shown that the asymmetries inferred from the velocity field of the TSC collapse are a general property of the rotating infalling envelopes. Let us consider a given isovelocity surface in an infalling and rotating cloud (in the following descriptions, we consider velocities relative to the systemic velocity of the cloud). If rotation was absent (pure radial infall, as described in Section 1) the isovelocity surface considered, which would correspond to a sole l.o.s. infall velocity, would be symmetric with respect to any plane containing the centre of the core. However, rotation will produce l.o.s. velocity components of opposite sign at both sides of the rotation axis that, together with the l.o.s. infall component, will contribute to the total l.o.s. velocity. On the side where the l.o.s. rotation and infall velocities have the same sign (left side for the red-shifted channels or right side for the blue-shifted channels, as seen from the observer in Fig. 5a) rotation increases the magnitude of the total l.o.s. velocity. Thus, a contribution of the infall component smaller than in the case of no rotation would be required to obtain the same total l.o.s. velocity considered. Infall isovelocity surfaces are nested like matryoshka dolls, with decreasing magnitude of the l.o.s. infall velocity in the outer, larger surfaces (Fig. 1a); therefore, this side of the isovelocity surface considered will consist of points located at outer positions with respect to the case of pure radial infall. On the other hand, on the side where l.o.s. rotation and infall velocities have opposite sign, a larger contribution of the infall component will be required to compensate the decrease in the magnitude of total l.o.s. velocity produced by rotation. Therefore, on this side, the isovelocity surface considered will consist of points closer to the centre (higher magnitude of the l.o.s. infall velocity) than in the case of pure radial infall. In summary, the observed emission at a given l.o.s. velocity from a collapsing and rotating cloud will be asymmetric with respect to the rotation axis, being more extended on the side where rotation and infall velocity components have the same sign and more compact on the opposite side. Likewise, the spatial intensity profiles are stretched on the side where the channel and rotation velocities are both red-shifted or both blue-shifted while they are shrunk on the opposite side. It should also be noted that the set of spatial intensity profiles for different velocity channels is actually equivalent to a position-velocity (P-V) diagram. However, in a conventional P-V diagram the spatially resolved infall signatures are usually more difficult to identify. To study whether the rotation signatures discussed above are in qualitative agreement with what is observed in G31 HMC, we obtained the average intensity of the ammonia main line as a function of projected distance to the centre (ring-averaged values, as in Section 3.1, to increase the signal-to-noise ratio), but we calculated this intensity separately in the half of the source that rotation shifts to the blue (SW) and in the half that rotation shifts to the red (NE). From Fig. 2 we infer that the rotation axis (i.e., the angular momentum of the core) has a PA $\simeq 150\degr$. So, for every observed velocity channel, we obtained the average intensity over half-annuli of different radii corresponding to the NE ($-30\degr < {\rm PA} < 150\degr$) and SW ($150\degr < {\rm PA} < 330\degr$) halves of the source. In Fig. 6 we plotted the resulting intensity profiles for different velocity channels of the observed ammonia transitions. As can be seen in the figure, the intensity profiles show the general trends predicted for a rotating collapsing core: (i) The spatial intensity profiles of the blue-shifted channels sharply increase towards the centre, while those of the red-shifted channels are flatter; (ii) The profiles become more narrow as the magnitude of the infall velocity increases; (iii) The intensity profiles are asymmetric with respect to the central position, being the NE part (negative offsets) in the red-shifted channels more extended than the SE part (positive offsets), while the opposite is true for the blue-shifted channels. This behaviour is in agreement with what is expected for an infalling source with clockwise rotation (as seen from north). It should be noted that, so far, we only have constructed a spherically symmetric model for G31 HMC (Osorio et al. 2009); therefore, we can only evaluate the effects of rotation in a qualitative way. We are working on an integral modelling of the source, where the observational data are fitted to an infalling rotating envelope, whose results will be presented elsewhere (Mayen-Gijon et al., in preparation). \begin{figure*} \includegraphics[width=\linewidth]{fig6.eps} \caption{Observed intensity as a function of angular offset from the centre of G31.41 HMC, for different pairs of blue-shifted (solid blue line) and red-shifted (dotted red line) channel maps. Intensities have been averaged over half-annuli at both sides of the rotation axis, assumed to be at PA = 150$\degr$. Negative offsets correspond to the NE half of the source ($-30\degr < {\rm PA} < 150\degr$), where rotation velocities are red-shifted, and positive offsets correspond to the SW half of the source ($150\degr < {\rm PA} < 330\degr$), where rotation velocities are blue-shifted. Results for the ammonia (2,2), (3,3), (4,4), (5,5) and (6,6) transitions (from left to right) are shown. The sizes of the synthesized beams are given in Table 1. Colour labels indicate the channel velocity (in km~s$^{-1}$) relative to the assumed systemic velocity of the cloud ($V_{\rm LSR}=97.4$ km~s$^{-1}$).} \label{fig6} \end{figure*} Thus, we conclude that spatially resolved rotation and infall signatures can be identified in the images of the molecular ammonia emission of G31 HMC. It is important to note that rotation does not mask the mapping signatures of infall proposed by A91, but just modifies these signatures by making the spatial intensity profiles asymmetric. \section{Conclusions} We found that the high angular resolution images of the ammonia (2,2), (3,3), (4,4), (5,5), and (6,6) inversion transitions in G31 HMC present the mapping signatures of gravitational infall predicted by A91: (i) Given a pair of red/blue-shifted channels at symmetric velocities with respect to the systemic velocity of the cloud, in the blue-shifted channels the intensity as a function of projected distance to the core centre is stronger and sharply increases towards the centre, while the red-shifted channels show a flatter intensity distribution; (ii) The emission becomes more compact as the velocity increases with respect to the systemic velocity of the cloud. We introduced a third mapping signature of infall, that we call the ``central blue spot'', consisting in a spot of blue-shifted emission in the first-order moment towards the position of the zero-order moment peak. We found that the central blue-spot infall signature is also present in all the observed ammonia transitions of G31 HMC. These three signatures are robust infall signatures that, to our knowledge, have been observationally identified here for the first time. These signatures cannot be easily mimicked by other motions, and indicate that infall motions play a fundamental role in the gas kinematics of G31 HMC. There is a good quantitative agreement, both in the shape and range of values, between the spatial intensity profiles predicted by the model for G31 HMC developed by \citet{oso09} and those observed in the channel maps of the ammonia (2,2) to (6,6) inversion transitions. This is particularly remarkable, as only the SED and (4,4) transition data were used in fitting the model parameters. We studied how rotation affects the spatial intensity profiles of an infalling envelope. We concluded that the rotation signature makes the spatial intensity profiles asymmetric with respect to the central position but it does not mask the A91 infall signatures. The spatial intensity profile of the image in a given velocity channel (red- or blue-shifted) is stretched towards the side where rotation has the same sign (red- or blue-shifted, respectively), and it is shrunk on the opposite side. These rotation signatures are present in the spatial intensity profiles of the images of the ammonia transitions observed in G31 HMC. In summary, G31 HMC shows a quantitative agreement with the A91 infall signatures, it presents the ``central blue spot'' infall signature, and shows a qualitative agreement with the signatures expected in a rotating infalling envelope. G31 HMC is an exceptional source because of its strong ammonia emission that allowed us to identify these infall signatures for the first time. However, with the advent of new and improved high angular resolution facilities, it will become possible to identify these signatures in more sources. This kind of observations, combined with a detailed modelling, can be used to determine the kinematic properties of infalling protostellar envelopes. \section*{Acknowledgments} JMM-G, GA, MO, and JFG acknowledge support from MICINN (Spain) AYA2008-06189-C03-01 and AYA2011-30228-C03-01 grants (co-funded with FEDER funds). SL acknowledges support from PAPIIT-UNAM IN100412. We thank the referee, Eric Keto, for his valuable comments. We would like to thank Riccardo Cesaroni for his advise in the reduction of the ammonia (4,4) data.
\section{Historical context} The most prominent, most useful, and most-studied cardinal invariants associated with topological spaces are the weight, density character, and Souslin number. Countless papers and monographs over the decades have given estimates, in some cases even precise evaluations, of the value of these invariants for the usual Tychonoff product $X_I=\Pi_{i\in I}\,X_i$ of a set of spaces $(X_i)_{i\in I}$ in terms of the values for the initial spaces $X_i$. But in the case of $\kappa$-box topologies (defined in Section~1 below) on spaces of the form $X_I$, very little is known, and that is fragmentary and nowhere systematically assembled. In this paper we study with considerable thoroughness those three cardinal invariants for these modified box products, in each case seeking (as usual) estimates for the product in terms of the values for the initial spaces. Our methods are largely topological and set-theoretic, although as expected certain computations are made precise only when ZFC is enhanced with appropriate additional (consistent) axioms. Our work draws upon, and in some cases extends, published theorems of R.~Engelking and M.~Kar\l owicz \cite{EnKa}, W.~W. Comfort and S.~Negrepontis \cite{comfneg72}, \cite{comfneg74}, \cite{comfneg82}, F.~S.~Cater, P.~Erd{\H{o}}s and F.~Galvin \cite{ceg}, W.~W. Comfort and L.~C. Robertson \cite{comfrobii} and M.~Gitik and S.~Shelah \cite{gitikshelah}. We give details at appropriate points in the paper. We acknowledge with thanks helpful e-mail correspondence from (a)~Ist\-v\'an Juh\'asz, (b)~Stevo Todor\v{c}evi\'c, and (c) Santi Spadaro. \section{Introduction} Hypothesized topological spaces here are not subjected to standing separation properties. Special hypotheses are imposed locally, as required. $\alpha$, $\beta$, $\gamma$ and $\lambda$ are cardinals, $\kappa$ is an infinite cardinal, $\omega$ is the least infinite cardinal, and $\cc$ is the cardinality of the interval $[0,1]$. As usual, for $\alpha\geq\omega$ we write $\alpha^+:=\min\{\beta:\beta>\alpha\}$. $\eta$ and $\xi$ are ordinals. The symbols $w(X)$ and $d(X)$ denote respectively the weight and density character of the space $X$. A \emph{cellular} family in a space $X$ is a family of pairwise disjoint nonempty open subsets of $X$ and $S(X)$, the {\it Souslin number} of $X$, is the cardinal number \begin{equation}\nonumber \min\{\lambda~\mbox{: no cellular family}~\sA~\mbox{in}~X~\mbox{satisfies}~|\sA|=\lambda\}. \end{equation} We here follow many authors \cite{comfneg74}, \cite{comfneg82}, though not all \cite{E1}, \cite{juh71}, \cite{juh}, in allowing $w$, $d$ and $S$ to assume finite values. If for example $X$ is a discrete space of cardinality $17$, then $w(X)=d(X)=17$ and $S(X)=17^+=18$. For $I$ a set, we write $[I]^\lambda:=\{J\subseteq I:|J|=\lambda\}$; the notations $[I]^{<\lambda}$, $[I]^{\leq\lambda}$ are defined analogously. It is clear that if $\lambda>|I|$ then (a)~$[I]^\lambda=\emptyset$ and (b)~$[I]^{<\lambda}$ is the full power set $\sP(I)$. For a set $\{X_i:i\in I\}$ of sets we write $X_I:=\Pi_{i\in I}\,X_i$. For $A=\Pi_{i\in I}\,A_i\subseteq X_I$ the {\it restriction set of} $A$ is the set $R(A):=\{i\in I:A_i\neq X_i\}$. When each $X_i=(X_i,\sT_i)$ is a space, we use the symbol $(X_I)_\kappa$ to denote $X_I$ with the $\kappa$-{\it box topology}; this is the topology for which the set $$\sU:=\{\Pi_{i\in I}\,U_i:U_i\in\sT_i,|R(U)|<\kappa\}$$ is a base. (The $\omega$-box topology on $X_I$, then, is the usual product topology.) We refer to $\sU$ as {\it the canonical base} for $(X_I)_\kappa$, and to the elements of $\sU$ as {\it canonical open sets}. By way of caution to the reader, we note that even when $\kappa$ is regular, the intersection of fewer than $\kappa$-many sets, each open in $(X_I)_\kappa$, may fail to be open in $(X_I)_\kappa$. (Indeed each space $X_i$ embeds homeomorphically as a (closed) subspace of $(X_I)_\kappa$, so if some $X_i$ lacks that intersection property then so does $(X_I)_\kappa$.) For simplicity we denote by the symbol $\two$ the discrete space of cardinality $2$, and for cardinals $\alpha\geq2$ we denote by $D(\alpha)$ the discrete space of cardinality $\alpha$. For spaces $X$ and $Y$, the symbol $Y=_h X$ means that $Y$ and $X$ are homeomorphic; the symbol $Y\subseteq_h X$ means that $X$ contains a homeomorphic copy of the space $Y$. \begin{definition}\label{DefSLC} {\rm A cardinal $\kappa$ is a {\it strong limit cardinal} if $\lambda<\kappa\Rightarrow2^\lambda<\kappa$. } \end{definition} In \ref{beth}--\ref{expon} we cite the basic tools and facts we need from the elementary theory of cardinal arithmetic. For motivation, discussion and proofs where appropriate, see \cite[\S1]{comfneg74}, \cite[Appendix~A]{comfneg82} or \cite{jech}. The familiar {\it beth} cardinals $\beth_\xi(\alpha)$ are defined recursively as follows. \begin{definition}\label{beth} {\rm Let $\alpha\geq2$ be a cardinal. Then (a) $\beth_0(\alpha):=\alpha$; (b) $\beth_{\xi+1}(\alpha):=2^{\beth_\xi(\alpha)}$ for each ordinal $\xi$; and (c) $\beth_\xi(\alpha):=\Sigma_{\eta<\xi}\,2^{\beth_\eta(\alpha)}$ for limit ordinals $\xi>0$. } \end{definition} \begin{remarks}\label{bethprops} {\rm Let $\xi$ be a limit ordinal and let $\alpha\geq2$ and $\lambda\geq\omega$ be cardinals. Then (a) a set $S\subseteq\xi$ is cofinal in $\xi$ if and only if $\{\beth_\eta(\alpha):\eta\in S\}$ is cofinal in $\beth_\xi(\alpha)$; hence (b) $\cf(\beth_\lambda(\alpha))=\cf(\lambda)$. } \end{remarks} \begin{definition}\label{deflog} {\rm For $\alpha\geq\omega$, $\log(\alpha)$ is the cardinal number $$\log(\alpha):=\min\{\beta:2^\beta\geq\alpha\}.$$ } \end{definition} \begin{notation}\label{defweakexp} {\rm Let $\kappa\ge\omega$ and $\alpha\ge 2$. Then $\alpha^{<\kappa}:=\Sigma_{\lambda<\kappa}\,\alpha^\lambda$. } \end{notation} It is well known and easy to prove that $|[\alpha]^{\lambda}|=\alpha^\lambda$ when $\lambda\leq\alpha$, so $|[\alpha]^{<\kappa}|=\Sigma_{\lambda<\kappa}\,\alpha^\lambda$ when $\kappa\leq\alpha^+$. For ease of reference later, we build some redundancy into the statement of Theorem~\ref{expon}. \begin{theorem}\label{expon} Let $\alpha\geq2$ and $\kappa\geq\omega$. Then {\rm (a)} $\kappa\leq2^{<\kappa}\leq\alpha^{<\kappa}$; {\rm (b)} if $\kappa$ is regular then $\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$; {\rm (c)} if $\kappa$ is singular, then $(\alpha^{<\kappa})^{<\kappa}=\alpha{^\kappa}$; {\rm (d)} $((\alpha^{<\kappa})^{<\kappa})^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$. \end{theorem} \begin{remark} {\rm It is clear that the useful relation given in part (d) of Theorem~\ref{expon} is immediate from parts (b) and (c). The authors are not acquainted with other examples in mathematics of operators which, as in Theorem~\ref{expon}(d), first stabilize at the third iteration. Responding to a request from one of us (speaking in a seminar) for terminology suitable for this phenomenon, Peter Johnstone promptly proposed the expression ``sesquipotent". } \end{remark} The condition $(\alpha^{<\kappa})^{<\kappa}=\alpha^{<\kappa}$, satisfied by many pairs of cardinals $\alpha$ and $\kappa$, will play a role frequently in this paper. An alternate characterization is often useful. \begin{theorem}\label{weak<kappa} Let $\alpha\geq2$ and $\kappa\geq\omega$. Then \begin{itemize} \item[{\rm (a)}] these conditions are equivalent: \begin{itemize} \item[{\rm (i)}] $\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$; and \item[{\rm (ii)}] either $\kappa$ is regular or there is $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. \end{itemize} \item[{\rm (b)}] If the conditions in (a) fail, then $\kappa$ and $\alpha^{<\kappa}$ are singular cardinals and $\cf(\alpha^{<\kappa})=\cf(\kappa)$. \end{itemize} \end{theorem} \begin{proof} (a) ((i)~$\Rightarrow$~(ii)). If (ii) fails then $\kappa$ is a limit cardinal and for every $\nu<\kappa$ there is a cardinal $\lambda<\kappa$ such that $\alpha^\nu<\alpha^\lambda$, so also $\alpha^{<\kappa}$ is a limit cardinal. It is easily checked that $\cf(\kappa)=\cf(\alpha^{<\kappa})$, so $$(\alpha^{<\kappa})^{<\kappa}\geq(\alpha^{<\kappa})^{\cf(\kappa)} =(\alpha^{<\kappa})^{\cf(\alpha^{<\kappa}))}>\alpha^{<\kappa}.$$ ((ii)~$\Rightarrow$~(i)). If $\kappa$ is regular we have $(\alpha^{<\kappa})^{<\kappa}=\alpha^{<\kappa}$ by Theorem~\ref{expon}(b). Suppose then that $\kappa$ is singular, hence a limit cardinal, and that there is $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. Then $$\begin{array} {r@{=}l} (\alpha^{<\kappa})^{<\kappa}&(\alpha^\nu)^{<\kappa} =\Sigma_{\lambda<\kappa}\,(\alpha^\nu)^\lambda =\Sigma_{\nu<\lambda<\kappa}\,(\alpha^\nu)^\lambda\\ &\Sigma_{\nu<\lambda<\kappa}\,\alpha^\lambda =\alpha^{<\kappa}. \end{array}$$ (b) Clearly $\kappa$ is singular, $\alpha^{<\kappa}$ is limit, and $$\cf(\alpha^{<\kappa})=\cf(\kappa)<\kappa\le \alpha^{<\kappa}.$$ Hence $\alpha^{<\kappa}$ is singular. \end{proof} \begin{remarks}\label{R1.2} {\rm (a) As our title and our Abstract indicate, we are concerned here with the weight, density character, and Souslin number of (sometimes specialized) products of the form $(X_I)_\kappa$; the corresponding results are contained in Sections 2, 3 and 4, respectively. (b) As the reader knows well, the ``functions" $w$, $d$ and $S$ enjoy specific useful monotonicity properties; we have in mind these familiar phenomena: \begin{itemize} \item[({\it i})] If $X$ and $Y$ are spaces and $Y\subseteq_h X$, then $w(Y)\leq w(X)$; \item[({\it ii})] If $\sT_1$ and $\sT_2$ are topologies on a set $X$ with $\sT_1\subseteq\sT_2$, then $d(X,\sT_1)\leq d(X,\sT_2)$ and $S(X,\sT_1)\leq S(X,\sT_2)$. \end{itemize} On the other hand, both the analogue of ({\it i}) for $d$ and $S$ and of ({\it ii}) for $w$ can fail. For example, with $X=\two^\cc$ and $$Y:=\{x\in X:|\{i\in I:x_i\neq0\}|=1\},$$ one has $Y$ discrete in $X$ with $|Y|=\cc$ and $d(X)=\omega<\cc=d(Y)$, also $S(X)=\omega^+<\cc^+=S(Y)$. And with $X=\two^\cc$ and $Y'$ a countable dense subset of $X$ one has $w(Y')=w(X)=\cc$ when the usual product topology $\sT_1$ is considered, but $w(Y',\sT_2)=\omega<\cc$ when $Y'$ is given the discrete topology $\sT_2\supseteq\sT_1$. We use the indicated monotonicity properties ({\it i}) and ({\it ii}) frequently in this paper, without warning or comment. We use also the fact that if $X$ is a space and $Y$ is dense in $X$, then necessarily $S(Y)=S(X)$.} \end{remarks} \section{On the weight of $\kappa$-box products} \begin{discussion}\label{oldweight} {\rm It is well known \cite[2.3.F(a)]{E1} for each set $\{X_i:i\in I\}$ of $T_1$-spaces with $w(X_i)\geq2$ that $X_I:=\Pi_{i\in I}\,X_i$ satisfies $$w(X_I)=\max\{\sup_{i\in I}\,w(X_i),|I|\}.$$ In particular, \begin{equation}\label{Eq1} w(X_I)=|I|~\mbox{if each}~X_i~\mbox{satisfies}~w(X_i)\leq|I|. \end{equation} In Theorem~\ref{T2.9} we give the correct analogue of (\ref{Eq1}) for $\kappa$-box topologies. } \end{discussion} \begin{lemma}\label{LC} Let $\alpha\geq\omega$ and $\kappa\geq\omega$. Then $$w((\two^\alpha)_\kappa)\ge\alpha.$$ \end{lemma} \begin{proof} Let $Y\subseteq X=\two^\alpha$ be as in Remarks~\ref{R1.2}(b). Then $Y$ is discrete in $\two^\alpha$, hence is discrete in $(\two^\alpha)_\kappa$, so $$w((\two^\alpha)_\kappa)\ge w(Y)=\alpha.$$ \vskip-18pt \end{proof} \begin{theorem}\label{T1} Let $\alpha\geq\omega$ and $\kappa\geq\omega$ and let $\{X_i:i\in I\}$ be a set of spaces such that $w(X_i)\le\alpha$ for each $i\in I$. Then {\rm (a)} $\sup_{i\in I}\,w(X_i)\le w((X_I)_\kappa)\le\alpha^{<\kappa}\cdot|I|^{<\kappa}$; and {\rm (b)} if in addition $\two\subseteq_h X_i$ for each $i\in I$, then also $w((X_I)_\kappa)\geq|I|$. \end{theorem} \begin{proof} (a) Let $\sB_i$ be a base for $X_i$ with $|\sB_i|\leq\alpha$ and with $X_i\in\sB_i$, and for $\lambda<\kappa$ let $$\sB(\lambda):=\{B=\Pi_{i\in I}\,B_i:B_i\in\sB_i,|R(B)|=\lambda\}.$$ Then $|\sB(\lambda)|\leq|[I]|^\lambda\cdot\alpha^\lambda$, and since $\sB:=\bigcup_{\lambda<\kappa}\,\sB(\lambda)$ is a base for $(X_I)_\kappa$ we have $$w((X_I)_\kappa)\leq|\sB|\leq\Sigma_{\lambda<\kappa}\,|[I]^\lambda|\cdot\alpha^\lambda =|I|^{<\kappa}\cdot\alpha^{<\kappa}.$$ Since $X_i\subseteq_h (X_I)_\kappa$, we have $w((X_I)_\kappa)\ge w(X_i)$ for each $i\in I$. Hence $w((X_I)_\kappa)\ge \sup_{i\in I}w(X_i)$. (b) It follows from $\two\subseteq_h X_i$ that $\two^I\subseteq_h X$ and from Lemma \ref{LC} we have $w((X_I)_\kappa)\ge w((\two^I)_\kappa)\ge |I|.$ \end{proof} For future reference we re-state this portion of Theorem~\ref{T1}. \begin{corollary}\label{C2.4} Let $\alpha$ and $\kappa$ be infinite cardinals and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|\leq\alpha$ and $w(X_i)\leq\alpha$ for each $i\in I$. Then $w((X_I)_\kappa)\leq\alpha^{<\kappa}$. \end{corollary} \begin{discussion}\label{D2.5} {\rm If $\omega\leq\alpha<\alpha^+<\kappa$, then $w((\two^\alpha)_\kappa)=2^\alpha$, while $\alpha^{<\kappa}\geq\alpha^{(\alpha^+)}=2^{(\alpha^+)}$. In many models of set theory and for many cardinals $\alpha$ one has $2^{(\alpha^+)}>2^\alpha$, and in such cases the inequality $w((\two^\alpha)_\kappa)\leq\alpha^{<\kappa}$ of Corollary~\ref{C2.4} becomes strict. That explains why the formula $w(X_\kappa)=\alpha^{<\kappa}$ cannot be asserted without restraint in Corollary~\ref{C2.4}, even when $|I|=\alpha$. Our next goal in this section is to show that, subject only to the simple restrictions $\kappa\leq\alpha^+$ and $|I|=\alpha$, the inequality $w((X_I)_\kappa)\leq\alpha^{<\kappa}$ of Corollary~\ref{C2.4} becomes an equality (Theorem~\ref{T2.9}). } \end{discussion} \begin{lemma}\label{L14} Let $\alpha$ and $\kappa$ be infinite cardinals such that $\kappa\leq\alpha^+$. Then {\rm (a)} if $\lambda<\kappa$ and $\lambda\leq\alpha$, then $w((\two^\alpha)_\kappa)\ge 2^\lambda$; and {\rm (b)} $w((\two^\alpha)_\kappa)\geq2^{<\kappa}$. \end{lemma} \begin{proof} (a) If $\kappa=\alpha^+$ then $(\two^\alpha)_\kappa$ is the discrete space $D(2^\alpha)$, which has weight $2^\alpha=2^{<\kappa}\geq2^\lambda$. We assume therefore that $\kappa\leq\alpha$. The space $(\two^\lambda)_{\kappa}$ is then homeomorphic to a discrete subspace of $(\two^\alpha)_\kappa$, so $w((\two^\alpha)_\kappa)\geq w((\two^\lambda)_{\kappa})=|\two^\lambda|=2^\lambda$. (b) is immediate from (a). \end{proof} \begin{theorem}\label{T2.7} Let $\kappa$ and $\alpha$ be infinite cardinals. Then {\rm (a)} if $\kappa\geq\alpha^+$ then $w((\two^\alpha)_\kappa)=2^\alpha$; {\rm (b)} if $\kappa\leq\alpha^+$ then $w((\two^\alpha)_\kappa)=\alpha^{<\kappa}$. \end{theorem} \begin{proof} (a) is obvious, since $(\two^\alpha)_\kappa$ is discrete. (b) The inequality $\leq$ is given by Corollary~\ref{C2.4}. We show $\geq$. If $\kappa=\alpha^+$ then $(\two^\alpha)_\kappa$ is the discrete space $D(2^\alpha)$, which has weight $2^\alpha=\alpha^\alpha=\alpha^{<\kappa}$. We assume in what follows that $\kappa\leq\alpha$ and we consider two cases. \underline{Case 1}. $2^{<\kappa}\leq\alpha$. If $w((\two^\alpha)_\kappa)\geq\alpha^{<\kappa}$ fails then there is $\lambda<\kappa$ such that $w((\two^\alpha)_\kappa)<\alpha^{\lambda}$; we fix such $\lambda$ and we choose in $(\two^\alpha)_\kappa$ a base $\sB$ of canonical open sets such that $|\sB|<\alpha^{\lambda}$. From Lemma~\ref{L14}(b) we have $|\sB|\geq2^{<\kappa}$. For every $A\in[\alpha]^{<\kappa}$ there is $B\in\sB$ such that $A\subseteq R(B)$. (To check that, it is enough to choose in $(\two^\alpha)_\kappa$ a canonical open set $U=\Pi_{i\in I}\,U_i$ with $R(U)=A$ and $x\in U$, and then to find $B\in\sB$ such that $x\in B\subseteq U$. Then $B$ is as required.) Thus \begin{equation}\label{Eq2} [\alpha]^{<\kappa}\subseteq\bigcup\{[R(B)]^{<\kappa}:B\in\sB\}. \end{equation} For each $B\in\sB$ we have $|R(B)|<\kappa$ and hence $[R(B)]^{<\kappa}=\sP(R(B))$. Therefore $|[R(B)]^{<\kappa}|=2^{|R(B)|}\leq2^{<\kappa}$. From (\ref{Eq2}), then, we have the contradiction $$\alpha^{<\kappa}=|[\alpha]^{<\kappa}|\leq\Sigma\{[R(B)]^{<\kappa}|:B\in\sB\}\leq$$ $$2^{<\kappa}\cdot|\sB|=|\sB|<\alpha^{\lambda}\leq\alpha^{<\kappa}.$$ \underline{Case 2}. Case 1 fails. Then there is $\lambda<\kappa$ such that $2^\lambda>\alpha$. If the desired inequality $w((\two^\alpha)_\kappa)\geq\alpha^{<\kappa}$ fails then there is $\mu<\kappa$ such that $w((\two^\alpha)_\kappa)<\alpha^\mu$, and then with $\delta:=\max\{\lambda,\mu\}$ we have the contradiction $$w((\two^\alpha)_\kappa)<\alpha^\delta\leq(2^\delta)^\delta=2^\delta=w((\two^\delta)_\kappa)\leq w((\two^\alpha)_\kappa).$$ \vskip-20pt \end{proof} \begin{remark} {\rm When $\kappa=\alpha^+$ in Theorem~\ref{T2.7}, parts (a) and (b) are compatible since $\alpha^{<\kappa}=\alpha^\alpha=2^\alpha$ in that case. } \end{remark} \begin{corollary}\label{C2.8} Let $\kappa$ and $\alpha$ be infinite cardinals. Then $$w((\two^{(\alpha^{<\kappa})})_\kappa)=(\alpha^{<\kappa})^{<\kappa}= w((\two^{((\alpha^{<\kappa})^{<\kappa})})_\kappa).$$ \end{corollary} \begin{proof} From Theorem~\ref{expon}(a) we have $\kappa\leq\alpha^{<\kappa}\leq(\alpha^{<\kappa})^{<\kappa}$. The first equality then results by replacing $\alpha$ by $\alpha^{<\kappa}$ in Theorem~\ref{T2.7}(b), the second equality results by making the same substitution one more time. \end{proof} \begin{theorem}\label{T2.9} Let $\alpha$ and $\kappa$ be infinite cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|=\alpha$, and $\two\subseteq_h X_i$ and $w(X_i)\leq\alpha$ for each $i\in I$. Then {\rm (a)} if $\kappa\leq\alpha^+$ then $w((X_I)_\kappa)=\alpha^{<\kappa}$; {\rm (b)} if $\kappa\geq\alpha^+$ then $w((X_I)_\kappa)=2^\alpha$. \end{theorem} \begin{proof} We have $\two^\alpha\subseteq_h X$ and hence $(\two^\alpha)_\kappa\subseteq_h (X_I)_\kappa$. Then from Theorem~\ref{T2.7} and Corollary~\ref{C2.4} it follows that $$\alpha^{<\kappa}=w((\two^\alpha)_\kappa)\leq w((X_I)_\kappa)\leq\alpha^{<\kappa}$$ in (a), and $$2^\alpha=w((\two^\alpha)_\kappa)\leq w((X_I)_\kappa)\leq\alpha^{<\kappa}\leq\alpha^\alpha=2^\alpha$$ in (b). \end{proof} \begin{corollary}\label{wXIkappa} Let $\alpha$ and $\kappa$ be infinite cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $\two\subseteq_h X_i$ for each $i\in I$. If $w(X_i)\leq(\alpha^{<\kappa})^{<\kappa}$ for each $i\in I$ and $\alpha^{<\kappa}\leq|I|\leq(\alpha^{<\kappa})^{<\kappa}$, then $w((X_I)_\kappa)=(\alpha^{<\kappa})^{<\kappa}$. \end{corollary} \begin{proof} For $\leq$, replace $\alpha$ by $(\alpha^{<\kappa})^{<\kappa}$ in Theorem~\ref{T1}(a) and use Theorem~\ref{expon}(d). For $\geq$, it is enough to note from Corollary~\ref{C2.8} that $$w((X_I)_\kappa)\geq w((\two^I)_\kappa)\geq w((\two^{(\alpha^{<\kappa})})_\kappa)=(\alpha^{<\kappa})^{<\kappa}.$$ \vskip-20pt \end{proof} Like the authors, the reader will have noted already at this stage a distinction in kind between the pleasing, clear-cut result given in Discussion~\ref{oldweight} concerning the weight of a product in the usual product topology and the less satisfactory statement given in Corollary~\ref{wXIkappa}; in this latter, the weight of spaces of the form $(\two^I)_\kappa$ is determined by $|I|$, but unexpectedly such products which differ in size may have the same weight. \begin{corollary}\label{betaleqalpha} Let $\alpha$, $\kappa$ and $\lambda$ be infinite cardinals such that $\lambda\leq\kappa$, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|=\alpha$, and $\two\subseteq_h X_i$ and $w(X_i)\leq\alpha$ for each $i\in I$. Then $w((X_I)_\lambda)\leq w((X_I)_\kappa)$. \end{corollary} \begin{proof} Necessarily we have $\lambda\leq\kappa\leq\alpha^+$, or $\lambda\leq\alpha^+\leq\kappa$, or $\alpha^+\leq\lambda\leq\kappa$. In those three cases, Theorem~\ref{T2.9} gives respectively $w((X_I)_\lambda)=\alpha^{<\lambda}\leq\alpha^{<\kappa}=w((X_I)_\kappa)$, $w((X_I)_\lambda)=\alpha^{<\lambda}\leq\alpha^\alpha=2^\alpha=w((X_I)_\kappa)$, and $w((X_I)_\lambda)=2^\alpha=w((X_I)_\kappa)$. \end{proof} \begin{remarks}\label{R2.9} {\rm (a) Surely Corollary \ref{betaleqalpha} is as expected. Presumably a short, direct proof is available but the authors' search for that was unsuccessful. We note however that, as the simple example in Remark~\ref{R1.2}(b)({\it ii}) shows, a larger topology (for example, the discrete topology) on a given set may have a strictly smaller weight than does a smaller Tychonoff topology. (b) The authors find surprising both the extent of validity of the formula given in Theorem~\ref{T2.9} and the simplicity of its proof. We had anticipated finding an explicit formula for $w((X_I)_\kappa)$ only under special axioms and assumptions (perhaps GCH, for example), and we had anticipated the necessity to consider, at the least, such cardinals as $\cf(\kappa)$, $\cf(\alpha)$ and $\log(\alpha)$, as well as the least cardinal $\gamma$ such that $\alpha^\gamma>\alpha$. } \end{remarks} \section{On the density character of $\kappa$-box products} In this section we continue to investigate spaces of the form $(X_I)_\kappa$, focusing now on the invariant $d$ rather than on $w$. Our point of departure and motivation is the paradigmatic trilogy of Theorems~\ref{HMP}, \ref{HMPconv} and \ref{dexact}, which for the usual product topology give respectively upper bounds, lower bounds, and conditions of equality for (certain) numbers of the form $d((X_I)_\kappa)$. To avoid unnecessary restrictions, we state these three familiar results in considerable generality. Standard treatments often impose stronger separation properties according to authors' conventions, but the published proofs (of Theorems~\ref{HMPconv} and \ref{dexact} in \cite[3.19 and 3.20]{comfneg74}, for example) suffice to establish Theorems~\ref{HMP}--\ref{dexact} in the form we have chosen. Theorem~\ref{HMP} is, of course, the classic theorem of Hewitt, Marczewski and Pondiczery~\cite{hewitt46a}, \cite{marc47}, \cite{pondi}, stated here in two useful equivalent forms; and Theorem~\ref{HMPconv} is its converse. Our $\kappa$-box analogues to Theorems~\ref{HMP} and \ref{HMPconv} are given in \ref{maindlemma}--\ref{HMPgen} and \ref{d(Xkappa)}--\ref{optimal|I|}, respectively. The quest for the exact $\kappa$-box analogue of Theorem~\ref{dexact}---that is, the search for a specific cardinal number $\delta$ depending on the variables $|I|$, $d(X_i)$ ($i\in I$) and $\kappa$ so that $d((X_I)_\kappa)=\delta$---is elusive, perhaps unattainable. For example, answering a question from \cite{comfneg72}, \cite{comfneg74}, Cater, Erd\"os, and Galvin \cite{ceg} have shown that in some models for $\beta = \aleph_\omega$ the inequalities $$d((\two^{(\beta^+)})_{\omega^+})=\cc<\beta=\log(2^\beta)<d((\two^{(2^\beta)})_{\omega^+})$$ occur. Furthermore, it has been known for some time~\cite{ceg}, \cite{comfrobii} that consistently $d((\two^\beta)_{\omega^+})=(\log(\beta))^\omega$ for every infinite cardinal $\beta$. The question whether that equality holds in (all models of) ZFC, raised in ~\cite{ceg}, was answered in the negative by Gitik and Shelah \cite{gitikshelah}; we discuss their models in \ref{gitik-shelah}(d)--(g). The foregoing paragraph explains why we are able for $\kappa>\omega$ to offer exact computations of the form $d((X_I)_\kappa)=\delta$, in parallel with Theorem~\ref{dexact}, only for spaces $X_i$ ($i\in I$) and $\kappa$ subject to severe constraints. Our (few) contributions of this sort are given in Corollary~\ref{cortomain}(a) and Theorems~\ref{HMPkappa}---\ref{T3.17} below. \begin{theorem}\label{HMP} {\rm [Version 1]} Let $\alpha\geq\omega$, $X_I=\Pi_{i\in I}\,X_i$ with $d(X_i)\leq\alpha$ for each $i\in I$ and with $|I|\leq2^\alpha$. Then $d(X_I)\leq\alpha$. {\rm [Version 2]} Let $I$ be an infinite set and $\{X_i:i\in I\}$ a set of spaces. Then $d(X_I)\leq\max\{\sup\{d(X_i):i\in I\},\log|I|\}$. \end{theorem} \begin{theorem}\label{HMPconv} Let $\alpha\geq\omega$ and let $X_I=\Pi_{i\in I}\,X_i$ with $S(X_i)\geq3$ for each $i\in I$. If $d(X_i)>\alpha$ for some $i\in I$, or if $|I|>2^\alpha$, then $d(X_I)>\alpha$. \end{theorem} \begin{theorem}\label{dexact} If $\{X_i:i\in I\}$ is a family of spaces such that $S(X_i)\ge 3$ for each $i\in I$ and $|I|\ge\omega$, then $$d(X_I)=\max\{\sup\{d(X_i):i\in I\},\log|I|\}.$$ \end{theorem} \centerline{{\sc Part A}. {\sc Upper Bounds for } $d((X_I)_\kappa)$.} We say that a subset $A$ of a space $X$ is {\it strongly discrete} (in $X$) if there is a family $\{U(a):a\in A\}$ of pairwise disjoint open subsets of $X$ such that $a\in U(a)$ for each $a\in A$. Simple examples show that a strongly discrete set need not be closed. It is clear, however, that if $\kappa$ is fixed and every $A\in[X]^{<\kappa}$ is strongly discrete, then also every $A\in[X]^{<\kappa}$ is closed in $X$. That motivates the following terminology. \begin{definition}\label{strongdis} {\rm Let $\kappa\geq\omega$ and let $X$ be a space. Then $X$ is {\it strongly} $\kappa$-{\it discrete} if every $A\in[X]^{<\kappa}$ is strongly discrete. } \end{definition} \begin{remarks} {\rm (a) The terminology in Definition~\ref{strongdis} is not in universal usage. Note that the separation requirement applies only to sets $A\in[X]^{<\kappa}$, not to all $A\in[X]^{\leq\kappa}$. Note also that since in a strongly $\kappa$-discrete space $X$ each set $A\in[X]^{<\kappa}$ is both closed and discrete, the condition is strictly stronger than the condition that each discrete set $A\in[X]^{<\kappa}$ is strongly discrete. (b) We note that a space which for some $\kappa\geq\omega$ is strongly $\kappa$-discrete is a Hausdorff space. (c) We give the following lemma in the generality it warrants, but in fact we will use it only when each of the spaces $E_i$ is discrete. } \end{remarks} \begin{lemma}\label{L3.3} Let $\kappa\geq\omega$ and let $E=E_I=\Pi_{i\in I}\,E_i$ with each space $E_i$ strongly $\kappa$-discrete. Then $(E_I)_\kappa$ is strongly $\kappa$-discrete. \end{lemma} \begin{proof} Given $A\in[E]^{<\kappa}$ there is $J\in[I]^{<\kappa}$ such that the projection $\pi_J:E\twoheadrightarrow\Pi_{i\in J}\,E_i$, when restricted to $A$, is an injection. (If $\kappa>|I|$ we may take $J=I\in[I]^{<\kappa}$.) Now for $i\in I$ and $a\in A$ we choose a neighborhood $U_i(a)$ of $a_i$ in $X_i$ such that (a) $U_i(a_i)=U_i(b_i)$ if $a,b\in A$ and $a_i=b_i$, and (b) $U_i(a_i)\cap U_i(b_i)=\emptyset$ if $a,b\in A$ and $a_i\neq b_i$. \noindent [Such a family $\{U_i(a_i):a\in A\}$ exists in $X_i$ since $\pi_i[A]\in[X_i]^{<\kappa}$.] Then the sets $U(a):=(\Pi_{i\in J}\,U_i(a_i))\times\Pi_{i\in I\backslash J}\,E_i$ are open in $(E_I)_\kappa$ and are pairwise disjoint with $a\in U(a)$ for each $a\in A$. \end{proof} The principal result of this section is given in Theorem~ \ref{maindtheorem}. The following lemma does most of the work. \begin{lemma}\label{maindlemma} Let $\alpha\ge 2$, $\beta\ge\omega$, and $\kappa\ge\omega$ be cardinals and let $E:=(D(\alpha))^{2^{\beta}}$. Then $d(E_\kappa)\leq\alpha^{<\kappa}\cdot(\beta^{<\kappa})^{<\kappa}$. \end{lemma} \begin{proof} We set $X=\two^{\beta}$. Since $w(\two)=2\leq\beta$ there is (by Corollary~\ref{C2.4}) a base $\sB$ for $(\two^{\beta})_\kappa$ such that $|\sB|=\beta^{<\kappa}$. We assume without loss of generality that the elements of $\sB$ are drawn from the canonical base for $(\two^{\beta})_\kappa$ (see~\cite[1.1.15]{E1}). Let $\CC:=\{\sC\subseteq\sP(\sB):\sC~\mbox{is cellular in}~(\two^{\beta})_\kappa~\mbox{and}~~|\sC|<\kappa\},$ and for each $\sC\in\CC$ and $f:\sC\rightarrow D(\alpha)$ define $p(\sC,f)\in E=(D(\alpha))^{\two^{\beta}}$ by $$(p(\sC,f))_x= \left\{ \begin{array} {r@{~\mathrm{if}~}l} f(x)&\mbox{there is~} C\in\sC \mbox{~such that~} x\in C\\ 0&x\in\two^{\beta}\backslash\bigcup\sC \end{array} \right\}.$$ We set $A:=\{p(\sC,f):\sC\in\CC,f:\sC\rightarrow D(\alpha)\}$. Since $\kappa\leq\beta^{<\kappa}=|\sB|$ we have $|[\sB]^{<\kappa}|=(\beta^{<\kappa})^{<\kappa}$. Then since $|\CC|\leq|[\sB]^{<\kappa}|=(\beta^{<\kappa})^{<\kappa}$ and for $\sC\in\CC$ we have $|\alpha^\sC|=\alpha^{|\sC|}\leq\alpha^{<\kappa}$-many functions $f:\sC\rightarrow D(\alpha)$, it follows that $$|A|\leq\alpha^{<\kappa}\cdot(\beta^{<\kappa})^{<\kappa}.$$ It suffices then to show that $A$ is dense in $E_\kappa=((D(\alpha))^{2^{\beta}})_\kappa$. Let $U=\Pi_{x\in\two^{\beta}}\,U_x$ be a canonical open subset of $E_\kappa$. Without loss of generality we take $|U_x|=1$ when $x\in R(U)\in[\two^{\beta}]^{<\kappa}$ (and necessarily $U_x=D(\alpha)$ when $x\in\two^{\beta}\backslash R(U)$). We define $f:R(U)\rightarrow D(\alpha)$ so that $U_x=\{f(x)\}$. Since $(\two^{\beta})_\kappa$ is strongly $\kappa$-discrete (by Lemma~\ref{L3.3}) and $R(U)\in[\two^{\beta}]^{<\kappa}$, there is a family $\sC=\{C(x):x\in R(U)\}\in\CC$ of pairwise disjoint open subsets of $(\two^{\beta})_\kappa$ such that $x\in C(x)$ for each $x\in R(U)$. Then for $x\in R(U)$ we have $x\in C(x)\in\sC\in\CC$, and $(p(\sC,f))_x=f(x)\in U_x$; it follows that $p(\sC,f)\in A\cap U$, as required. \end{proof} The proof of Lemma~\ref{maindlemma} seemed so natural that for some time we considered its statement to be optimal. However, a stronger statement is available. This is the principal result of this section, given now in two equivalent formulations. \begin{theorem}\label{maindtheorem} Let $\alpha\ge 2$, $\beta\ge\omega$, and $\kappa\ge\omega$ be cardinals. {\rm (a)} Let $E:=(D(\alpha))^{2^{\beta}}$. Then $d(E_\kappa)\leq(\alpha\cdot\beta)^{<\kappa}$. {\rm (b)} Let $E:=(D(\alpha))^\beta$. Then $d(E_\kappa)\leq(\alpha\cdot\log(\beta))^{<\kappa}$. \end{theorem} \begin{proof} \,[When (a) is known, (b) follows upon replacing $2^\beta$ in (a) by $\beta$ and using the inequality $\beta\leq2^{\log(\beta)}$. To derive (a) from (b), replace $\beta$ in (b) by $2^\beta$ and use that $\log(2^\beta)\le\beta$.] To prove (a), we consider two cases. \underline{Case 1}. $\kappa$ is regular. Then it follows from Lemma \ref{maindlemma} and Theorem \ref{expon}(b) that $$d(E_\kappa)\leq\alpha^{<\kappa}\cdot(\beta^{<\kappa})^{<\kappa}= \alpha^{<\kappa}\cdot\beta^{<\kappa}=(\alpha\cdot\beta)^{<\kappa}.$$ \underline{Case 2}. $\kappa$ is singular (hence, a limit cardinal). [Here we use a trick taken from ~\cite[p.~308]{ceg}.] For $\lambda<\kappa$ there is, by Case~1 applied to the regular cardinal $\lambda^+$, a dense set $A(\lambda)\subseteq E_{\lambda^+}$ such that $|A(\lambda)|\leq(\alpha\cdot\beta)^{<\lambda^+}=(\alpha\cdot\beta)^\lambda$. The set $A:=\bigcup_{\lambda<\kappa}\,A(\lambda)$ is clearly dense in $E_\kappa$, so $$d(E_\kappa)\leq|A|\leq\Sigma_{\lambda<\kappa}\,(\alpha\cdot\beta)^\lambda=(\alpha\cdot\beta)^{<\kappa}.$$ \vskip-18pt \end{proof} \begin{corollary}\label{cortomain} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals, and let $1\leq\lambda\leq(\alpha^{<\kappa})^{<\kappa}$ and $1\le\mu\leq2^{((\alpha^{<\kappa})^{<\kappa})}$. Then {\rm (a)} $d(((D((\alpha^{<\kappa})^{<\kappa}))^{2^{((\alpha^{<\kappa})^{<\kappa})}})_\kappa)=(\alpha^{<\kappa})^{<\kappa}$; and {\rm (b)} $d(((D(\lambda))^\mu)_\kappa)\leq(\alpha^{<\kappa})^{<\kappa}$. \end{corollary} \begin{proof} Clearly (b) is immediate from (a). To prove (a), it is enough to replace $\alpha$ and $\beta$ in Theorem~\ref{maindtheorem} by $(\alpha^{<\kappa})^{<\kappa}$ and then to use Theorem~\ref{expon}(d). \end{proof} \begin{discussion}\label{D3.10} {\rm A convenient method of proof of the Hewitt-Marczewski-Pondiczery theorem (Theorem~\ref{HMP}), adopted by many expositors, is to prove first that the tractable space $E:=(D(\alpha))^{2^\alpha}$ has a dense subset $A$ with $|A|=\alpha$; since evidently there is a continuous function $f$ from $E$ onto a dense subset of $X$, the set $f[A]$ is dense in $X$, with $|f[A]|\leq|A|=\alpha$. The identical argument suffices to derive Corollary~\ref{HMPgen} from Theorem~\ref{maindtheorem} and Corollary~\ref{cortomain}. } \end{discussion} \begin{corollary}\label{HMPgen} Let $\alpha\ge 2$, $\beta\geq\omega$ and $\kappa\ge\omega$ be cardinals and let $\{X_i:i\in I\}$ be a set of spaces. \begin{itemize} \item[(a)] If $d(X_i)\leq\alpha$ for each $i\in I$ and $|I|\leq2^\beta$, then $d((X_I)_\kappa)\leq(\alpha\cdot\beta)^{<\kappa}$; \item[(b)] if $d(X_i)\leq\alpha$ for each $i\in I$ and $|I|\leq\beta$, then $d((X_I)_\kappa)\leq(\alpha\cdot\log(\beta))^{<\kappa}$; and \item[(c)] if $d(X_i)\leq(\alpha^{<\kappa})^{<\kappa}$ for each $i\in I$ and $|I|\leq2^{((\alpha^{<\kappa})^{<\kappa})}$, then $d((X_I)_\kappa)\leq(\alpha^{<\kappa})^{<\kappa}$. \end{itemize} \end{corollary} \begin{remark} {\rm For cardinals $\alpha$ and $\kappa$ such that $\alpha^{<\kappa}=\alpha=\beta$, Corollary \ref{HMPgen}(a) is \cite[3.18]{comfneg74} and was also mentioned in \cite[p.~76]{comfneg82}. } \end{remark} We restate Corollary~\ref{HMPgen}(b) in the form most easily comparable with Version~2 of Theorem~\ref{HMP}. \begin{theorem}\label{moreHMP} Let $\{X_i:i\in I\}$ be a set of spaces with each $d(X_i)=\alpha_i$, and let $\alpha:=\sup_{i\in I}\,\alpha_i$. Then $$d((X_I)_\kappa)\leq\max\{\alpha^{<\kappa},(\log(|I|))^{<\kappa}\}.$$ \end{theorem} As we see in Discussion~\ref{gitik-shelah}(d), however, the inequality in Theorem~\ref{moreHMP} can be strict. Thus consistently the obvious $\kappa$-box analogue of Theorem~\ref{dexact} can fail. \begin{discussion}\label{gitik-shelah} {\rm The two results $$d(((D(\alpha))^{2^\alpha})_\kappa)\leq\alpha^{<\kappa}$$ and $$d(((D((\alpha^{<\kappa})^{<\kappa}))^{2^{((\alpha^{<\kappa})^{<\kappa})}})_\kappa)=(\alpha^{<\kappa})^{<\kappa},$$ valid for $\alpha\geq2$ and $\kappa\geq\omega$ and given by Theorem~\ref{maindtheorem}(a) and Corollary~\ref{cortomain}(a) respectively, suggest the attractive ``intermediate" speculation \begin{equation}\label{Eq3} d(((D(\alpha^{<\kappa}))^{2^{(\alpha^{<\kappa})}})_\kappa)\leq\alpha^{<\kappa} \end{equation} which, if valid, would yield these two weaker statements: \begin{equation}\label{Eq4} d(((D(\alpha))^{2^{(\alpha^{<\kappa})}})_\kappa)\leq\alpha^{<\kappa} \end{equation} and \begin{equation}\label{Eq5} d(((D(\alpha^{<\kappa}))^{2^\alpha})_\kappa)\leq\alpha^{<\kappa}. \end{equation} We discuss what we do and do not know about the truth value of (\ref{Eq3}), (\ref{Eq4}), and (\ref{Eq5}). (a) (\ref{Eq3}) (\emph{hence also} (\ref{Eq4}) \emph{and} (\ref{Eq5})) \emph{holds for all $\alpha$ and $\kappa$ satisfying $(\alpha^{<\kappa})^{<\kappa}=\alpha^{<\kappa}$}. This is obvious from Corollary~\ref{cortomain}(a). Thus by Theorem~\ref{weak<kappa}(a) the conditions (\ref{Eq3}), (\ref{Eq4}) and (\ref{Eq5}) hold (for all $\alpha\geq2$) when $\kappa$ is regular or there is $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. (b) (\ref{Eq5}) \emph{holds in} ZFC, \emph{for all} $\kappa\geq\omega$, \emph{when} $2\leq\alpha<\omega$. This is obvious, since $|D(\alpha^{<\kappa})^{2^\alpha}|=\alpha^{<\kappa}$ in that case. (c) \emph{for all $\alpha\geq\kappa$,} (\ref{Eq5}) \emph{fails} (\emph{hence} (\ref{Eq3}) \emph{fails}) \emph{in} ZFC \emph{for certain $\kappa$.} In fact, we prove this statement: {\it Let $\alpha\geq\omega$. There are arbitrarily large cardinals $\kappa$ such that the space $E:=(D(\alpha^{<\kappa}))^\alpha$ satisfies $d(E_\kappa)>\alpha^{<\kappa}$.} To prove that, choose $\lambda\geq\omega$ such that $\cf(\lambda)\leq\alpha$ (for example, set $\lambda:=\omega$). Then, set $\kappa:=\beth_\lambda(\alpha)$. Since $\kappa>\alpha$ the space $E_\kappa$ is discrete, so from Remarks~\ref{bethprops}(b) we have $$d(E_\kappa)=|E|=\kappa^\alpha\geq\kappa^{\cf(\lambda)} =\kappa^{\cf(\kappa)}>\kappa=\alpha^{<\kappa}.$$ (d) \emph{Consistently,} (\ref{Eq4}) \emph{fails} (\emph{hence} (\ref{Eq3}) \emph{fails}) \emph{when $\alpha=2$ and $\kappa=\aleph_1$.} Indeed, Gitik and Shelah~\cite{gitikshelah}, answering a question left unresolved in \cite{ceg} and \cite{comfrobii}, have constructed models $\VV_1$ and $\VV_2$ of {\rm ZFC} such that $$d((\two^{\aleph_\omega})_{\aleph_1})= \left\{ \begin{array} {r@{\quad \mathrm{in}\quad}l} \aleph_{\omega+1} & \VV_1 \\ \aleph_{\omega+2} & \VV_2 \end{array} \right. ,$$ with $2^{\aleph_\omega}=\aleph_\omega^\omega=\aleph_{\omega+2}$ in each case and with ``GCH below $\aleph_\omega$", so that $2^{<\aleph_\omega}=\aleph_\omega$. Then, taking $\alpha=2$ and $\kappa=\aleph_\omega$ in Theorem~\ref{maindtheorem}(a) we have $$2^{\aleph_\omega}\ge d((\two^{(2^{(2^{<\aleph_\omega})})})_{\aleph_\omega})\ge d((\two^{\aleph_\omega})_{\aleph_1})=\aleph_{\omega+2}=2^{\aleph_\omega}>\aleph_\omega=2^{<\aleph_\omega}$$ in the Gitik-Shelah model $\VV_2$, while in $\VV_1$ we have $$2^{\aleph_\omega}\ge d((\two^{(2^{(2^{<\aleph_\omega})})})_{\aleph_\omega})\ge d((\two^{\aleph_\omega})_{\aleph_1})=\aleph_{\omega+1}>\aleph_\omega=2^{<\aleph_\omega}.$$ Thus in both $\VV_1$ and $\VV_2$ we have $$2^{\aleph_\omega}\ge d((\two^{(2^{(2^{<\aleph_\omega})})})_{\aleph_\omega})>2^{<\aleph_\omega},$$ \noindent so (\ref{Eq4}) (hence (\ref{Eq3})) fails there. (e) We interpret the cited results of Gitik and Shelah, where the density character of so simple a space as $(\two^{\aleph_\omega})_{\aleph_1}$ is not determined by the axioms of ZFC (even when $2^{\aleph_n}=\aleph_{n+1}$ for all $n<\omega$ and $2^{\aleph_\omega}=\aleph_{\omega+2}$) as indicating the difficulty, perhaps even the futility, of finding a pleasing and definitive $\kappa$-box analogue of Theorem~\ref{dexact}. (f) It is clear from the relations $$|\two^{\aleph_\omega}|\geq d((\two^{\aleph_\omega})_{\aleph_\omega})\geq d((\two^{\aleph_\omega})_{\aleph_1}) =\aleph_{\omega+2}=|\two^{\aleph_\omega}|$$ in $\VV_2$ that $d((\two^{\aleph_\omega})_{\aleph_\omega})=\aleph_{\omega+2}$ there. For the value of $d((\two^{\aleph_\omega})_{\aleph_\omega})$ in the model $\VV_1$ we have $$\aleph_{\omega+2}=|2^{\aleph_\omega}|\geq d((\two^{\aleph_\omega})_{\aleph_\omega})\geq d((\two^{\aleph_\omega})_{\aleph_1}) =\aleph_{\omega+1}$$ there, i.e., $$\aleph_{\omega+1}\leq d((\two^{\aleph_\omega})_{\aleph_\omega})\leq\aleph_{\omega+2}.$$ As it was noted in \cite[p. 236]{gitikshelah} there exist models of {\rm ZFC} such that $$d((D(\alpha)^{\aleph_\omega})_{\kappa})=\aleph_{\omega+1}$$ for every $\alpha,\kappa<\aleph_{\omega}$ and therefore in such models $d((\two^{\aleph_\omega})_{\aleph_\omega})=\aleph_{\omega+1}$. We do not know if there exist models of {\rm ZFC} such that $d((\two^{\aleph_\omega})_{\aleph_1})=\aleph_{\omega+1}$ and $d((\two^{\aleph_\omega})_{\aleph_\omega})=\aleph_{\omega+2}$. (g) For an exact computation of the weight and Souslin number of the spaces $(\two^{\aleph_\omega})_{\aleph_1}$ and $(\two^{\aleph_\omega})_{\aleph_\omega}$ in the Gitik-Shelah models $\VV_1$ and $\VV_2$, see Remark~{\ref{S(G-S)} below. (h) While we do not pretend to follow every detail of the arguments from \cite{gitikshelah}, nor to frame maximal generalizations, we note that the consistent failure of (\ref{Eq3}) and (\ref{Eq4}) is not restricted to the case $\alpha=2<\omega$. In both $\VV_1$ and $\VV_2$ one evidently has $\aleph_n^{<\aleph_\omega}=\aleph_\omega$ for $0<n<\omega$, so (\ref{Eq3}) and (\ref{Eq4}) fail in those models (with $\kappa=\aleph_\omega$) for every $\alpha$ such that $2\leq\alpha<\aleph_\omega$. (i) In passing we note the existence of two misprints in \cite{gitikshelah} which have confused at least two readers: Reference in Theorem~1.1(c) should be only to {\it uncountable} cardinals $\gamma$, and in Theorem~4.2(4) the symbol $<\aleph_0$ should be $<\aleph_1$. }} \end{discussion} \begin{remark}\label{R3.18} {\rm The arguments developed to prove \ref{maindlemma}--\ref{moreHMP} follow the general pattern of classical arguments used to prove the original Hewitt-Mar\-czewski-Pondiczery Theorem~\ref{HMP}, albeit with combinatorial modifications necessary to accommodate to the $\kappa$-box topology. (When $\kappa=\omega$, Lemma~\ref{L3.3} reduces to the simple observations that (1)~the product of Hausdorff spaces is a Hausdorff space, and (2)~in a Hausdorff space, the points of any finite set can be separated by disjoint open sets.) Quite likely, it was reasoning similar to ours which over 40 years ago provoked from Engelking and Kar{\l}owicz \cite[p.~285]{EnKa}, after they had completed their own proof of the Hewitt-Mar\-czewski-Pondiczery theorem, the cavalier sta\-te\-ment (here we quote faithfully, but using the notation of the present paper) ``We can also derive theorems analogous to those above for $\kappa$-box topologies$\,\ldots$. We shall not formulate these theorems since they are less interesting, but the reader, if he wishes, will be able to do so without the least difficulty." OK, fair enough. We do note, however, that in the several treatments known to us of the Hewitt-Marczewski-Pondiczery theorem, we have found no mention of the cardinal number $(\alpha^{<\kappa})^{<\kappa}$ which figures prominently and naturally in our development. (This is hardly surprising with respect to the paper \cite{EnKa}, since those authors restrict attention to box products of the form $(X_I)_{\kappa^+}$.) Nor have we found an indication, as in Theorem~\ref{HMPkappa} below, that the upper bound $\alpha^{<\kappa}\geq d(((D(\alpha))^{\beta})_\kappa)$ given in Theorem~\ref{maindtheorem}(a) is in fact assumed in every case with $\kappa\leq\beta\leq2^\alpha$. } \end{remark} \centerline{{\sc Part B}. {\sc Lower Bounds for $d((X_I)_\kappa)$}} In Part A, seeking $\kappa$-box analogues and generalizations of Theorem~\ref{HMP}, for specific function pairs $f$ and $g$ of two variables we have sought a function $h$ so that $$d(((D(f(\alpha,\kappa)))^{g(\alpha,\kappa)})_\kappa)\leq h(\alpha,\kappa)$$ holds. Now in Part~B, again for hand-picked $f$ and $g$, we seek $h'$ so that \begin{equation}\label{Eq6} d(((D(f(\alpha,\kappa)))^{g(\alpha,\kappa)})_\kappa)\geq h'(\alpha,\kappa). \end{equation} In some cases the choice $h=h'$ is accessible, so $$d(((D(f(\alpha,\kappa)))^{g(\alpha,\kappa)})_\kappa)$$ is computed exactly. In other cases, in parallel with Theorem~\ref{HMPconv}, we find several conditions sufficient to ensure that the inequality (\ref{Eq6}) is strict. \begin{lemma}\label{lowerbound} Let $\alpha\geq2$ and $\kappa\ge\omega$ be cardinals and let $E:=(D(\alpha))^\kappa$. Then $d(E_\kappa)\ge\alpha^{<\kappa}$. \end{lemma} \begin{proof} If the inequality fails there is $\lambda<\kappa$ such that $d(E_\kappa)<\alpha^\lambda$. The space $((D(\alpha))^\lambda)_\kappa$ is then discrete, and since the projection from $E_\kappa$ onto $((D(\alpha))^{\lambda})_{\kappa}$ is continuous we have the contradiction $$\alpha^\lambda=d(((D(\alpha))^\lambda)_\kappa)\leq d(E_\kappa)<\alpha^\lambda.$$ \vskip-18pt \end{proof} As we noted in Discussion~\ref{D3.10}, the Hewitt-Marczewski-Pondiczery theorem may be regarded as a routine generalization of this startling special case: $d((D(\alpha))^\beta)=\alpha$ when $\alpha\geq\omega$ and $1\le\beta\le 2^\alpha$. We draw specific attention therefore to the correct $\kappa$-box analogue of that result. We note that no regularity hypothesis is imposed here on the cardinal number $\kappa$. \begin{theorem}\label{HMPkappa} Let $\omega\leq\kappa\leq \beta\le 2^\alpha$. Then $d((D(\alpha))^\beta)_\kappa)=\alpha^{<\kappa}$. \end{theorem} \begin{proof} The inequalities $\geq$ and $\leq$ are immediate from Lemma \ref{lowerbound} and Theorem \ref{maindtheorem}(a), respectively. \end{proof} In the following theorems we compute the density character of certain specific spaces. \begin{theorem}\label{regularSLC} Let $\omega\leq\kappa$ and $2\leq\alpha\leq\kappa$. If either $\log(\kappa)<\kappa$ or $\kappa$ is a regular strong limit cardinal, then {\rm (a)} $2^{<\kappa}=\alpha^{<\kappa}=\kappa^{<\kappa}$; and {\rm (b)} $d(((D(\alpha))^\kappa)_\kappa)= 2^{<\kappa}=\alpha^{<\kappa}=\kappa^{<\kappa}$. \end{theorem} \begin{proof} (a) That $2^{<\kappa}\leq\alpha^{<\kappa}\leq\kappa^{<\kappa}$ is clear, since $2\leq\alpha\leq\kappa$. Now if $\log(\kappa)<\kappa$ then $$\kappa^{<\kappa}\leq(2^{\log(\kappa)})^{<\kappa} =\Sigma_{\lambda<\kappa}\,(2^{\log(\kappa)})^\lambda=\Sigma_{\lambda<\kappa}\,2^\lambda=2^{<\kappa};$$ and if $\kappa$ is regular then since no set in $[\kappa]^{<\kappa}$ is cofinal in $\kappa$ we have $[\kappa]^{<\kappa}\subseteq\bigcup_{\eta<\kappa}\,\sP(\eta)$, so if in addition $\kappa$ is a strong limit cardinal then $$\kappa^{<\kappa}=|[\kappa]^{<\kappa}|\leq\Sigma_{\eta<\kappa}\,|\sP(\eta)| =\Sigma_{\eta<\kappa}\,2^{|\eta|}\leq\Sigma_{\eta<\kappa}\,\kappa =\kappa\leq2^{<\kappa}.$$ (b) From Theorem~\ref{HMPkappa} (with $\alpha=\beta=\kappa$ there) and Lemma~\ref{lowerbound} (with $\alpha=2$ there) we have $$\kappa^{<\kappa}=d(((D(\kappa))^\kappa)_\kappa)\geq d(((D(\alpha))^\kappa)_\kappa)\geq d((\two^\kappa)_\kappa)\geq2^{<\kappa},$$ so the asserted equations follow from (a). \end{proof} Here is our most comprehensive result for numbers of the form $d((X_I)_\kappa)$. \begin{theorem}\label{upper-lower} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals, and let $\alpha\leq\lambda\leq(\alpha^{<\kappa})^{<\kappa}$ and $\kappa\le\mu\leq2^{((\alpha^{<\kappa})^{<\kappa})}$. Then {\rm (a)} $\alpha^{<\kappa} \leq d(((D(\lambda))^\mu)_\kappa)\leq(\alpha^{<\kappa})^{<\kappa}$; {\rm (b)} if $\kappa$ is regular or some $\nu<\kappa$ satisfies $\alpha^\nu=\alpha^{<\kappa}$, then $d(((D(\lambda))^\mu)_\kappa)=\alpha^{<\kappa}$. \end{theorem} \begin{proof} (a) This is clear from Lemma~\ref{lowerbound} and Corollary~\ref{cortomain}(b). (b) From Theorem~\ref{weak<kappa}(a) we have $\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$, so (b) follows from (a). \end{proof} We note next that for $\lambda=\alpha$ and $\kappa\leq\mu\leq2^{(\alpha^{<\kappa})}$, the conclusion of Theorem~\ref{upper-lower}(b) can be established with a supplementary hypothesis weaker than the existence of $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. (The ZFC-consistent existence of instances to which Theorem~\ref{T3.17} applies, while Theorem~\ref{upper-lower}(b) does not, is shown in Remark~\ref{R3.20}.) \begin{theorem}\label{T3.17} Let $\alpha\geq2$, $\kappa\geq\omega$ and $\kappa\leq\mu\leq2^{(\alpha^{<\kappa})}$, and set $E:=(D(\alpha))^\mu$. If there is $\nu<\kappa$ such that $2^{(\alpha^\nu)}=2^{(\alpha^{<\kappa})}$, then $d(E_\kappa)=\alpha^{<\kappa}$. \end{theorem} \begin{proof} That $d(E_\kappa)\geq\alpha^{<\kappa}$ is immediate from Lemma~\ref{lowerbound}. Now for $\nu\leq\lambda<\kappa$ we have $2^{(\alpha^\lambda)}=2^{(\alpha^{<\kappa})}$ and there is, by Theorem~\ref{maindtheorem}(a) with $\alpha$, $\alpha^\lambda$, and $\lambda^+$ in the role of $\alpha$, $\beta$, and $\kappa$ there, a dense set $A(\lambda)\subseteq ((D(\alpha))^{2^{(\alpha^{\lambda})}})_{\lambda^+}$ such that $|A(\lambda)|\leq(\alpha\cdot\alpha^\lambda)^\lambda=\alpha^{\lambda}$. The set $A:=\bigcup_{\nu\le\lambda<\kappa}\,A(\lambda)$ is clearly dense in $(((D(\alpha))^{(2^{(\alpha^{<\kappa})})})_\kappa$, so $d(E_\kappa)\leq d((((D(\alpha))^{(2^{(\alpha^{<\kappa})})})_\kappa)\leq |A|\leq\Sigma_{\nu\le\lambda<\kappa}\,\alpha^\lambda\le \kappa\cdot\alpha^{<\kappa}=\alpha^{<\kappa}.$ \end{proof} \begin{remarks}\label{R3.20} {\rm (a) We indicate that there are models $\MM$ of ZFC in which, for suitably chosen $\alpha$ and $\kappa$ as in Theorem~\ref{T3.17} (specifically for $\alpha=2$, $\kappa=\aleph_\omega$) there exist $\nu<\kappa$ such that $2^{\alpha^\nu}=2^{(\alpha^{<\kappa})}$ but there is no $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. To that end, using the fundamental consistency theorem of Easton~\cite{easton70} as exposed by Kunen~\cite[VIII]{kunen}, let $\MM$ be a model of ZFC in which \begin{itemize} \item[(1)]$2^{\aleph_n}=\aleph_{\omega+n+1}$ for $n<\omega$, \item[(2)]$2^{\aleph_\omega}=\aleph_{\omega+\omega+1}$, and \item[(3)]$2^{\aleph_{\omega+n+1}}=2^{\aleph_{\omega+\omega}}=\aleph_{\omega+\omega+2}$ for $n<\omega$. \end{itemize} It is clear in $\MM$, taking $\alpha=2$ and $\kappa=\aleph_\omega$, that $$\alpha^{<\kappa}=2^{<\aleph_\omega}=\aleph_{\omega+\omega},$$ so for every $\nu=\aleph_n<\aleph_\omega=\kappa$ we have $$\alpha^\nu=2^{\aleph_n}=\aleph_{\omega+n+1}< \aleph_{\omega+\omega}=\alpha^{<\kappa}$$ and $$2^{\alpha^\nu}=2^{(2^{\aleph_n})}=\aleph_{\omega+\omega+2} =2^{\aleph_{\omega+\omega}}=2^{(\alpha^{<\kappa})}.$$ (b) We note in passing that the existence of $\nu<\kappa$ such that $2^{\alpha^\nu}=2^{(\alpha^{<\kappa})}$ holds in all models in which $\log(2^{(\alpha^{<\kappa})})<\alpha^{<\kappa}$. Indeed if $\log(2^{(\alpha^{<\kappa})})=\beta<\alpha^{<\kappa}$ then there is $\lambda<\kappa$ such that $\beta<\alpha^\lambda$, and then $2^\beta=2^{\alpha^\nu}=2^{(\alpha^{<\kappa})}$ for all $\nu$ satisfying $\lambda\leq\nu<\kappa$. } \end{remarks} Next as promised we give a couple of generalizations of Theorem~\ref{HMPconv} to the $\kappa$-box context. \begin{theorem}\label{d(Xkappa)} Let $\alpha\ge \omega$, $\beta\geq2$, $\kappa\ge \omega$ and let $S(X_i)>\beta$ for each $i\in I$. Suppose that either {\rm (i)} some $i\in I$ satisfies $d(X_i)>\alpha$; or {\rm (ii)} $|[I]^{<\kappa}|>2^\alpha$; or {\rm (iii)} $\beta^{<\kappa}>2^\alpha$; or {\rm (iv)} there is $J\in[I]^{<\kappa}$ such that $\beta^{|J|}>\alpha$. \noindent Then $d((X_I)_\kappa)>\alpha$. \end{theorem} \begin{proof} The sufficiency of (i) is clear: the natural projection from $(X_I)_\kappa$ to $X_i$ is continuous and surjective, so $d((X_I)_\kappa)\geq d(X_i)$. For the rest of the proof, for $i\in I$ let $\{U_i(\eta):\eta<\beta\}$ be a cellular family in $X_i$. We prove $d((X_I)_\kappa)>\alpha$, assuming that either (ii) or (iii) holds. For $A\in[I]^{<\kappa}$ and $f\in \beta^A$ set $$U(A,f):=\{x\in X_I:i\in A\Rightarrow x_i\in U_i(f(i))\}.$$ Let $T$ be dense in $(X_I)_\kappa$ with $|T|=d((X_I)_\kappa)$ and set $T(A,f):=T\cap U(A,f)$. We claim that the map $\phi:\bigcup_{A\in[I]^{<\kappa}}\,(A\times\beta^A)\longrightarrow\sP(T)$ given by $\phi(A,f)=T(A,f)$ is injective. Let $(A,f)\neq(B,g)$ with $A,B\in[I]^{<\kappa}$, $f\in\beta^A$, and $g\in\beta^B$. We consider two cases. \underline{Case 1}. $A=B$. Then there is $i\in A=B$ such that $f(i)\neq g(i)$, so $T(A,f)\cap T(B,g)=\emptyset$ (since $U_i(f(i))\cap U_i(g(i))=\emptyset$). \underline{Case 2}. $A\neq B$. Without loss of generality there is then $i\in A\backslash B$. Choose $\eta<\beta$ such that $f(i)\neq\eta$. The set $V:=U(B,g)\cap\pi_i^{-1}(U_i(\eta))$ is then nonempty and open in $(X_I)_\kappa$, and with $p\in T\cap V$ we have $p\in T(B,g)\backslash T(A,f)$. The claim is proved. For $\lambda<\kappa$ and $A\in[I]^\lambda$ we have $A\times\beta^A\subseteq \dom(\phi)$, so $|\dom(\phi)|\geq|[I]^\lambda|$ and $|\dom(\phi)|\geq\beta^\lambda$. Then it follows that $|\dom(\phi)|\geq|[I]^{<\kappa}|\cdot\beta^{<\kappa}$, so if $d((X_I)_\kappa)=|T|\leq\alpha$ and (ii) or (iii) holds we would have the contradiction $$2^\alpha<|[I]^{<\kappa}|\cdot\beta^{<\kappa}\leq|\dom(\phi)| \leq|\sP(T)|\leq2^\alpha.$$ It remains to derive $d((X_I)_\kappa)>\alpha$ from (iv). Let $J$ be as hypothesized, and for $f\in\beta^J$ set $$V(f):=(\Pi_{i\in J}\,U_i(f(i)))\times(\Pi_{i\in I\backslash J}\,X_i).$$ Then $\sV:=\{V(f):f\in\beta^J\}$ is cellular in $(X_I)_\kappa$, so $$d((X_I)_\kappa)\geq|\sV|=|\beta^J|=\beta^{|J|}>\beta.$$ \vskip-18pt \end{proof} We note that the hypothesis in Theorem~\ref{d(Xkappa)} on the family $\{X_i:I\in I\}$ can be relaxed in places. In connection with (iv), for example, it is clear that the condition $S(X_i)>\beta$ need hold only for $i$ in some set $J\in[I]^{<\kappa}$ such that $\beta^{|J|}>\alpha$. Taking $\beta=2$ in Theorem~\ref{d(Xkappa)} and replacing $\alpha$ there first by $\alpha^{<\kappa}$ and then by $(\alpha^{<\kappa})^{<\kappa}$, we obtain respectively parts (a) and (b) of the following corollary. \begin{corollary}\label{optimal|I|} Let $\alpha\ge 2$ and $\kappa\ge\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $S(X_i)\geq3$ for each $i\in I$. {\rm(a)} If $|[I]^{<\kappa}|>2^{(\alpha^{<\kappa})}$ then $d((X_I)_\kappa))>\alpha^{<\kappa}$; and {\rm(b)} if $|[I]^{<\kappa}|> 2^{((\alpha^{<\kappa})^{<\kappa})}$ then $d((X_I)_\kappa)>(\alpha^{<\kappa})^{<\kappa}$. \end{corollary} Corollary~\ref{optimal|I|} shows that the inequalities given in Corollary~\ref{HMPgen} are sharp. The following simple combinatorial result offers reformulations of some of the hypotheses of Corollary~\ref{optimal|I|}. \begin{theorem}\label{L3.25} Let $\alpha\ge 2$ and $\kappa\ge\omega$ be cardinals and let $I$ be a set. {\rm (a)} These three conditions are equivalent: {\rm (1)} $|I|>2^{(\alpha^{<\kappa})}$; {\rm (2)} $|[I]^{<\kappa}|>2^{(\alpha^{<\kappa})}$; {\rm (3)} $|I|^{<\kappa}>2^{(\alpha^{<\kappa})}$. {\rm(b)} These three conditions are equivalent. {\rm (1)} $|I|>2^{((\alpha^{<\kappa})^{<\kappa})}$; {\rm (2)} $|[I]^{<\kappa}|>2^{((\alpha^{<\kappa})^{<\kappa})}$; {\rm (3)} $|I|^{<\kappa}>2^{((\alpha^{<\kappa})^{<\kappa})}$. \end{theorem} \begin{proof} The implications (1)~$\Rightarrow$~(2) and (2)~$\Rightarrow$~(3) are clear in both (a) and (b). To see that (3)~$\Rightarrow$~(1) in (a), note that if $|I|\leq2^{(\alpha^{<\kappa})}$ then $$|I|^{<\kappa}\leq(2^{(\alpha^{<\kappa})})^{<\kappa} \leq(2^{(\alpha^{<\kappa})})^{\kappa}= 2^{(\alpha^{<\kappa})\cdot\kappa}=2^{(\alpha^{<\kappa})}.$$ The proof that (3)~$\Rightarrow~$(1) in (b) is similar. \end{proof} \begin{remarks} {\rm (a) The authors of \cite[3.16]{comfneg74}, improving their results from~\cite{comfneg72}, show that if $E=(D(\alpha))^{2^\alpha}$ or $E=(D(\alpha))^{\alpha^+}$, then $d(E_\kappa)=\alpha$ if and only if $\alpha=\alpha^{<\kappa}$. Clearly Theorem~\ref{upper-lower}(b) above improves that statement. Similarly, Corollary~\ref{HMPgen}(a) improves \cite[3.18]{comfneg74}, which asserts the conclusion of \ref{HMPgen} only under the assumption that $\alpha=\alpha^{<\kappa}$. (b) The investigation by Hu~\cite{hu06} of cardinals of the form $d((X_I)_\kappa)$ is from a different perspective: Rather than beginning with the set $E=\Pi_{i\in I}\,D(\alpha_i)$ and seeking dense subsets of the space $E_\kappa$, Hu~\cite{hu06} uses (maximal) generalized independent families of partitions of a given set $S$ to map $S$ faithfully onto dense subsets of spaces of the form $E_\kappa$ (one writes $S\subseteq E_\kappa$). The emphasis is on finding conditions so that $S\subseteq E_\kappa$ is irresolvable. Hu~\cite{hu06} shows, for example, that if each $\alpha_i$ is less than the first cardinal which is strongly $\kappa$-inaccessible, and $E_\kappa$ contains a dense, irresolvable subspace, then $\kappa=2^{<\kappa}$, and consistently a measurable cardinal exists. } \end{remarks} \section{On the Souslin number of $\kappa$-box products} We remind the reader of our standing convention that hypothesized spaces are not assumed to enjoy any special separation properties. This complicates our exposition slightly, since it is convenient for us to cite some basic familiar results from sources where, for simplicity and often unnecessarily, such properties as Hausdorff separation are assumed throughout. We mention in particular the following two useful results, both valid for every space. These will be used frequently in what follows, without explicit restatement. {\it Let $X$ be a space. Then {\rm (a)} $S(X)\neq\omega$; and {\rm (b)} either $S(X)<\omega$ or $S(X)$ is an (infinite) regular cardinal.}\\ \noindent The proof given in~\cite[2.10]{comfneg82} of (a), although long-winded and unnecessarily complicated, is valid without separation assumptions; (b) is a fundamental result of Erd{\H{o}}s and Tarski~\cite{erdtarski43} (see \cite[2.10]{comfneg74}, \cite[2.14]{comfneg82} for other treatments). As with Sections 2 and 3 concerning weight and density character respectively, we begin this section by citing those classical Souslin-related theorems (pertaining to the usual product topology) whose $\kappa$-box analogues we study here. As usual, when a set $\{X_i:i\in I\}$ of spaces is given we write $X_I:=\Pi_{i\in I}\,X_i$. \begin{theorem}\label{firstSP} Let $\{X_i:i\in I\}$ be a set of nonempty spaces and set $$\alpha:=\sup\{S(X_F):\emptyset\neq F\in[I]^{<\omega}\}.$$ Then $$S(X_I)= \left\{ \begin{array} {r@{~\mathrm{~}~}l} \alpha&\mbox{\rm{if~(a)~$\alpha<\omega$ or (b)~$\alpha$ is regular and }~}\alpha>\omega\\ \alpha^+&\mbox{\rm{in all other cases }~} \end{array} \right\}.$$ \end{theorem} The thrust of Theorem~\ref{firstSP} is that the Souslin number of a product space $X_I$ (in the usual product topology) is completely determined by the Souslin numbers of the various subproducts $X_F$ with $F\in[I]^{<\omega}$. Much of this Section is devoted to the presentation of $\kappa$-box analogues of Theorem~\ref{firstSP}. See in particular Theorems~\ref{CN}, \ref{GCN}, \ref{T4.15}, \ref{GCH}, and Corollaries \ref{S(power)}(b), \ref{CGCH}. The proof of Theorem~\ref{firstSP} depends on nontrivial combinatorial machinery in which, reflecting the restriction to the usual product topology, the cardinal numbers $\omega$ and $\omega^+$ figure prominently. The key to the proof is the theory of quasi-disjoint families as developed by Erd{\H{o}}s and Rado~\cite{erdosrado60}, \cite{erdosrado69} (the ``$\Delta$-system lemma"); this is used in the proof of Theorem~\ref{CN}. For a thorough development of that result and of several other Souslin-related consequences, the reader may consult \cite[3.8]{comfneg74} and \cite[3.25]{comfneg82}. As we noted in Theorem~\ref{HMP}, for $\alpha\geq\omega$ the product of $2^\alpha$-many (or fewer) spaces $X_i$, with each $d(X_i)\leq\alpha$, satisfies $d(X_I)\leq\alpha$. From that and Theorem~\ref{firstSP}, one can derive this well-known theorem (see for example \cite[2.3.17]{E1}; or see Theorem \ref{T4.12} for the general $\kappa$-box statement of which Theorem~\ref{secondSP} is the case $\kappa=\omega$). \begin{theorem}\label{secondSP} Let $\alpha\geq\omega$ and $\{X_i:i\in I\}$ be a set of spaces with $d(X_i)\leq\alpha$ for each $i\in I$. Then $S(X_I)\leq\alpha^+$. \end{theorem} \begin{discussion} {\rm Theorems \ref{firstSP} and \ref{secondSP} leave unanswered even for the usual product topology a question which arises naturally in their wake: {\it Given $\alpha\geq\omega$ and a finite set $\{X_i:i\in F\}$ of spaces with each $S(X_i)\leq\alpha$, is necessarily $S(X_F)\leq\alpha$?} The brief response is that the question is not settled by the axioms of ZFC, even in the case $\alpha=\omega^+$. Referring the reader to \cite{comfneg82} for extensive comments and relevant bibliographic citations, we remark simply that it has been known in ZFC for many years that while the Souslin number may ``jump" in passing from a space $X$ to $X\times X$, roughly speaking that jump is bounded by a single exponential. To be more precise we give below a theorem taken from \cite[3.13]{comfneg74} that gives one possible generalization of that claim for the $\kappa$-box topology. For its full generalization to the $\kappa$-box context, see Theorem \ref{GKurepa}. } \end{discussion} \begin{theorem} If $\alpha\ge\omega$ and $\{X_i:i\in I\}$ is a family of spaces such that $S(X_i)\le\alpha^+$ for $i\in I$, then $S((X_I)_{\alpha^+})\le(2^\alpha)^+$. \end{theorem} The following notational device (see \cite[p.~254]{comfneg82}) is useful as we seek $\kappa$-box analogues of Theorems~\ref{firstSP} and \ref{secondSP}. \begin{notation} {\rm Let $\alpha$ and $\kappa$ be infinite cardinals. Then $\alpha$ is {\it strongly} $\kappa$-{\it inaccessi\-ble} (in symbols: $\kappa\ll\alpha$) if (a)~$\kappa<\alpha$ and (b)~$\beta^\lambda<\alpha$ whenever $\beta<\alpha$ and $\lambda<\kappa$. } \end{notation} \begin{remark}\label{help} {\rm To help the reader fix ideas, we note that the condition $\kappa\ll\alpha$ occurs for many pairs of cardinals. For example, (1) every uncountable cardinal $\alpha$ satisfies $\omega\ll\alpha$; (2) every infinite cardinal $\alpha$ satisfies $\alpha^+\ll(2^\alpha)^+$, since if $\lambda<\alpha^+$ and $\beta<(2^\alpha)^+$ then $\lambda\leq\alpha$ and $\beta\leq2^\alpha$ and hence $\beta^\lambda\leq(2^\alpha)^\alpha=2^\alpha<(2^\alpha)^+$; and (3) every pair $\kappa,\alpha$ with $\alpha\geq2$ and $\kappa\geq\omega$ satisfies $\kappa\ll((\alpha^{<\kappa})^{<\kappa})^+$, since if $\lambda<\kappa$ and $\beta<((\alpha^{<\kappa})^{<\kappa})^+$ then $\beta\leq(\alpha^{<\kappa})^{<\kappa}$ and hence $\beta^\lambda\leq((\alpha^{<\kappa})^{<\kappa})^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$ (by Theorem~\ref{expon}(d)). (4) every pair $\kappa,\alpha$ with $\alpha\geq2$ and $\kappa\geq\omega$ singular satisfies $\kappa^+\ll((\alpha^{<\kappa})^{<\kappa})^+$, since if $\lambda<\kappa^+$ and $\beta<((\alpha^{<\kappa})^{<\kappa})^+$ then $\lambda\le\kappa$ and $\beta\leq(\alpha^{<\kappa})^{<\kappa}$ and hence $$\beta^\lambda\leq((\alpha^{<\kappa})^{<\kappa})^{\kappa}=\alpha^{\kappa}=(\alpha^{<\kappa})^{<\kappa}$$ (by Theorem~\ref{expon}(c)). } \end{remark} \begin{theorem}[{cf. \cite[3.8]{comfneg74}, \cite[3.25(a)]{comfneg82}}]\label{CN} Let $\omega\le\kappa\ll\alpha$ with $\alpha$ regular and let $\{X_i:i\in I\}$ be a family of nonempty spaces. Then $S((X_I)_\kappa)\le\alpha$ if and only if $S((X_J)_\kappa)\le\alpha$ for each nonempty $J\in [I]^{<\kappa}$. \end{theorem} From the relations $\alpha^+\ll(2^\alpha)^+$ and $\kappa\ll((\alpha^{<\kappa})^{<\kappa})^+$ we have these consequences of Theorem~\ref{CN}. \begin{theorem}\label{GCN} Let $\kappa\ge \omega$ and $\alpha\ge 2$ be cardinals and let $\{X_i:i\in I\}$ be a family of nonempty spaces. Then {\rm (a)} If $\alpha\geq\omega$, then $S((X_I)_{\alpha^+})\leq(2^\alpha)^+$ if and only if $$S((X_J)_{\alpha^+})\leq(2^\alpha)^+$$ for each nonempty $J\in[I]^{\leq\alpha}$; and {\rm (b)} $S((X_I)_\kappa)\le((\alpha^{<\kappa})^{<\kappa})^+$ if and only if $$S((X_J)_\kappa)\le((\alpha^{<\kappa})^{<\kappa})^+$$ for each nonempty $J\in [I]^{<\kappa}$. \end{theorem} The following result, another immediate consequence of Theorem~\ref{CN}, furnishes in certain cases an exact formula for the numbers $S((X_I)_\kappa)$. \begin{corollary}\label{S=sup} Let $\kappa\geq\omega$, let $\{X_i:i\in I\}$ be a set of nonempty spaces, and set $\alpha:=\sup\{S((X_J)_\kappa):\emptyset\neq J\in[I]^{<\kappa}\}.$ Then {\rm (a)} {\rm (cf. \cite[3.27]{comfneg82})} If $\alpha$ is regular and $\kappa\ll\alpha$ then $S((X_I)_\kappa)=\alpha$; {\rm (b)} if $\alpha$ is singular and $\kappa\ll\alpha^+$, then $S((X_I)_\kappa)=\alpha^+$. \end{corollary} The following result, taken from \cite[3.28]{comfneg82}, is given here for the reader's convenience. Since every infinite cardinal $\alpha$ satisfies $\omega\ll\alpha^+$ with $\alpha^+$ regular, the implication (a)~$\Rightarrow$~(b) is a suitable $\kappa$-box analogue of Theorem~\ref{secondSP}. \begin{theorem}\label{iffk<<a} Let $\omega\leq\kappa<\alpha$ with $\alpha$ regular. Then these conditions are equivalent. {\rm (a)} $\kappa\ll\alpha$; {\rm (b)} if $\{X_i:i\in I\}$ is a set of spaces with each $d(X_i)<\alpha$, then $S((X_I)_\kappa)\leq\alpha$. \end{theorem} \begin{proof} (a)~$\Rightarrow$~(b). According to Theorem~\ref{CN}, it suffices to show that $S((X_J)_\kappa)\leq\alpha$ whenever $\emptyset\neq J\in[I]^{<\kappa}$. Fix such $J$, for $i\in J$ let $D_i$ be dense in $X_i$ with $|D_i|=\beta_i<\alpha$, and set $D:=\Pi_{i\in J}\,D_i$ and $\beta:=\sup_{i\in J}\,\beta_i$. Since $|J|<\kappa<\alpha=\cf(\alpha)$ we have $\beta<\alpha$, and from $\kappa\ll\alpha$ follows $|D|\leq\beta^{|J|}<\alpha$. Clearly $D$ is dense in $(X_J)_\kappa$, and from $d((X_J)_\kappa)<\alpha$ it then follows that $S((X_J)_\kappa)\leq\alpha$, as required. (b)~$\Rightarrow$~(a). Fix $\beta<\alpha$ and $\lambda<\kappa$, and set $X:=(D(\beta))^\lambda$. Then $(X)_\kappa$ is discrete, and from $S((X)_\kappa)\leq\alpha$ it follows that $\beta^\lambda=|X|=|(X)_\kappa|<\alpha$. \end{proof} \begin{corollary}\label{S(Xkappa)} Let $\alpha\geq2$ and $\kappa\geq\omega$, and let $\{X_i:i\in I\}$ be a set of spaces. {\rm (a)} If $\alpha\geq\omega$ and $d(X_i)\leq2^\alpha$ for each $i\in I$, then $S((X_I)_{\alpha^+})\leq(2^\alpha)^+$; and {\rm (b)} if $d(X_i)\leq(\alpha^{<\kappa})^{<\kappa}$ for each $i\in I$, then $S((X_I)_\kappa)\leq((\alpha^{<\kappa})^{<\kappa})^+$. \end{corollary} \begin{proof} As noted in Remark~\ref{help}(3) and \ref{help}(4) we have $\alpha^+\ll(2^\alpha)^+$ and $\kappa\ll(\alpha^{<\kappa})^{<\kappa})^+$, so Theorem~\ref{iffk<<a} applies (with $(2^\alpha)^+$ replacing $\alpha$ in (a) and with $((\alpha^{<\kappa})^{<\kappa})^+$ replacing $\alpha$ in (b)). \end{proof} The following result, which we are going to use frequently, shows a relationship between the Souslin number of product spaces of the type $(X_I)_\kappa$ and the Souslin numbers $S(X_i)$, $i\in I$. \begin{lemma}\label{obs1} Let $\alpha\geq3$ and $\kappa\ge\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $S(X_i)\geq\alpha$ for each $i\in I$. Let $\mu=\min\{\kappa,|I|^+\}$. Then $S((X_I)_{\kappa})>\beta^{<\mu}$ for each $\beta<\alpha$. \end{lemma} \begin{proof} We show first that if $\kappa\ge|I|$ then $S((X_I)_\kappa)>\kappa$. Indeed, in this case there is $J\in[I]^\kappa$, say $J=\{i_\eta:\eta<\kappa\}$. For $i_\eta\in J$ let $U(i_\eta,0)$ and $U(i_\eta,1)$ be nonempty, disjoint open subsets of $X_{i_\eta}$ and for $\eta<\kappa$ set $$U(\eta):=(\Pi_{\xi<\eta}\,U(i_\xi,0)\times (U(i_\eta,1))\times(\Pi_{i\in I\backslash \{i_\xi:\xi\le\eta\}}\,X_i).$$ Then $\sC(\kappa):=\{U(\eta):\eta<\kappa\}$ is cellular in $(X_I)_\kappa$, so $S((X_I)_\kappa)>|\sC(\kappa)|=\kappa$. Now fix $\beta<\alpha$, $\lambda<\mu$ and $J\in[I]^\lambda$. For $i\in J$ let $\{U(i,\eta):\eta<\beta\}$ be a cellular family in $X_i$, and for $f\in\beta^J$ set $U(f):=(\Pi_{i\in J}\,U(i,f(i)))\times(X_{I\backslash J})$. Then $\sC:=\{U(f):f\in\beta^J\}$ is cellular in $(X_I)_{\kappa}$, and \begin{equation}\label{EqN} S((X_I)_{\kappa})>|\sC|=|\beta^J|=\beta^\lambda. \end{equation} Since $S((X_I)_\kappa)<\beta^\lambda$ for each $\lambda<\kappa$, we have $$S((X_I)_{\kappa})\ge\beta^{<\mu} \mbox{ for each } \beta<\alpha.$$ To show that $S((X_I)_{\kappa})>\beta^{<\mu}$ for each $\beta<\alpha$ we consider three cases. \underline{Case 1}. The cardinal number $\beta^{<\mu}$ is singular. Then clearly $S((X_I)_{\kappa})>\beta^{<\mu}$. \underline{Case 2}. The cardinal number $\beta^{<\mu}$ is regular and there is $\nu<\mu$ such that $\beta^\nu = \beta^{<\mu}$. Then $S((X_I)_{\kappa})>\beta^\nu=\beta^{<\mu}$ by (\ref{EqN}). \underline{Case 3}. Cases 1 and 2 fail. Then, according to Lemma \ref{L4.13}(b), $\beta^{<\mu}=\mu$ and $\mu$ is a regular strong limit cardinal. Since $\mu=\min\{\kappa,|I|^+\}$, we have $\mu=\kappa$, hence $|I|\ge\kappa$ and since in that case $S((X_I)_{\kappa})>\kappa$ we conclude that $S((X_I)_{\kappa})>\beta^{<\mu}$. \end{proof} \begin{theorem}\label{Obsn} Let $\kappa\ge\omega$ be a limit cardinal and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|\geq\kappa$ and $S(X_i)\ge 3$ for each $i\in I$. Let also $$\alpha:=\sup\{S((X_I)_\gamma):\gamma<\kappa\} \mbox{ for }\kappa>\omega \mbox{ and }$$ $$\alpha:=\sup\{S(X_J):J\in[I]^{<\kappa}\} \mbox{ for }\kappa=\omega.$$ Then {\rm (a)} $\kappa\leq\alpha\leq S((X_I)_\kappa)\leq\alpha^+$, and $\kappa^+\leq S((X_I)_\kappa)$; {\rm (b)} if $\alpha$ is regular and $\kappa<\alpha$ then $S((X_I)_\kappa)=\alpha$; and {\rm (c)} if $\alpha$ is singular or $\kappa=\alpha$ then $S((X_I)_\kappa)=\alpha^+$. \end{theorem} \begin{proof} In each of (a), (b), or (c) the case $\kappa=\omega$ follows from Theorem \ref{firstSP}. Therefore below we consider only the case $\kappa>\omega$. (a) That $\alpha\leq S((X_I)_\kappa)$ is obvious. Let $\mu=\min\{\kappa,|I|^+\}$. Since $|I|\ge\kappa$, we have $\mu=\kappa$. Then it follows from Lemma \ref{obs1} that $S((X_I)_{\kappa})>2^{<\mu}=2^{<\kappa}\ge\kappa$, hence $S((X_I)_{\kappa})\ge\kappa^+$. Also, since $\kappa > \omega$, $S((X_I)_{\gamma^+})>2^{<\gamma^+}=2^{\gamma}\ge\gamma^+$ for every $\gamma<\kappa$, hence $\alpha\ge\kappa$. To prove $S((X_I)_\kappa)\leq\alpha^+$, suppose there is in $(X_I)_\kappa$ a basic cellular family $\sC$ such that $|\sC|=\alpha^+$, and for $\gamma<\kappa$ set $C(\gamma):=\{U\in\sC:|R(U)|<\gamma\}$. Then since $\alpha^+$ is regular with $\alpha^+>\kappa$, there is $\gamma<\kappa$ such that $|\sC(\gamma)|=\alpha^+$ and we have the contradiction $\alpha\geq S((X_I)_\gamma)\geq\alpha^{++}$. (b) A similar argument applies. If in $(X_I)_\kappa$ there is a basic cellular family $\sC$ such that $|\sC|=\alpha$ then with $$\sC(\gamma):=\{U\in\sC:|R(U)|<\gamma\} \mbox{ for } \gamma<\kappa$$ we have $\sC=\bigcup_{\gamma<\kappa}\,\sC(\gamma)$ and from the regularity of $\alpha$ and the relation $\kappa<\alpha$ we have $|\sC(\gamma)|=\alpha$ for some $\gamma<\kappa$ and then $$\alpha\geq S((X_I))_\kappa)\geq S((X_I)_\gamma)>\alpha,$$ a contradiction. (c) Since $S((X_I)_\kappa)$ is regular, this is immediate from (a). \end{proof} \begin{theorem}\label{T4.12} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals and let $\{X_i:i\in I\}$ be a set of spaces with $d(X_i)\leq\alpha$ for each $i\in I$. Then $S((X_I)_\kappa)\leq(\alpha^{<\kappa})^+$. \end{theorem} \begin{proof} We assume first that $\alpha\geq\omega$ and we consider two cases. \underline{Case 1}. $\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$. The conclusion is immediate from Corollary~\ref{S(Xkappa)} (even with the hypothesis $d(X_i)\leq\alpha$ weakened to $d(X_i)\leq\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$). \underline{Case 2}. Case 1 fails. Then $\kappa$ is singular (by Theorem~\ref{weak<kappa}(a)) and therefore a limit cardinal such that $\kappa>\omega$. If there exists $\gamma<\kappa$ such that $S((X_I)_\kappa)=S((X_I)_{\gamma^+})$ then since $\gamma^+$ is regular it follows from Corollary~\ref{S(Xkappa)}(b) that $$S((X_I)_\kappa)=S((X_I)_{\gamma^+})\le(\alpha^{\gamma})^{\gamma}\le\alpha^{<\kappa}<(\alpha^{<\kappa})^+.$$ If there is no $\gamma<\kappa$ such that $S((X_I)_\kappa)=S((X_I)_{\gamma^+})$ then for each $\gamma<\kappa$ we have $S((X_I)_{\gamma^+})\le(\alpha^{\gamma})^{\gamma}\le\alpha^{<\kappa}$ and hence $$S((X_I)_\kappa)\leq(\sup_{\gamma<\kappa}S((X_I)_\gamma))^+\le(\alpha^{<\kappa})^+$$ from Theorem~\ref{Obsn}(a). It remains to consider the case $\alpha<\omega$. Note that $d(X_i)\leq\omega$ for each $i\in I$. Then if $\kappa=\omega$ we have $$S((X_I)_\kappa)=S(X_I)\leq\omega^+=(\alpha^{<\kappa})^+$$ from Theorem~\ref{secondSP}, and if $\kappa>\omega$ then the preceding paragraphs apply to give $$S((X_I)_\kappa)\leq(\omega^{<\kappa})^+\leq((2^\omega)^{<\kappa})^+ =(2^{<\kappa})^+\leq(\alpha^{<\kappa})^+.$$ \vskip-18pt \end{proof} \begin{remark} {\rm (a) If the hypothesis $d(X_i)\leq\alpha$ of Theorem~\ref{T4.12} is weakened to $d(X_i)\leq\alpha^{<\kappa}$, the conclusion can fail. To see that, it is enough to refer to Discussion~\ref{gitik-shelah}, where we noted that for every pre-assigned $\alpha\geq\omega$ the choice $\kappa:=\beth_\lambda(\alpha)$ with $\lambda\leq\cf(\alpha)$ guarantees that the space $E:=(D(\alpha^{<\kappa}))^I$ with $|I|=\alpha$ has $E_\kappa$ discrete (since $\kappa>|I|$) and $|E_\kappa|=(\alpha^{<\kappa})^\alpha=\kappa^\alpha>\alpha^{<\kappa}$, hence $S(E_\kappa)=|E_\kappa|^+>(\alpha^{<\kappa})^+$. (b) With Theorem \ref{T4.12} in hand the implication (a)$\Rightarrow$(b) in Theorem \ref{iffk<<a} becomes now a direct corollary. Indeed, if $\alpha=\beta^+$ in Theorem \ref{iffk<<a} then $d(X_i)\le\beta$ and according to Theorem \ref{T4.12} we have $$S((X_I)_\kappa)\le(\beta^{<\kappa})^+=(\Sigma_{\lambda<\kappa}\, \beta^\lambda)^+\le(\beta\cdot\kappa)^+=\alpha$$ since $\kappa\ll\alpha$. And if $\alpha$ is a regular limit cardinal in Theorem \ref{iffk<<a} then Theorem \ref{T4.12} gives $S((X_I)_\kappa)\leq(\alpha^{<\kappa})^+$. But in this case, since $\alpha$ is regular and no set in $[\alpha]^{<\kappa}$ is cofinal in $\alpha$ we have $$\alpha^{<\kappa}=|[\alpha]^{<\kappa}|\leq\Sigma_{\eta<\alpha}\, |[\eta]^{<\kappa}|=\Sigma_{\eta<\alpha}\,|\eta|^{<\kappa}\leq\Sigma_{\eta<\alpha} \,\alpha=\alpha,$$ since $|\eta|^{<\kappa}=\Sigma_{\zeta<\kappa}\,|\eta|^\zeta\le\kappa\cdot\alpha=\alpha$ for every $\eta<\alpha$ whenever $\kappa\ll\alpha$. } \end{remark} Theorem \ref{largeS}, using some of those same ideas, strengthens that result. For use in its proof and frequently thereafter we adopt henceforth the following notational convention concerning limit cardinals $\kappa$. We do not exclude here the possibility that $\kappa$ is regular, but this convention will be invoked chiefly in cases where it is known that $\cf(\kappa)<\kappa$. \begin{notation}\label{defkappaeta} {\rm Let $\kappa\geq\omega$ be a limit cardinal. Then $\{\kappa_\eta:\eta<\cf(\kappa)\}$ is a set of cardinals such that (a) $\kappa_\eta<\kappa_{\eta'}<\kappa \mbox{ when }\eta<\eta'<\cf(\kappa)$, and (b) $\Sigma_{\eta<\cf(\kappa)}\,\kappa_\eta=\kappa.$ } \end{notation} \begin{theorem}\label{largeS} Let $\alpha\geq2$, let $\kappa>\omega$ be a (possibly regular) limit cardinal, and let $\{\kappa_\eta:\eta<\cf(\kappa)\}$ be a family of cardinals as in Notation~\ref{defkappaeta}. For $\eta<\cf(\kappa)$ let $\{X(\eta):\eta<\cf(\kappa)\}$ be a (not necessarily faithfully indexed) set of spaces such that $S(X(\eta))\geq\alpha^{\kappa_\eta}$ for each $\eta<\cf(\kappa)$, and let $X:=\Pi_{\eta<\cf(\kappa)}\,X(\eta)$. Then {\rm (a)} $S(X_{\cf(\kappa)})\geq\alpha^{<\kappa}$; and {\rm(b)} if $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$, then $S(X_{(\cf(\kappa))^+})>\alpha^\kappa\geq(\alpha^{<\kappa})^+$. \end{theorem} \begin{proof} (a) is obvious, since $S(X_{\cf(\kappa)})\geq S(X(\eta))\geq\alpha^{\kappa_\eta}$ for each $\eta<\cf(\kappa)$. (b) The topology of $X_{(\cf(\kappa))^+}$ is the (full) box topology. Since $\cf(\alpha^{<\kappa})=\cf(\kappa)<\kappa$ by Theorem~\ref{weak<kappa}, we may assume without loss of generality that $\alpha^{\kappa_\eta}<\alpha^{\kappa_{\eta'}}$ for $\eta<\eta'<\cf(\kappa)$. Let $\sC(\eta):=\{X(\eta)\}$ for limit ordinals $\eta<\cf(\kappa)$, and for $\eta<\cf(\kappa)$ let $\sC(\eta+1)$ be cellular in $X(\eta+1)$ with $|\sC(\eta+1)|\ge\alpha^{\kappa_{\eta}}$. Then $\sC:=\{\Pi_{\eta<\cf(\kappa)}\,C_\eta:C_\eta\in\sC(\eta)\}$ is cellular in $X_{(\cf(\kappa))^+}$, with $$|\sC|=\Pi_{\eta<\cf(\kappa)}\,|\sC(\eta)| \geq\Pi_{\eta<\cf(\kappa)}\,\alpha^{\kappa_\eta} =\alpha^{\Sigma_{\eta<\cf(\kappa)}\,\kappa_\eta} =\alpha^\kappa,$$ \noindent so $S(X_{(\cf(\kappa))^+})>\alpha^\kappa\geq(\alpha^{<\kappa})^+$. \end{proof} The following simple lemma, strictly set-theoretic (non-topological) in nature, is one of several preliminaries required for the proof of Theorem~\ref{T4.15}. \begin{lemma}\label{L4.13} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals. {\rm (a)} If $\alpha^{<\kappa}$ is a successor cardinal then there is $\nu<\kappa$ such that $\alpha^\nu = \alpha^{<\kappa}$. {\rm (b)} If $\alpha^{<\kappa}$ is a regular cardinal and there is no $\nu<\kappa$ such that $\alpha^\nu = \alpha^{<\kappa}$ then $\alpha^{<\kappa}=\kappa$ and $\kappa$ is a regular strong limit cardinal. \end{lemma} \begin{proof} (a) Let $\alpha^{<\kappa}=\lambda^+$. If $\kappa=\alpha^{<\kappa}$ then $\alpha^{<\kappa}=\alpha^\lambda$ (with $\lambda<\kappa$). If $\kappa<\alpha^{<\kappa}$ and $\alpha^\nu<\alpha^{<\kappa}$ for each $\nu<\kappa$, then we have the contradiction $$\alpha^{<\kappa}=\Sigma_{\nu<\kappa}\,\alpha^\nu\leq\lambda\cdot\kappa= \lambda<\lambda^+=\alpha^{<\kappa}.$$ (b) It follows from (a) that if (b) fails and there is no $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$, then $\alpha^{<\kappa}$ is a (regular) limit cardinal and we have $$\alpha^{<\kappa}=\cf(\alpha^{<\kappa})\leq\cf(\kappa) \leq\kappa\leq\alpha^{<\kappa}.$$ Hence $\alpha^{<\kappa}=\kappa$, and for each $\nu<\kappa$ we have $$2^\nu\leq\alpha^\nu<\alpha^{<\kappa}=\kappa,$$ as required. \end{proof} \begin{remark}\label{nonvac} {\rm It is not difficult to show, as in \cite[3.12]{comfneg74}, that for every uncountable regular cardinal $\alpha$ there is a product space $X_I$ such that $S(X_I)=S((X_I)_\omega)=\alpha$; indeed, as noted there, with $Y:=\Pi_{\beta<\alpha}\,D(\beta)$ one has $S(Y^I)=\alpha$ for all nonempty sets $I$. Thus the instance $S(X_I)=\alpha$ allowed by Theorem~\ref{firstSP} does in fact arise in non-trivial circumstances, provided that uncountable regular limit cardinals $\alpha$ do exist. In any case it is immediate from Theorem~\ref{firstSP} that for every infinite cardinal $\alpha$ of the form $\alpha=\beta^+$ one has $S((D(\beta))^I)=\alpha$ for all nonempty sets $I$. The $\kappa$-box analogue of these statements holds for suitable regular cardinals $\alpha$ (see Theorem~\ref{T4.15}(a) and Remark~\ref{condsi--iv}(b) below), but the full analogue fails consistently (see Remark~\ref{G-Smodels}). We continue with results preparatory to the proof of Theorem~\ref{T4.15}. } \end{remark} \begin{theorem}\label{T4.14} Let $\alpha\ge 3$ and $\kappa\ge\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$ and $\alpha\leq S(X_i)$ for each $i\in I$. Then {\rm (a)} $\alpha^{<\kappa}\leq S((X_I)_\kappa);$ and {\rm (b)} if in addition $\alpha<\kappa$ then $(2^{<\kappa})^+=(\alpha^{<\kappa})^+\le S((X_I)_\kappa).$ \end{theorem} \begin{proof} (a) We consider two cases. \underline{Case 1}. $\alpha$ is singular. Then for each $i\in I$ we have $S(X_i)\ge\alpha^+$ and it follows from Lemma \ref{obs1} (with $\alpha^+$ now replacing $\alpha$) that for each $\lambda<\kappa$ we have $S((X_I)_\kappa)>\alpha^\lambda$. Thus $$S((X_I)_\kappa)\geq\sup_{\lambda<\kappa}\,(\alpha^\lambda)^+\ge \Sigma_{\lambda<\kappa}\,\alpha^\lambda=\alpha^{<\kappa}.$$ \underline{Case 2}. $\alpha$ is regular. (We consider here only the case $\alpha\ge\kappa$ since the case $\alpha<\kappa$ is considered in (b).) Fix $\beta<\alpha$. Since $S(X_i)>\beta$ for each $i\in I$, it follows from Lemma \ref{obs1} that $S((X_I)_\kappa)\ge(\beta^\lambda)^+$ for every $\lambda<\kappa$. Therefore \begin{equation}\label{Eq7} S((X_I)_\kappa)\geq\sup_{\beta<\alpha}\,(\beta^\lambda)^+ \ge\Sigma_{\beta<\alpha}\,\beta^\lambda. \end{equation} Since $\alpha$ is regular and $\lambda<\kappa\leq\alpha$, for each $A\in[\alpha]^\lambda$ there is $\xi<\alpha$ such that $A\subseteq\xi$ (with $|\xi|<\alpha$), so $\alpha^\lambda=\Sigma_{\beta<\alpha}\,\beta^\lambda$. It follows from (\ref{Eq7}) that $S((X_I)_\kappa)\geq\alpha^\lambda$ for each $\lambda<\kappa$. Hence $S((X_I)_\kappa)\geq\alpha^{<\kappa}$, as required. (b) Since $$2^{<\kappa}\leq\alpha^{<\kappa}\leq(2^\alpha)^{<\kappa}= \Sigma_{\lambda<\kappa}\,(2^\alpha)^\lambda= \Sigma_{\lambda<\kappa}\,2^\lambda=2^{<\kappa},$$ we have $$2^{<\kappa}=\alpha^{<\kappa}.$$ Now, fix $\lambda<\kappa$. Since $S(X_i)>2$ for each $i\in I$, it follows from Lemma \ref{obs1} that \begin{equation}\label{Eq8new} S((X_I)_\kappa)\geq S((X_I)_{\lambda^+})\geq(2^\lambda)^+. \end{equation} \underline{Case 1}. There exists $\nu < \kappa$ such that $\alpha^\nu = \alpha^{<\kappa}$. Since $\alpha<\kappa$, without loss of generality, we can assume that $\nu\ge \alpha$. Then $\alpha^\nu=2^\nu$ and from (\ref{Eq8new}) we get $$S((X_I)_\kappa)\geq (2^\nu)^+ > 2^\nu=\alpha^{<\kappa}.$$ \underline{Case 2}. Case 1 fails. If $\alpha^{<\kappa}$ is regular then it follows from Lemma \ref{L4.13}(b) that $\kappa=\alpha^{<\kappa}$ and $\kappa$ is a regular strong limit cardinal. If $\kappa=\omega$ then surely $S((X_I)_\kappa)\geq\kappa^+$, and if $\kappa>\omega$ then Theorem \ref{Obsn}(a) applies to give $$S((X_I)_\kappa)\ge\kappa^+=(\alpha^{<\kappa})^+.$$ Now let $\alpha^{<\kappa}$ be singular. Since (\ref{Eq8new}) holds for every $\lambda<\kappa$ we have $$S((X_I)_\kappa)\geq\sup_{\lambda<\kappa}\,(2^\lambda)^+ \ge\Sigma_{\lambda<\kappa}\,2^\lambda=\alpha^{<\kappa}$$ and since $\alpha^{<\kappa}$ is singular we have $S((X_I)_\kappa)\geq(\alpha^{<\kappa})^+$, as required. \end{proof} \begin{corollary} Let $\alpha$, $\beta$ and $\kappa$ be cardinals with $\alpha\ge 3$ and $\kappa\ge\omega$, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$, and $d(X_i)\leq \beta$ and $\alpha\leq S(X_i)$ for each $i\in I$. Then $\alpha^{<\kappa}\leq S((X_I)_\kappa)\leq(\beta^{<\kappa})^+.$ \end{corollary} \begin{proof} Follows directly from Theorem~\ref{T4.12} and Theorem~\ref{T4.14}. \end{proof} \begin{corollary}\label{L4.14} Let $\alpha\geq3$ and $\kappa\geq\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$ and $d(X_i)\leq\alpha\leq S(X_i)$ for each $i\in I$. Then $\alpha^{<\kappa}\leq S((X_I)_\kappa)\leq(\alpha^{<\kappa})^+.$ \end{corollary} Corollary~\ref{L4.14} provides tight parameters, but leaves undetermined the question exactly when it is that the value of $S((X_I)_\kappa)$ is $\alpha^{<\kappa}$ and when it is $(\alpha^{<\kappa})^+$. In the following theorem we settle that matter completely. \begin{theorem}\label{T4.15} Let $\alpha\geq3$ and $\kappa\geq\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$ and $d(X_i)\leq\alpha\leq S(X_i)$ for each $i\in I$. Consider these conditions: {\rm (i)}~$\alpha$ is regular; {\rm (ii)}~$\alpha=\alpha^{<\kappa}$; {\rm (iii)} $\kappa\ll\alpha$; {\rm (iv)}~$S((X_J)_\kappa)=\alpha$ for all nonempty $J\in[I]^{<\kappa}$. Then: {\rm (a)} if conditions {\rm (i), (ii), (iii)} and {\rm (iv)} hold, then $S((X_I)_\kappa)=\alpha^{<\kappa}=\alpha$; and {\rm (b)} if one (or more) of conditions {\rm (i), (ii), (iii)} or {\rm (iv)} fails, then $S((X_I)_\kappa)=(\alpha^{<\kappa})^+$. \end{theorem} \begin{proof} (a) is immediate from Theorem~\ref{CN}, since $S(X_i)=\alpha=\alpha^{<\kappa}$ for each $i\in I$ under the present hypotheses. (b) It suffices, according to Corollary~\ref{L4.14}, to assume that $S((X_I)_\kappa)=\alpha^{<\kappa}$ and to show that conditions (i), (ii), (iii) and (iv) must hold. We consider two cases. \underline{Case 1}. There is $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. We fix such $\nu$. If (i) fails then $S(X_i)=\alpha^+$ and from Lemma \ref{obs1} (with $\alpha^+$ and $\nu$ in the roles of $\alpha$ and $\mu$, respectively) we have $S((X_I)_\kappa)\geq S((X_I)_{\nu^+})>\alpha^\nu=\alpha^{<\kappa}$, a contradiction. Thus (i) holds. To see that (ii) holds, suppose first that there are $\beta<\alpha$ and $\lambda<\kappa$ such that $\beta^\lambda\geq\alpha$. Then $$\alpha^{<\kappa}=\alpha^\nu\leq\beta^{\lambda\cdot\nu}\leq\alpha^{\lambda\cdot\nu}\leq\alpha^{<\kappa}$$ so from Lemma~\ref{obs1} we conclude that $S((X_I)_\kappa)>\beta^{<\kappa}\geq\beta^{\lambda\cdot\nu}=\alpha^{<\kappa}$, a contradiction. Thus \begin{equation}\label{Eq8} \beta^\lambda<\alpha \mbox{ for all } \beta<\alpha, \lambda<\kappa. \end{equation} It follows that $\kappa\leq\alpha$. Then each $\lambda<\kappa$ satisfies $\lambda<\alpha$ and from the regularity of $\alpha$ we have $[\alpha]^\lambda=\bigcup_{\beta<\alpha}\,[\beta]^\lambda$ for each such $\lambda$. Thus (\ref{Eq8}) gives $$\alpha^{<\kappa}=\Sigma_{\lambda<\kappa}\,\alpha^\lambda\leq \Sigma_{\lambda<\kappa}[\Sigma_{\beta<\alpha}\,\beta^\lambda] \leq\kappa\cdot\alpha\cdot\alpha=\alpha\leq\alpha^{<\kappa},$$ and (ii) is proved. To show (iii) we need only show $\kappa<\alpha$, since (\ref{Eq8}) then gives $\kappa\ll\alpha$. Suppose then that $\kappa=\alpha$. Then (\ref{Eq8}) shows that $\kappa$ is a (regular, strong) limit cardinal, so from Theorem \ref{Obsn} we have the contradiction $S((X_I)_\kappa)>\kappa=\alpha=\alpha^{<\kappa}$. Thus $\kappa<\alpha$ and the proof of (iii) is complete. To prove (iv), it suffices to note that if $S((X_J)_\kappa)>\alpha$ for some nonempty $J\in[I]^{<\kappa}$, then we have the contradiction $S((X_I)_\kappa)>\alpha=\alpha^{<\kappa}$. \underline{Case 2}. There is no $\nu<\kappa$ such that $\alpha^\nu=\alpha^{<\kappa}$. If $\alpha^{<\kappa}$ is singular then $S((X_I)_\kappa)=\alpha^{<\kappa}$ is impossible, so $\alpha^{<\kappa}$ is regular and Lemma~\ref{L4.13}(b) applies to show that $\alpha^{<\kappa}=\kappa$ is a (regular, strong) limit cardinal; from Theorem \ref{Obsn} we again have the contradiction $S((X_I)_\kappa)>\kappa=\alpha^{<\kappa}$. \end{proof} Although every infinite Souslin number is regular and uncountable, hence is either a successor cardinal or an uncountable regular limit cardinal, it is perhaps not clear from Theorem~\ref{T4.15} exactly which uncountable regular cardinals occur in the form $S((X_I)_\kappa)$ with $\{X_i:i\in I\}$ constrained as in Theorem~\ref{T4.15}. Is part (a) of that theorem potentially vacuous? Can every successor cardinal $\beta^+$ occur as $\beta^+=\alpha$ in Theorem~\ref{T4.15}(a)? For each $\kappa$, can some $\beta^+=\alpha$ so occur? Do there exist, for every regular limit cardinal $\alpha$, infinite $\kappa\ll\alpha$ and spaces $\{X_i:i\in I\}$ such that $S((X_I)_\kappa)=\alpha$? We address these questions in \ref{condsi--iv}---\ref{S=lim} below. \begin{remarks}\label{condsi--iv} {\rm (a) Let $\beta$ be a singular cardinal and set $\alpha:=\beta^+$. Let $I$ be an uncountable set and for $i\in I$ set $X_i:=D(\beta)$. Clearly (i), (ii) and (iii) are satisfied with $\kappa=\omega$; also (iv) is satisfied with $\kappa=\omega$, since if $J\in[I]^{<\omega}$ then $X_J=X^J$ is discrete with $|X_J|=\beta$, so $S(X_I)=S((X_I)_\omega)=\beta^+=\alpha$. Thus $S(X_I)=S((X_I)_\omega)=\alpha$ by Theorem~\ref{T4.15}(a). The same conclusion is available from Theorem~\ref{T4.15}(b) by replacing $\alpha$ everywhere in the statement of Theorem~\ref{T4.15} by $\beta$. In this case both (i) and (iv) fail for $\beta$, so $S(X_I)=S((X_I)_\omega)=\beta^+=\alpha$ by Theorem~\ref{T4.15}(b). (b) Similar examples exist in ZFC for every uncountable regular cardinal $\kappa$. Indeed, given such $\kappa$ let $\gamma\geq2$ be arbitrary and set $\beta:=\beth_\kappa(\gamma)$ and $\alpha:=\beta^+$. For $\lambda<\kappa$ we have $$[\beta]^\lambda=\bigcup_{\delta<\beta}\,[\delta]^\lambda \mbox{ and } [\alpha]^\lambda=\bigcup_{\xi<\alpha}\,[\xi]^\lambda,$$ so $$\beta^\lambda\leq\Sigma_{\delta<\beta}\,\delta^\lambda \leq\Sigma_{\delta<\beta}\,2^\delta\cdot2^\lambda=\beta$$ and $$\alpha^\lambda\le\alpha\cdot\beta^\lambda\leq\alpha\cdot\beta=\alpha.$$ Conditions (i), (ii) and (iii) are then clear and again, as in (a), if $|I|>\kappa$ and $X_i:=D(\beta)$ for each $i\in I$, then each space $(X^J)_\kappa=(X_J)_\kappa$ is discrete (when $|J|<\kappa$) with $|X_J|=\beta$, so $S((X_J)_\kappa)=\beta^+=\alpha$ and (iv) holds by Theorem~\ref{T4.15}(a). Also as in (a) above the same conclusion is available from Theorem~\ref{T4.15}(b) by replacing $\alpha$ everywhere in the statement of Theorem~\ref{T4.15} by $\beta$. In this case both (i) and (iv) fail for $\beta$, so $S((X_I)_\kappa)=\beta^+=\alpha$ by Theorem~\ref{T4.15}(b). (c) (a) and (b) above indicate that in all models of ZFC conditions (i), (ii), (iii) and (iv) of Theorem~\ref{T4.15}(a) are satisfied by suitably chosen cardinals and spaces, so part (a) of Theorem~\ref{T4.15} is not vacuous. Those examples depend, however, on choosing for $\alpha$ a regular cardinal of the form $\alpha=\beta^+$. Part (a) of Theorem~\ref{T4.20} shows exactly which successor cardinals $\gamma^+$ arise as $S((X_I)_\kappa)$ in Theorem~\ref{T4.15}, and part (b) indicates when it can occur that $S((X_I)_\kappa)$ is a limit cardinal. } \end{remarks} \begin{theorem}\label{T4.20} Let $\alpha\geq3$ and $\kappa\geq\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$ and $d(X_i)\leq\alpha\leq S(X_i)$ for each $i\in I$. {\rm (a)} If $S((X_I)_\kappa)$ is a successor cardinal---say $S((X_I)_\kappa)=\gamma^+$---then either conditions {\rm (i), (ii), (iii)} and {\rm (iv)} of Theorem \ref{T4.15} hold and $\gamma=\gamma^{<\kappa}$, or at least one of those conditions fails and $\gamma=\alpha^{<\kappa}$. {\rm (b)} If $S((X_I)_\kappa)$ is a (regular) limit cardinal, then $S((X_I)_\kappa)=\alpha$ and $d(X_i)=S(X_i)=\alpha$ for each $i\in I$. \end{theorem} \begin{proof} (a) If $\gamma^+=S((X_I)_\kappa)\neq(\alpha^{<\kappa})^+$ then by Theorem~\ref{T4.15} the indicated conditions (i), (ii), (iii) and (iv) all hold and $\gamma^+=S((X_I)_\kappa)=\alpha=\alpha^{<\kappa}$. Since $\kappa\ll\alpha=\gamma^+$ we have $\gamma^\lambda=\gamma$ for all $\lambda<\kappa$ and hence $\gamma^{<\kappa}\leq\kappa\cdot\gamma=\gamma$, so $\gamma=\gamma^{<\kappa}$ as asserted. (b) If $S((X_I)_\kappa)$ is a regular limit cardinal then conditions (i), (ii), (iii) and (iv) of Theorem~\ref{T4.15} all hold, so $S((X_I)_\kappa)=\alpha$ by Theorem~\ref{T4.15}(a); further for each $i\in I$ we have $S(X_i)=\alpha$ by condition (iv). If there is $i\in I$ such that $d(X_i)<\alpha$ then $(d(X_I))^+<\alpha=S(X_i)$, which is impossible. \end{proof} \begin{remark}\label{S=lim} {\rm It is well known that the existence of an uncountable regular strong limit cardinal cannot be established in ZFC \cite[12.12]{jech}, but it should be noted that in case such a cardinal $\alpha$ exists then there are cardinals $\kappa$ and spaces $X_i$ to which Theorem~\ref{T4.15}(a) and Theorem~\ref{T4.20} apply. Indeed, let $\alpha$ be a regular limit cardinal and suppose that $\kappa$ satisfies $\omega\leq\kappa\ll\alpha=\alpha^{<\kappa}$. (These latter conditions are satisfied by every infinite $\kappa<\alpha$, in case $\alpha$ is in addition assumed to be a strong limit cardinal.) Let $I$ be a nonempty set and for $\beta<\alpha$ and $i\in I$ set $D(\beta,i):=D(\beta)$. Then $\{D(\beta,i):\beta<\alpha,i\in I\}$ is a set of spaces with each $d(D(\beta,i))=\beta<\alpha$, so by Theorem~\ref{iffk<<a} the space $Y:=\Pi_{\beta<\alpha, i\in I}\,D(\beta,i)$ satisfies $S((Y)_\kappa)=\alpha$. As a set we have $Y=X^I$ with $X:=\Pi_{\beta<\alpha}\,D(\beta)$, and the topology of the space $Y_\kappa$ is finer than the topology of the space $(X^I)_\kappa$, so also the power space $X^I$ satisfies $S((X^I)_\kappa)=\alpha$, where $\alpha$ is a (regular strong) limit cardinal. Clearly $d(X)=\alpha$ and $S(X)=\alpha$ and therefore $(X^I)_\kappa$ is an example of a product space that satisfies all the hypotheses and the conclusion of Theorem \ref{T4.15}(a). } \end{remark} \begin{lemma}\label{SLC} Let $\kappa$ be a strong limit cardinal. {\rm (a)} If $2\leq\alpha<\kappa$ then $\alpha^{<\kappa}=\kappa$. {\rm (b)} If $\kappa$ is regular then $\kappa^{<\kappa}=\kappa$. {\rm (c)} If $\kappa$ is singular then $\kappa^{<\kappa}=2^\kappa$. \end{lemma} \begin{proof} (a) We have $$\kappa\leq\alpha^{<\kappa}\leq(2^\alpha)^{<\kappa}=\Sigma_{\lambda<\kappa}\,(2^\alpha)^\lambda =\Sigma_{\lambda<\kappa}\,2^\lambda\leq\kappa\cdot\kappa=\kappa.$$ (b) This is proved in Lemma~\ref{regularSLC}(a). (c) With $\{\kappa_\eta:\eta<\cf(\kappa)\}$ chosen as in Notation~\ref{defkappaeta} we have $$2^\kappa=2^{\Sigma_{\eta<\cf(\kappa)}\,\kappa_\eta}=\Pi_{\eta<\cf(\kappa)}\,2^{\kappa_\eta} \leq\kappa^{cf(\kappa)}\leq\kappa^{<\kappa}\leq2^\kappa.$$ \vskip-18pt \end{proof} \begin{theorem}\label{Spower} Let $\kappa$ be a strong limit cardinal and $I$ be an index set with $|I|\ge\kappa$. {\rm (a)} If $2\leq\alpha<\kappa$ then $S(((D(\alpha))^I)_\kappa)=\kappa^+$. {\rm (b)} If $\kappa$ is regular then $S(((D(\kappa))^I)_\kappa)=\kappa^+$. {\rm (c)} If $\kappa$ is singular then $S(((D(\kappa))^I)_\kappa)=(2^\kappa)^+$. \end{theorem} \begin{proof} In each case, condition (iii) of Theorem~\ref{T4.15} fails, so parts (a), (b) and (c) follow from Theorem~\ref{T4.15}(b) and from parts (a), (b) and (c) of Lemma~\ref{SLC}, respectively. \end{proof} \begin{remark}\label{S(G-S)} {\rm As we noted in Discussion \ref{gitik-shelah}(f) the value of $d((\two^{\aleph_\omega})_{\aleph_\omega}$ depends on the model of {\rm ZFC}, while the findings we have enunciated here are sufficiently powerful that the weight and Souslin number of such spaces as $(\two^{\aleph_\omega})_{\aleph_1}$ and $(\two^{\aleph_\omega})_{\aleph_\omega}$ in $\VV_1$ and $\VV_2$ now emerge painlessly. To make those computations, recall that $2^{\aleph_n}=\aleph_{n+1}$ and $2^{\aleph_\omega}=\aleph_\omega^\omega=\aleph_{\omega+2}$ there, so we have from Theorem~\ref{T2.7}(b) $$w((\two^{\aleph_\omega})_{\aleph_1})=\aleph_\omega^\omega=\aleph_{\omega+2},$$ also $$w((\two^{\aleph_\omega})_{\aleph_\omega})=(\aleph_\omega)^{<\aleph_\omega}=\aleph_{\omega+2}$$ in both those models. Concerning the Souslin number, it is clear that $S((\two^{\aleph_0})_{\aleph_1})=(2^{\aleph_0})^+=\cc^+$ in ZFC, so from Corollary~\ref{S(Xkappa)} we have for each nonempty set $I$ that $$\cc^+=S((\two^{\aleph_0})_{\aleph_1})\leq S((\two^I)_{\aleph_1})\leq\cc^+,$$ hence $S((\two^{\aleph_\omega})_{\aleph_1})=\cc^+$ in ZFC (with $\cc^+=\aleph_2$ in the models $\VV_1$ and $\VV_2$). Finally, from Theorem~\ref{Spower}(a) (or from Theorem \ref{T4.15}(b)) we have $$S((\two^{\aleph_\omega})_{\aleph_\omega})=(\aleph_\omega)^+=\aleph_{\omega+1}$$ in $\VV_1$ and $\VV_2$. } \end{remark} In Theorem \ref{GKurepa} below we give an upper bound for the Souslin number of a product space with the $\kappa$-box topology that depends only on the Souslin numbers of its coordinate spaces, rather than (as in Theorem \ref{GCN}) on the Souslin numbers of its ``small" sub-products. For that we need the following notation (see \cite{comfneg74}, \cite{comfneg82}, \cite{juh}). \begin{notation} {\rm Let $\alpha$, $\beta$, $\kappa$ and $\lambda$ be cardinals. The \emph{arrow notation} $\alpha\rightarrow(\kappa)^\beta_\lambda$ denotes the following partition relation: if $[\alpha]^\beta=\cup_{i<\lambda}\,P_i$ then there are $A\subseteq\alpha$ and $i<\lambda$ such that $|A|=\kappa$ and $[A]^\beta\subseteq P_i$. } \end{notation} Preliminary to Theorem~\ref{GKurepa} we give a combinatorial lemma which makes plain the relevance of the arrow relation $\alpha\rightarrow(\kappa)^2_\lambda$ to numbers of the form $S((X_I)_{\lambda^+})$. The (general) proof we give is as anticipated in \cite[page~73]{comfneg74}; it parallels in all its essentials that of the special case treated in \cite[3.13]{comfneg74}. We remark that results significantly stronger than that of Lemma~\ref{arrow}, which have perhaps not received the attention or the recognition they deserve, were developed by Negrepontis and his school in Athens in the 1970's. It is shown in \cite[5.17]{comfneg82} for example, using the hypothesis $\omega\leq\lambda<\kappa\ll\alpha$ with $\kappa$ and $\alpha$ regular, that if each $S(X_i)\leq\kappa$ then not only is $S((X_I)_{\lambda^+})\leq\alpha$, as in Lemma~\ref{arrow}, but in fact of every $\alpha$-many nonempty open subsets of $(X_I)_{\lambda^+}$ some $\alpha$-many have the finite intersection property. \begin{lemma}\label{arrow} Let $\alpha$, $\kappa$ and $\lambda$ be infinite cardinals such that $\alpha\rightarrow(\kappa)^2_\lambda$, and let $\{X_i:i\in I\}$ be a set of spaces such that $S(X_i)\leq\kappa$ for each $i\in I$. Then $S((X_I)_{\lambda^+})\leq\alpha$.\end{lemma} \begin{proof} Suppose that there is a faithfully indexed cellular family $\{U^\xi:\xi<\alpha\}$ of basic open subsets of $(X_I)_{\lambda^+}$, and for $\{\xi,\xi'\}\in[\alpha]^2$ let $i(\xi,\xi')\in I$ be such that $U^\xi_{i(\xi,\xi')}\cap U^{\xi'}_{i(\xi,\xi')}=\emptyset$. For $\xi<\alpha$ we define $$I(\xi):=\{i(\xi,\xi'):\xi'<\alpha \mbox{ and }\xi\ne\xi'\}.$$ Since $i(\xi,\xi')\in R(U^\xi)\cap R(U^{\xi'})$ for $\{\xi,\xi'\}\in[\alpha]^2$ we have $|I(\xi)|\le|R(U^\xi)|\leq\lambda$ for $\xi<\alpha$. Let $\{i_{\xi,\eta}:\eta<\lambda\}$ be an indexing of $I(\xi)$ for $\xi<\alpha$, and for $(\eta,\eta')\in \lambda\times\lambda$ set $$P_{\eta,\eta'}:=\{\{\xi,\xi'\}\in[\alpha]^2:\xi<\xi'\mbox{ and } i_{\xi,\eta}=i_{\xi',\eta'}\}$$ (some of the sets $P_{\eta,\eta'}$ might be empty). Since $$[\alpha]^2=\bigcup_{(\eta,\eta')\in\lambda\times\lambda}\,P_{\eta,\eta'}$$ and $\alpha\rightarrow(\kappa)^2_\lambda$, there are $A\in[\alpha]^\kappa$ and $(\overline{\eta},\overline{\eta'})\in\lambda\times\lambda$ such that $[A]^2\subseteq P_{\overline{\eta},\overline{\eta'}}$. Thus there is $\overline{i}\in I$ such that if $\{\xi,\xi'\}\in[A]^2$ and $\xi<\xi'$ then $i(\xi,\xi')=i_{\xi,\overline{\eta}}=i_{\xi',\overline{\eta'}}=\overline{i}$ and hence $U^\xi_{\overline{i}}\cap U^{\xi'}_{\overline{i}}=\emptyset$. It follows that $\{U^\xi_{\overline{i}}:\xi\in A\}$ is cellular in $X_{\overline{i}}$ and we have the contradiction $S(X_{\overline{i}})>|A|=\kappa$. \end{proof} The following theorem is \cite[Theorem 1.5(a)]{comfneg82}. It is noted in \cite{comfneg82} that preliminary formulations of Theorem~\ref{CNT1.5} appear (with different hypotheses) in Erd{\H{o}}s and Rado~\cite[39(iii)]{erdosrado56} and Kurepa~\cite{kurepa59a}. The important and motivating special case $(2^\alpha)^+\rightarrow(\alpha^+)^2_\alpha$ of Theorem~\ref{CNT1.5} appeared as early as 1942~\cite{erdos42}, while the seminal instance $\kappa=\omega$, $\alpha=\omega^+$ of Theorem~\ref{GKurepa}(a) was given by Kurepa~\cite{kurepa62} (see also \cite[Theorem 3.13 and remark on pp.~73-74]{comfneg74}). \begin{theorem}\label{CNT1.5} If $\omega\le \kappa \ll \alpha$ with $\alpha$ and $\kappa$ regular, then $\alpha\rightarrow(\kappa)^2_\lambda$ for all $\lambda < \kappa$. \end{theorem} \begin{theorem}\label{GKurepa} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals and let $\{X_i:i\in I\}$ be a family of nonempty spaces such that $S(X_i)\le\alpha^+$ for each $i\in I$. {\rm (a)} If $\alpha^+\ge\kappa$ then $S((X_I)_\kappa)\le(2^\alpha)^+$; and {\rm (b)} if $\alpha^+\leq \kappa$ then $S((X_I)_\kappa)\le((\alpha^{<\kappa})^{<\kappa})^+$. \end{theorem} \begin{proof} (a) Clearly $\alpha\geq\omega$ here, so Remark~\ref{help}(2) applies to give $\alpha^+\ll(2^\alpha)^+$. We then have $$(2^\alpha)^+\rightarrow(\alpha^+)^2_\alpha$$ by Theorem~\ref{CNT1.5}, so $S((X_I)_{\alpha^+})\leq(2^\alpha)^+$ by Lemma~\ref{arrow} (with $(2^\alpha)^+$, $\alpha^+$, and $\alpha$ in the role of $\alpha$, $\kappa$, and $\lambda$ there). Thus surely $S((X_I)_\kappa)\leq(2^\alpha)^+$ if $\kappa\leq\alpha^+$. (b) We consider three cases. \underline{Case 1}. $\kappa$ is singular. Then $\kappa^+\ll((\alpha^{<\kappa})^{<\kappa})^+$ by Remark~\ref{help}(4), hence we have from Theorem~\ref{CNT1.5} that $$((\alpha^{<\kappa})^{<\kappa})^+\rightarrow(\kappa^+)^2_\kappa.$$ Since $\alpha\leq\kappa$ we have $S(X_i)\leq\kappa^+$ for each $i\in I$, so in fact even $$S((X_I)_\kappa)\leq S((X_I)_{\kappa^+})\leq((\alpha^{<\kappa})^{<\kappa})^+$$ by Lemma~\ref{arrow} (with $((\alpha^{<\kappa})^{<\kappa})^+$, $\kappa^+$, and $\kappa$ in the role of $\alpha$, $\kappa$, and $\lambda$ there). \underline{Case 2}. $\kappa$ is a successor cardinal, say $\kappa=\lambda^+$. Since $\lambda^+\ll(\alpha^{\lambda})^+$, by Theorem~\ref{CNT1.5} we have $$(\alpha^\lambda)^+\rightarrow(\lambda^+)^2_\lambda,$$ so $$S((X_I)_{\kappa})\leq(\alpha^\lambda)^+=((\alpha^{<\kappa})^{<\kappa})^+$$ by Lemma~\ref{arrow} (with $(\alpha^\lambda)^+$, $\lambda^+$, and $\lambda$ in the role of $\alpha$, $\kappa$, and $\lambda$ there). \underline{Case 3}. $\kappa$ is regular limit cardinal. Then it follows from Remark \ref{help}(3) that $\omega\le \kappa \ll ((\alpha^{<\kappa})^{<\kappa})^+$ and therefore, according to Theorem \ref{CNT1.5}, $((\alpha^{<\kappa})^{<\kappa})^+\rightarrow(\kappa)^2_\lambda$ for all $\lambda<\kappa$. Then since $S(X_i)\leq\kappa$ for each $i\in I$ we have \begin{equation}\label{Eq9} S((X_I)_{\lambda^+})\le ((\alpha^{<\kappa})^{<\kappa})^+ \mbox{ for all } \lambda<\kappa \end{equation} \noindent from Lemma~\ref{arrow} (with $((\alpha^{<\kappa})^{<\kappa})^+$ in the role of $\alpha$ there). Now suppose that $\sC$ is a cellular family in $(X_I)_\kappa$ of canonical open sets such that $|\sC|=((\alpha^{<\kappa})^{<\kappa})^+$, and for $\lambda<\kappa$ let $\sC(\lambda):=\{U\in\sC:|R(U)|<\lambda\}$. Then $\sC=\bigcup_{\lambda<\kappa}\,\sC(\lambda)$ with $$\cf(((\alpha^{<\kappa})^{<\kappa})^+)=((\alpha^{<\kappa})^{<\kappa})^+\geq\kappa^+>\kappa,$$ so there is $\lambda<\kappa$ such that $|\sC(\lambda)|=((\alpha^{<\kappa})^{<\kappa})^+$. Then $$S((X_I)_{\lambda^+})>|\sC(\lambda)|=((\alpha^{<\kappa})^{<\kappa})^+,$$ contrary to (\ref{Eq9}). \end{proof} \begin{remark} {\rm We note that in those cases of Theorem~\ref{GKurepa} to which both (a) and (b) apply, namely when $\kappa=\alpha^+$, the upper bounds provided by the estimates in (a) and (b) coincide. Indeed with $\kappa=\alpha^+$ we have $$((\alpha^{<\kappa})^{<\kappa})^+=((\alpha^\alpha)^\alpha)^+=(2^\alpha)^+.$$ } \end{remark} Combining Theorem \ref{T4.14} and Theorem \ref{GKurepa} we obtain the following corollary. \begin{corollary}\label{C4.37} Let $\alpha\geq3$ and $\kappa\geq\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$ and $\alpha\le S(X_i)\le \alpha^+$ for each $i\in I$. {\rm (a)} If $\alpha^+\ge\kappa$ then $\alpha^{<\kappa}\le S((X_I)_\kappa)\le(2^\alpha)^+$; and {\rm (b)} if $\alpha^+\leq \kappa$ then $(2^{<\kappa})^+\le S((X_I)_\kappa)\le((2^{<\kappa})^{<\kappa})^+$. \end{corollary} \begin{corollary}\label{C4.40} Let $\alpha\geq3$ and $\kappa\geq\omega$ be cardinals, and let $\{X_i:i\in I\}$ be a set of spaces such that $|I|^+\geq\kappa$ and $3\le S(X_i)\le \alpha^+$ for each $i\in I$. If $\alpha^+\le\kappa$ then $S((X_I)_\kappa)=(2^{<\kappa})^+$. \end{corollary} \begin{proof} The case $3=\alpha<\kappa$ of Theorem~\ref{T4.14}(b) gives $S((X_I)_\kappa\geq(2^{<\kappa})^+$. If $\kappa$ is regular or there is $\nu<\kappa$ such that $2^\nu=2^{<\kappa}$ then $2^{<\kappa}=(2^{<\kappa})^{<\kappa}$ by Theorem~\ref{weak<kappa}(a) and the statement is immediate from Corollary \ref{C4.37}(b). Now we assume that $\kappa$ is singular and that $2^\nu<2^{<\kappa}$ for each $\nu<\kappa$ and we let $\{\kappa_\eta:\eta<\cf(\kappa)\}$ be a set of cardinals as in Notation~\ref{defkappaeta}. Suppose that $S((X_I)_\kappa)>(2^{<\kappa})^+$, and let $\sC$ be a cellular family of basic open sets in $(X_I)_\kappa$ with $|\sC|=(2^{<\kappa})^+$. Then with $$\sC(\eta):=\{C:C\in \sC \mbox{ and } C \mbox{ is open in } (X_I)_{\kappa_\eta}\}$$ for $\eta<\cf(\kappa)$ we have $\sC=\cup_{\eta<\cf(\kappa)}\,\sC(\eta)$, and since $\cf((2^{<\kappa})^+)>\cf(\kappa)$ there is $\eta<\cf(\kappa)$ such that $|\sC(\eta)| = (2^{<\kappa})^+$; for this $\eta$ we have \begin{equation}\label{Eq8'} S((X_I)_{\kappa_\eta})>(2^{<\kappa})^+. \end{equation} Since $\kappa$ is singular we have from $\alpha^+\leq\kappa$ that $\alpha<\kappa$, so there is $\eta'<\cf(\kappa)$ such that $\alpha<\kappa_{\eta'}$; we take $\eta'\geq\eta$. Then from Theorem~\ref{GKurepa}(b) with $\kappa_{\eta'}^+$ replacing $\kappa$ we have $$S((X_I)_{\kappa_\eta})\leq S((X_I)_{\kappa_{\eta'}^+})\leq(\alpha^{\kappa_{\eta'}})^+ \leq((2^\alpha)^{\kappa_{\eta'}})^+=(2^{\kappa_{\eta'}})^+<2^{<\kappa},$$ which contradicts (\ref{Eq8'}). \end{proof} \begin{remarks}\label{G-Smodels} {\rm (a) We noted in Remark~\ref{condsi--iv} that the conditions given in Theorem~\ref{T4.15} are satisfied by many pairs $\kappa$, $\alpha$ of cardinals and for many sets $\{X_i:i\in I\}$ of spaces; in particular (see condition (iv) of Theorem~\ref{T4.15}) for $\kappa\ll\alpha=\alpha^{<\kappa}$ there are spaces $X$ such that $S((X^I)_\kappa)=\alpha$ for all nonempty index sets $I$. We note now that consistently there are (regular) $\alpha$ and $\kappa$ for which the relation $S((X^I)_\kappa)=\alpha$ holds for no space $X$ and infinite index set $I$. Indeed, let $\VV$ be one of the Gitik-Shelah models whose salient cardinality properties are given in Discussion \ref{gitik-shelah}(d) and let $\kappa=\aleph_1$ and $\alpha=\aleph_{\omega+1}$. Suppose there is a space $X$ such that $S(X)=\alpha$ and $S((X^I)_\kappa)=\alpha$ for some infinite set $I$. Then $S(X)=\alpha=\aleph_{\omega+1}>\aleph_{\omega}$, and it follows from Lemma~\ref{obs1} that $S((X^I)_\kappa)>(\aleph_{\omega})^\omega=\aleph_{\omega+2}$, a contradiction. (b) The upper bound $S(X\times X)=(2^\alpha)^+$ for spaces $X$ such that $S(X)=\alpha^+$, allowed by Theorem~\ref{GKurepa}(a), is in fact achieved for many $\alpha$ and $X$. This was first shown by Galvin and Laver (cf.~\cite{galvin}) assuming $\alpha^+=2^\alpha$ (see \cite[7.13]{comfneg82} for a treatment of the construction) and by examples in ZFC by Todor\v{c}evi\'{c} \cite{todor85}, \cite{todori}, \cite{todorii}. When $\alpha^+=2^\alpha$ this strict increase from $S(X)$ to $S(X\times X)$ is minimal in the sense that $$S(X\times X)=(2^\alpha)^+=(\alpha^+)^+=(S(X))^+.$$ \noindent We note in contrast that in the models discussed in (a) there are spaces $X$ such that $S(X^I)\ge (S(X))^{++}$, for every infinite set $I$. For example, for any space $X$ in those models satisfying $S(X)=\alpha=\aleph_{\omega+1}$ and with $\kappa=\aleph_1$ we have $$((\aleph_{\omega+1})^+)^+=\aleph_{\omega+3}\leq S((X^I)_\kappa) \le(2^\alpha)^+=(2^{\aleph_{\omega+1}})^+$$ in these models. Similarly Fleissner~\cite[Section~5]{fleissner78}, in suitably defined Cohen models of ZFC, constructs spaces $X$ for which $S(X)=\omega^+=\aleph_1$ and $S(X\times X)=\aleph_{\omega+2}>\cc=\aleph_{\omega+1}$. } \end{remarks} The rest of this section is devoted to seeking definitive relations between and among the cardinals $S((X_I)_\kappa)$, $S((X_J)_\kappa)$ with $J\in[I]^{<\kappa}$, and $(\alpha^{<\kappa})^+$. Our success, though substantial, is only partial, since we have been unable to give a fully satisfactory answer to Question~\ref{Q1} in ZFC. \begin{theorem}\label{T4.16} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals such that $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$. If $\{X_i:i\in I\}$ is a set of nonempty spaces such that $S((X_I)_\kappa)>(\alpha^{<\kappa})^+$, then there are a cardinal $\lambda<\kappa$ and $J\in[I]^{<\lambda}$ such that $S((X_J)_{\lambda})\geq(\alpha^{<\kappa})^+$. \end{theorem} \begin{proof} Let $\sC$ be a cellular family of basic open subsets of $(X_I)_\kappa$ such that $|\sC|=(\alpha^{<\kappa})^+$. Let $\{\kappa_\eta:\eta<\cf(\kappa)\}$ be as in Notation~\ref{defkappaeta} and for $\eta<\cf(\kappa)$ set $\sC(\eta):=\{U\in\sC:|R(U)|<\kappa_\eta\}$. Since $|\sC|=\bigcup_{\eta<\cf(\kappa)}\,\sC(\eta)$ and $\cf((\alpha^{<\kappa})^+)=(\alpha^{<\kappa})^+\geq\kappa^+>\cf(\kappa)$, there is $\eta<\cf(\kappa)$ (henceforth fixed) such that $|\sC(\eta)|=(\alpha^{<\kappa})^+$. Then $\sC(\eta)$ is cellular in $(X_I)_{\kappa_\eta}$, hence in $(X_I)_{\kappa_\eta^+}$, and for $\eta<\eta'<\cf(\kappa)$ we have $$S((X_I)_{\kappa_\eta^+})>|\sC(\eta)| =(\alpha^{<\kappa})^+>(\alpha^{\kappa_{\eta'}})^+.$$ Then since $\kappa_\eta^+\ll(\alpha^{\kappa_{\eta'}})^+$ there is, by Theorem~\ref{CN} (with $\kappa_\eta^+$ and $(\alpha^{\kappa_{\eta'}})^+$ in the roles of $\kappa$ and $\alpha$ respectively), a set $J(\eta')\in[I]^{<\kappa_\eta^+}$ such that $S((X_{J(\eta')})_{\kappa_\eta^+})>(\alpha^{\kappa_{\eta'}})^+$. Then with $J:=\bigcup_{\eta<\eta'<\cf(\kappa)}\,J(\eta')$ we have $|J|\leq\kappa_\eta\cdot\cf(\kappa)<\kappa$, and $$S((X_J)_{\kappa_\eta^+})>(\alpha^{\kappa_{\eta'}})^+ \mbox{ when }\eta<\eta'<\cf(\kappa),$$ hence $$S((X_J)_{\kappa_\eta^+})\ge\sup_{\eta<\eta'<\cf(\kappa)}(\alpha^{\kappa_{\eta'}})^+=\Sigma_{\eta<\eta'<\cf(\kappa)}\, (\alpha^{\kappa_{\eta'}})^+=\alpha^{<\kappa}.$$ Since in our case $\alpha^{<\kappa}$ is singular (Theorem \ref{weak<kappa}(b)) we have $$S((X_J)_{\kappa_\eta^+})\geq(\alpha^{<\kappa})^+,$$ so the conclusion holds with $\lambda:=\max\{\kappa_\eta^+,|J|^+\}$. \end{proof} We continue in Corollary~\ref{GCNan} with a consequence of Theorem~\ref{T4.16} for which Lemma~\ref{newl} is preparatory. \begin{lemma}\label{newl} Let $\kappa\ge\omega$ be a limit cardinal, $\alpha\geq2$ be a cardinal and $\{X_i:i\in I\}$ be a set of nonempty spaces. Then $S((X_J)_\lambda)<\alpha$ for each $\lambda<\kappa$ and each nonempty $J\in[I]^{<\lambda}$ if and only if $S((X_J)_\kappa)<\alpha$ for each nonempty $J\in[I]^{<\kappa}$. \end{lemma} \begin{proof} Let $S((X_J)_\kappa)<\alpha$ for each nonempty $J\in[I]^{<\kappa}$ and let $\lambda < \kappa$ and $\emptyset\neq J_0\in[I]^{<\lambda}$. Then $S((X_{J_0})_\lambda)=S((X_{J_0})_\kappa)<\alpha$ since the spaces $(X_{J_0})_\lambda$ and $(X_{J_0})_\kappa$ have the full box topology and therefore coincide. For the converse, let $S((X_J)_\lambda)<\alpha$ for each $\lambda<\kappa$ and each nonempty $J\in[I]^{<\lambda}$ and let $\emptyset\neq J_0\in[I]^{<\kappa}$. Since $|J_0|<\kappa$ and $\kappa$ is a limit cardinal, there exists $\lambda<\kappa$ such that $|J_0|<\lambda$. Then $S((X_{J_0})_\kappa)=S((X_{J_0})_\lambda)<\alpha$ since the spaces $(X_{J_0})_\kappa$ and $(X_{J_0})_\lambda$ have the full box topology and therefore coincide. \end{proof} \begin{corollary}\label{GCNan} Let $\alpha\geq2$ and $\kappa\ge\omega$ be cardinals such that $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$ and let $\{X_i:i\in I\}$ be a set of nonempty spaces such that $S((X_J)_\kappa)<(\alpha^{<\kappa})^+$ for each nonempty $J\in[I]^{<\kappa}$. Then $S((X_I)_\kappa)\le(\alpha^{<\kappa})^+$. \end{corollary} \begin{proof} The cardinal $\kappa$ is singular by Theorem~\ref{weak<kappa}(b), hence is a limit cardinal. Then by Lemma~\ref{newl} (with $\alpha$ there replaced by $(\alpha^{<\kappa})^+)$ we have $S((X_J)_\lambda)<(\alpha^{<\kappa})^+$ whenever $\lambda<\kappa$ and $J\in[I]^{<\lambda}$. Then $S((X_I)_\kappa)\leq(\alpha^{<\kappa})^+$ by Theorem~\ref{T4.16}. \end{proof} Theorem~\ref{GCNab}, like Corollary~\ref{S(power)}, is a miscellaneous stand-alone result based on the homeomorphisms developed in Lemma~\ref{homeos}. To see that those results are (consistently) nonvacuous, we need a model of ZFC where $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$ and $\alpha^\kappa>(\alpha^{<\kappa})^+$. For that, see Remark~\ref{nonvac2}. In \ref{homeos}--\ref{S(power)}, given a set $\{X_i:i\in I\}$ of spaces, for $i\in I$ we write $$\widetilde{i}:=\{j\in I:X_i=_h X_j\}.$$ We note that if $\kappa\geq\omega$ and $|\widetilde{i}|\geq\cf(\kappa)$ for each $i\in I$, then there is a partition $\{I(\eta):\eta<\cf(\kappa)\}$ of $I$ such that $|I(\eta)\cap\widetilde{i}|=|\widetilde{i}|$ for each $i\in I$. Indeed, it is enough for each $i\in I$ to choose a partition $\{A(\widetilde{i},\eta):\eta<\cf(\kappa)\}$ of $\widetilde{i}$ with each $|A(\widetilde{i},\eta)|=|\widetilde{i}|$ and to take $I(\eta):=\bigcup_{i\in I}\,A(\widetilde{i},\eta)$. \begin{lemma}\label{homeos} Let $\kappa\geq\omega$ and $\cf(\kappa)\le \lambda\le \kappa$ with $\lambda$ regular, and let $\{X_i:i\in I\}$ be a set of spaces with each $|\widetilde{i}|\geq\cf(\kappa)$. Let $\{I(\eta):\eta<\cf(\kappa)\}$ be a partition of $I$ such that $|I(\eta)\cap\widetilde{i}|=|\widetilde{i}|$ for each $i\in I$. Then {\rm (a)} $(X_I)_\lambda=_h(X_{I(\eta)})_\lambda$ for each $\eta<\cf(\kappa)$; {\rm (b)} $(X_I)_\lambda=_h(\Pi_{\eta<\cf(\kappa)}\,(X_{I(\eta)})_\lambda)_\lambda$; and {\rm (c)} $(X_I)_\lambda=_h(((X_I)_\lambda)^{\cf(\kappa)})_\lambda$. \end{lemma} \begin{proof} (a) Given $\eta<\cf(\kappa)$, let $\phi:I\twoheadrightarrow I(\eta)$ be a bijection such that $\phi[\widetilde{i}]=\widetilde{i}\cap I(\eta)$ for each $i\in I$. Then the map $\Phi:X_I\twoheadrightarrow X_{I(\eta)}$ given by $\Phi(x_i)=x_{\phi(i)}\in X_{I(\eta)}$ is a homeomorphism, with $\phi[R(A)]=R(\Phi[A])$ for each generalized rectangle $A=\Pi_{i\in I}\,A_i\subseteq X_I$. (b) We show that the natural map from $\Pi_{\eta<\cf(\kappa)}\,X_{I(\eta)}$ onto $X_I$ is a homeomorphism from $(\Pi_{\eta<\cf(\kappa)}\,(X_{I(\eta)})_\lambda)_\lambda$ onto $(X_I)_\lambda$ when $\lambda$ is regular. Indeed, an (open) generalized rectangle $U=\Pi_{\eta<\cf(\kappa)}\,U(\eta)$ in $\Pi_{\eta<\cf(\kappa)}\,X_{I(\eta)}$ with $U(\eta)=\Pi_{i\in I(\eta)}\,U(\eta,i)$ satisfies $R(U)=\bigcup_{\eta<\cf(\kappa)}\,R(U(\eta))$, so $|R(U)|<\lambda$ if and only if each $|R(U(\eta))|<\lambda$. (c) follows immediately from (a) and (b). \end{proof} \begin{theorem}\label{GCNab} Let $\alpha\geq2$ and $\kappa\ge\omega$ be cardinals and let $\{X_i:i\in I\}$ be a set of nonempty spaces. Suppose that $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$ and that $|\widetilde{i}|\geq\cf(\kappa)$ for each $i\in I$. If $S((X_I)_\kappa)>(\alpha^{<\kappa})^+$ then $S((X_I)_\kappa)>\alpha^\kappa$. \end{theorem} \begin{proof} By Theorem~\ref{T4.16} there are $J\in[I]^{<\kappa}$ and a regular cardinal $\lambda<\kappa$ such that $S((X_J)_\lambda)\geq(\alpha^{<\kappa})^+$. Let $\{I(\eta):\eta<\cf(\kappa)\}$ be a partition of $I$ as in Lemma~\ref{homeos}, and for $\eta<\cf(\kappa)$ let $\sC(\eta)$ be a cellular family in $X_{I(\eta)}$ such that $|\sC(\eta)|=\alpha^{<\kappa}$. Then $$\sC:=\Pi_{\eta<\cf(\kappa)}\,\sC(\eta)=\{\Pi_{\eta<\cf(\kappa)}\,C(\eta):C(\eta)\in\sC(\eta)\}$$ is cellular in $(\Pi_{\eta<\cf(\kappa)}\,(X_{I(\eta)})_\lambda)_\lambda=_h(X_I)_\lambda$, so $$S((X_I)_\kappa)\geq S((X_I)_\lambda)>|\sC|=(\alpha^{<\kappa})^{\cf(\kappa)}=\alpha^\kappa.$$ \vskip-18pt \end{proof} \begin{corollary}\label{S(power)} Let $\alpha\geq2$ and $\kappa\geq\omega$ be cardinals, and let $X$ be a space and $I$ a set. {\rm (a)} Suppose that $\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$. Then $S((X^I)_\kappa)\leq(\alpha^{<\kappa})^+$ if and only if $S((X^J)_\kappa)\leq(\alpha^{<\kappa})^+$ for every nonempty $J\in[I]^{<\kappa}$. {\rm (b)} Suppose that $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$. If $S((X^I)_\kappa)>(\alpha^{<\kappa})^+$ then \begin{itemize} \item[{\rm (1)}] there is nonempty $J\in[I]^{<\kappa}$ such that $S((X^J)_\kappa)\geq(\alpha^{<\kappa})^+$; and \item[{\rm (2)}] if $|I|\geq\cf(\kappa)$, then $S((X^I)_\kappa)>\alpha^\kappa$. \end{itemize} \end{corollary} \begin{proof} In view of Theorem~\ref{GCN} and Theorem~\ref{T4.16}, only (b)(2) requires attention. This follows from Theorem~\ref{GCNab}, since now $\widetilde{i}=I$ for each $i\in I$. \end{proof} \begin{remarks}\label{nonvac2} {\rm (a) It is easy to see that in many models of ZFC, for example under GCH, the equality $\alpha^\kappa=(\alpha^{<\kappa})^+$ holds for all cardinals $\alpha$ and $\kappa$ for which $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$. In such models, Theorem~\ref{GCNab} and Corollary~\ref{S(power)}(b)(2) become tautologies. To see that Theorem~\ref{GCNab} and Corollary~\ref{S(power)}(b)(2) are not vacuous in every setting, it is enough to refer to the models $\VV_1$ and $\VV_2$ of Gitik and Shelah described in Discussion~\ref{gitik-shelah}(d), taking now $\alpha=2$ and $\kappa=\aleph_\omega$. In those models we have $$\alpha^{<\kappa}=2^{<\aleph_\omega}=\aleph_\omega,$$ while (using Theorem~\ref{expon}(c), for example) $$\alpha^\kappa=(\alpha^{<\kappa})^{<\kappa} =2^{\aleph_\omega}=\aleph_{\omega+2}>\aleph_{\omega+1}=(\alpha^{<\kappa})^+.$$ (b) It is a consequence of Theorem~\ref{GCNab} and Corollary~\ref{S(power)}(b)(2) that under the hypotheses there the relation $S((X_I)_\kappa)=\alpha^\kappa$ is impossible (even when $\alpha^\kappa$ is regular). } \end{remarks} Now we consider two questions. The first of these arises naturally from Corollary~\ref{GCN}(b) and Corollary \ref{GCNan}, and a version of the second, attributed to Argyros and Negrepontis, appears in \cite{comfneg82}. Theorem~\ref{GCH} shows a relation between these. For what we do and do not not know about the status of these questions in ZFC and in augmented systems, see Remarks \ref{R4.50}((a) and (b)). \begin{question}\label{Q1} {\rm Let $\alpha\geq2$, $\kappa\geq\omega$, and let $\{X_i:i\in I\}$ be a set of spaces such that $S((X_J)_\kappa)\leq(\alpha^{<\kappa})^+$ for each nonempty $J\in[I]^{<\kappa}$. Is then necessarily $S((X_I)_\kappa)\leq(\alpha^{<\kappa})^+$? } \end{question} \begin{question}[{\cite[7.15(a)]{comfneg82}}]\label{Q} {\rm Are there spaces $X$ and $Y$ with $S(X\times Y)>S(X)>S(Y)$? } \end{question} \begin{remark}\label{Qa} {\rm By Theorem~\ref{GCN}(b), the answer to Question~\ref{Q1} is affirmative in case $\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$. } \end{remark} \begin{theorem}\label{GCH} Let $\alpha\geq2$, $\kappa\geq\omega$, and $\{X_i:i\in I\}$ witness a negative answer to Question \ref{Q1}. If $\alpha<\kappa$ then the answer to Question~\ref{Q} is positive. \end{theorem} \begin{proof} Set \begin{center} $L:=\{i\in I:S(X_i)=(\alpha^{<\kappa})^+\}$ and $M:=\{i\in I:S(X_i)<(\alpha^{<\kappa})^+\}$. \end{center} Note first from Remark \ref{Qa} that $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$, so $\kappa>\omega$, and $\kappa$ and $\alpha^{<\kappa}$ are singular by Theorem~\ref{weak<kappa}. Further, it follows directly from our hypothesis that $|I|\ge\kappa$. Clearly we may assume without loss of generality that $S(X_i)\ge 3$ for every $i\in I$. For each infinite cardinal $\lambda$ we have $S((X_I)_{\lambda^+})=S((X_L)_{\lambda^+}\times(X_M)_{\lambda^+})$. Thus to prove the theorem it suffices to show that there exists an infinite cardinal $\lambda<\kappa$ such that \begin{itemize} \item[(i)] $S((X_L)_{\lambda^+})=(\alpha^{<\kappa})^+$, \item[(ii)] $S((X_M)_{\lambda^+})<\alpha^{<\kappa}$, and \item[(iii)] $S((X_I)_{\lambda^+})>(\alpha^{<\kappa})^+.$ \end{itemize} Let $\{\kappa_\eta:\eta<\cf(\kappa)\}$ be a family of cardinals as in Notation \ref{defkappaeta}. We note that \begin{equation}\label{Eq15} |L|<\cf(\kappa), \end{equation} and \begin{equation}\label{Eq16} \mbox{there is }\eta<\cf(\kappa)\mbox{ such that } S(X_i)\leq\alpha^{\kappa_\eta} \mbox{ for each }i\in M. \end{equation} \noindent (Indeed if (\ref{Eq15}) [resp., (\ref{Eq16})] fails then by Theorem~\ref{largeS}(b) there is $J\in[L]^{\cf(\kappa)}$ [resp., $J\in[M]^{\cf(\kappa)}$] such that $$S((X_J)_\kappa)\geq S((X_J)_{(\cf(\kappa))^+})>(\alpha^{<\kappa})^+,$$ a contradiction since $|J|=\cf(\kappa)<\kappa$.) It follows from (\ref{Eq15}) that $|M|=|I|\ge \kappa$; further, according to Lemma \ref{obs1} (with $M$, $\gamma^+$, $2$ and $(\alpha^{<\kappa})^+$ in place of $I$, $\kappa$, $\beta$ and $\alpha$, respectively), we have \begin{equation}\label{Eq14} S((X_M)_{\gamma^+})> 2^\gamma \mbox{ for every infinite }\gamma<\kappa. \end{equation} We claim that \begin{equation}\label{Eq17} S((X_L)_\kappa)=S((X_M)_\kappa)=(\alpha^{<\kappa})^+. \end{equation} To see that, fix $\eta$ as in (\ref{Eq16}) and let $\gamma$ be such that $\kappa_\eta<\gamma<\kappa$. Since $\alpha<\kappa$, we have $(\alpha^\gamma)^+<\alpha^{<\kappa}$; then \begin{equation}\label{Eq18} S((X_M)_{\gamma^+})\leq(((\alpha^{\kappa_\eta})^\gamma)^\gamma)^+= (\alpha^\gamma)^+<\alpha^{<\kappa} \end{equation} by (\ref{Eq16}) and Theorem~\ref{GKurepa}(b) (with $\alpha$, $\kappa$ and $I$ replaced by $\alpha^{\kappa_\eta}$, $\gamma^+$ and $M$, respectively). It then follows from (\ref{Eq14}) and (\ref{Eq18}) that $$\sup\{S((X_M)_\gamma):\gamma<\kappa\}=\alpha^{<\kappa},$$ hence from Theorem \ref{Obsn}(c) we have $$S((X_M)_{\kappa})=(\alpha^{<\kappa})^+.$$ Since $S((X_M)_\kappa)=(\alpha^{<\kappa})^+<S((X_I)_\kappa)$ we have $M\neq I$, so $L\neq\emptyset$. Clearly then $S((X_L)_\kappa)\geq(\alpha^{<\kappa})^+$, while $S((X_L)_\kappa)\leq(\alpha^{<\kappa})^+$ follows from (\ref{Eq15}) and our hypothesis. Therefore $S((X_L)_\kappa)=(\alpha^{<\kappa})^+$ and claim (\ref{Eq17}) is proved. From (\ref{Eq17}) we have $$(\alpha^{<\kappa})^+\leq S((X_L)_{\lambda^+})\leq S((X_L)_\kappa)=(\alpha^{<\kappa})^+$$ for each infinite $\lambda$, so (i) holds (for all infinite $\lambda$). That (ii) holds for all $\lambda$ such that $\kappa_\eta<\lambda<\kappa$ is given by (\ref{Eq18}). It follows that there is $\lambda$ such that $\kappa_\eta<\lambda<\kappa$ and $S((X_I)_{\lambda^+})>(\alpha^{<\kappa})^+$, since otherwise we have $$\sup\{S((X_I)_{\lambda^+}):\lambda<\kappa\}=\sup\{S((X_I)_\lambda):\lambda<\kappa\}=(\alpha^{<\kappa})^+$$ and Theorem \ref{Obsn}(b) gives the contradiction $S((X_I)_\kappa)=(\alpha^{<\kappa})^+$. Then (iii) holds for that specific $\lambda$. \end{proof} \begin{corollary}\label{CGCH} Let $\MM$ be a model of ZFC in which every singular cardinal is a strong limit cardinal (e.g. $\MM$ is a model of ZFC+GCH). If the answer to Question \ref{Q1} is negative in $\MM$ then the answer to Question~\ref{Q} is positive in $\MM$. \end{corollary} \begin{proof} With $\alpha$, $\kappa$ and $\{X_i:i\in I\}$ chosen as in Theorem \ref{GCH} it suffices to show that $\alpha<\kappa$. Since $\alpha^{<\kappa}<(\alpha^{<\kappa})^{<\kappa}$ by Remark \ref{Qa}, by Theorem \ref{weak<kappa}(b), both $\kappa$ and $\alpha^{<\kappa}$ are singular. If $\alpha=\alpha^{<\kappa}$ then $\alpha=\alpha^{<\kappa}=(\alpha^{<\kappa})^{<\kappa}$, a contradiction; therefore $\alpha<\alpha^{<\kappa}$. Suppose now that $\kappa\leq\alpha$. Then from Theorem~\ref{expon}(c) we have $(\alpha^{<\kappa})^{<\kappa}=\alpha^\kappa\le\alpha^\alpha=2^\alpha$, and the relation $\alpha<\alpha^{<\kappa}<2^\alpha$ contradicts the hypothesis that $\alpha^{<\kappa}$ is a strong limit cardinal. \end{proof} \begin{remarks}\label{R4.50} {\rm (a) The proof of the previous corollary does not need the full hypothesis that every singular cardinal in $\MM$ is strong limit. It is enough to know just that $\alpha^{<\kappa}$ is strong limit. (b) ZFC-consistent examples of spaces as requested in Question~\ref{Q} are available in the literature. ~~(1) In the Cohen models of Fleissner \cite{fleissner78} (see Remark~\ref{G-Smodels}(b)) there are spaces $X$ and $Y$ such that $$S(Y\times Y)=\aleph_{\omega+2}>\cc=\aleph_{\omega+1}>\aleph_1=S(Y),$$ and then with $X$ the ``disjoint union" of $D(\aleph_1)$ and $Y$ we have $S(X\times Y)>S(X)>S(Y)$. ~~(2) It is shown by Shelah \cite[4.4]{shelah} that if $\kappa$ is a singular strong limit cardinal such that $\lambda:=\kappa^+=2^\kappa$ then there are spaces $X$ and $Y$ such that $$S(X\times Y)\ge \lambda^{++}>\lambda^+=S(X)>\lambda>\kappa>(2^{\cf(\kappa)})^{++}\ge S(Y).$$ (c) We do not know if there are models of ZFC in which no spaces as in Question~\ref{Q} exist. We do not know if the answer to Question~\ref{Q1} is absolutely or consistently ``Yes", absolutely or consistently ``No". } \end{remarks}
\section{Introduction and basic def\/initions} \label{sec1} Let~$u$ be a~function of integers~$i$ and~$j$. By $u_{p,q}$ denote the shifted value $u(i+p,j+q)$ of this function (in particular, $u:=u_{0,0}=u(i,j)$). Let $F(u,u_{1,0},u_{0,1})$ be a~single-valued function. Integrable (in various senses) equations of the form \begin{gather} \label{uij} u_{1,1}=F(u,u_{1,0},u_{0,1}) \end{gather} are actively studied in recent years (see, e.g., \cite{ABS,LS,LY,Mikh,R2001} and references within). In the present article we mostly consider so-called {\it Darboux integrable} equations, i.e.~equations~\eqref{uij} for which there exist functions~$I$ and~$J$ such that the relations \begin{gather} \label{iib} I(u_{0,1},u_{1,1},\dots,u_{m,1})=I(u,u_{1,0},\dots,u_{m,0}), \\ \label{ijb} J(u_{1,0},u_{1,1},\dots,u_{1,n})=J(u,u_{0,1},\dots,u_{0,n}) \end{gather} hold true for any solution~$u$ of the equation. (In other words, relationships~\eqref{iib},~\eqref{ijb} mean that~$I$ and~$J$ remain unchanged after the shifts in~$j$ and~$i$, respectively.) The functions~$I$ and~$J$ are respectively called an~$i$-integral of order~$m$ and a~$j$-integral of order~$n$. The discrete wave equation $u_{1,1}=u_{1,0}+u_{0,1}-u$ gives us the simplest example of a~Darboux integrable equation. Here $I=u_{1,0}-u$ and $J=u_{0,1}-u$. It should be noted that this work (in contrast to, for instance,~\cite{GY}) deals only with the autonomous integrals, i.e.~with the integrals that do not depend explicitly on the discrete variables~$i$ and~$j$. The equations~\eqref{uij} can be considered as dif\/ference analogues of the partial dif\/ferential equations \begin{gather} u_{xy}=F(u,u_x,u_y). \label{hyp} \end{gather} The concept of the Darboux integrability was initially introduced for partial dif\/ferential equations back in the 19th century. (The more recent term~$C$-integrability, which was of\/fered in~\cite{Cal}, in some sense generalizes this concept.) Searching for Darboux integrable equations of the form~\eqref{hyp} was started in classical works such as~\cite{Guq}, and the most recent and complete classif\/ication result was obtained in~\cite{ZhSok} more than a~century later. At present, similar classif\/ication results are absent for Darboux integrable equations~\eqref{uij}, only separate examples (see, for instance,~\cite{GY,HZS2,Sit}) and the description~\cite{Stn} of a~special case with $n=1$ in~\eqref{ijb} are known. Therefore, it seems reasonable to consider a~subclass of Darboux integrable equations~\eqref{uij} as an intermediate goal. To def\/ine this subclass, in the present paper we assume that the right-hand side of~\eqref{uij} satisf\/ies the inequalities \begin{gather} \label{hipc} \frac{\partial F}{\partial u} \ne 0, \qquad \frac{\partial F}{\partial u_{1,0}} \ne 0, \qquad \frac{\partial F}{\partial u_{0,1}} \ne 0 \end{gather} and a~relationship of the form \begin{gather} \label{qua} \varphi\left(u_{1,0},F(u,u_{1,0},u_{0,1})\right)=\psi(u,u_{0,1}), \end{gather} where $\varphi(w,z)$ and $\psi(w,z)$ are functionally independent. It is easy to see that the left-hand side of~\eqref{qua} is the shift of $\varphi(u,u_{0,1})$ in~$i$ by virtue of~\eqref{uij}. If we denote the operator of this shift by~$T_{\rm i}$, then~\eqref{qua} reads as \begin{gather} \label{invt} T_{\rm i}(\varphi (u, u_{0,1})) = \psi(u, u_{0,1}), \qquad \frac{\partial \varphi}{\partial u} \frac{\partial \psi}{\partial u_{0,1}} - \frac{\partial \varphi}{\partial u_{0,1}} \frac{\partial \psi}{\partial u} \ne 0. \end{gather} Thus, we consider formal dif\/ference analogues of the quasilinear partial dif\/ferential equations $u_{xy}=a(u,u_y) u_x + b(u,u_y)$ (because any of these dif\/ferential equations can be represented as $D_{x}(\varphi(u,u_y))=\psi(u,u_y)$ in terms of the total derivative $D_x$). For shorter formulations, below we refer to equation~\eqref{invt} for designating an equation of the form~\eqref{uij} that satisf\/ies~\eqref{qua}. In a~part of the article we use the stronger assumptions that the map $(w,z)\!\rightarrow\! (\varphi(w,z),\psi(w,z))$ is injective, the integral~$J$ has second order and $\varphi(u,u_{0,1})$ is uniquely expressed in terms of~$J$ and $\varphi(u_{0,1},u_{0,2})$. (Since we prove below that~$j$-integrals are expressed in terms of $\varphi(u,u_{0,1})$ and its shifts in~$j$, the last assumption is not so restrictive as it seems; moreover, it can be omitted if a~certain general statement will be proved~-- see the last paragraph of this section for more details.) It should be noted that one of the most known integrable equations on the quad-graph, the discrete Liouville equation \begin{gather} \label{dle} u_{1,1}=\frac{(u_{1,0}-1)(u_{0,1}-1)}{u} \end{gather} from~\cite{Hirota}, satisf\/ies all of the above assumptions. Indeed, the work~\cite{AdS} demonstrates that \begin{gather*} I=\left(\frac{u_{2,0}}{u_{1,0}-1} +1 \right) \left(\frac{u-1}{u_{1,0}} + 1 \right), \qquad J=\left(\frac{u_{0,2}}{u_{0,1}-1} +1 \right) \left(\frac{u-1}{u_{0,1}} + 1 \right) \end{gather*} for~\eqref{dle}, the corresponding functions $\varphi(w,z)=z/(w-1)$, $\psi(w,z)=(z-1)/w$ def\/ine the injective map and $\varphi(u,u_{0,1})=(\varphi(u_{0,1},u_{0,2})+1)/(J - \varphi(u_{0,1},u_{0,2}) - 1)$. Thus, the present article is devoted to `nearest relatives' of the discrete Liouville equation~\eqref{dle}. More precisely, we focus on interactions of a~non-point invertible transformation~\cite{Sit} with Darboux integrability and, to a~lesser extent, higher symmetries of the quad-graph equations. This transformation is def\/ined in the following way. We can rewrite~\eqref{invt} in the form of the system \begin{gather} \label{absd} v=\varphi (u,u_{0,1}), \qquad v_{1,0}:=T_{\rm i}(v)=\psi (u,u_{0,1}), \end{gather} express~$u$ and $u_{0,1}$ in terms of~$v$, $v_{1,0}$ from~\eqref{absd} and obtain \begin{gather} \label{pqsd} u=\Omega(v,v_{1,0}), \qquad u_{0,1}=\Upsilon(v,v_{1,0}), \end{gather} where~$\Omega$ and~$\Upsilon$ are functionally independent. According to~\eqref{hipc}, the functions~$\varphi$ and~$\psi$ are assumed here to be essentially depending on both their arguments, and~\eqref{pqsd} therefore implies that~$\Omega$ and~$\Upsilon$ essentially depend on both their arguments too. The system~\eqref{pqsd} is equivalent to the equation \begin{gather} \label{pqd} \Omega(v_{0,1},v_{1,1})=\Upsilon(v,v_{1,0}). \end{gather} Generally speaking, the above procedure is well-def\/ined only locally and the right-hand sides of~\eqref{pqsd} may be dif\/ferent for dif\/ferent pairs $(u,u_{0,1})$. We avoid this if no more than one pair $(u,u_{0,1})$ satisf\/ies the system~\eqref{absd} for any given~$v$, $v_{1,0}$. Under this assumption, the equation~\eqref{pqd} is well-def\/ined and the transformation $v=\varphi (u,u_{0,1})$ maps all solutions of~\eqref{invt} into solutions of~\eqref{pqd}. Because the set of solutions to~\eqref{pqd} may be wider than the image of solutions to~\eqref{invt} under the transformation $v=\varphi (u,u_{0,1})$, we can not guarantee that the inverse transformation $u=\Omega(v,v_{1,0})$ maps any solution of~\eqref{pqd} into a~solution of~\eqref{invt} (some pairs $(v,v_{1,0})$ may generate another equations of the form~\eqref{invt} when we perform the above procedure in the inverse order). But we can restore all equations related to~\eqref{pqd} by using the inverse procedure. The described transformation is the direct analogue of that was of\/fered in~\cite{Yam90} for dif\/ferential-dif\/ference equations. Some applications of the transformation~\eqref{absd}--\eqref{pqd} can also be found in~\cite{Sit,Yam94}. It is almost obvious that this transformation preserves Darboux integrability. But a~formal proof of this fact is still needed to demonstrate, for example, that no~$j$-integral of~\eqref{invt} becomes constant after substituting $u=\Omega(v,v_{1,0})$ into it. Such a~proof is given in Section~\ref{s2}. More precisely, we prove that equation~\eqref{pqd} has a~$j$-integral of order $n-1$ and an~$i$-integral of order $m+1$ if equation~\eqref{invt} possesses~$j$- and~$i$-integrals of orders~$n$ and~$m$, respectively. This reduces the classif\/ication problem for Darboux integrable equations~\eqref{invt} admitting~$n$-th order~$j$-integrals to the classif\/ication of equations~\eqref{pqd} possessing~$j$-integrals of order $n-1$. As an example of such kind, in Section~\ref{s2} we completely describe the Darboux integrable equations~\eqref{invt} admitting a~second-order~$j$-integral such that $\varphi(u,u_{0,1})$ is uniquely expressed in terms of this integral and $\varphi(u_{0,1},u_{0,2})$. In Section~\ref{s3} we study the interaction between the transformation~\eqref{absd}--\eqref{pqd} and symmetries of the quad-graph equations. Analogically to the case of semi-discrete equations~\cite{Yam90}, the transformation $v=\varphi (u,u_{0,1})$ def\/ines a~dif\/ference substitution for special symmetries of~\eqref{invt} and, under additional assumptions, maps any higher symmetry of~\eqref{invt} into a~higher symmetry of~\eqref{pqd} (i.e.~the transformation preserves integrability in the sense of the symmetry test\footnote{This test is described, for example, in~\cite{GY,LY}.}). The former fact allows us to obtain necessary conditions of Darboux integrability for the equations of the form~\eqref{invt} if~$u$ is uniquely expressed in terms of $\varphi (u,u_{0,1})$ and $u_{0,1}$. Now, let us introduce notation and more formal def\/initions. Due to the conditions~\eqref{hipc}, we can express any argument of the right-hand side~$F$ of~\eqref{uij} in terms of the others and rewrite this equation, after appropriate shifts in~$i$ and~$j$, in any of the following forms \begin{gather} \label{umm} u_{-1,-1}=\overline{F}(u,u_{-1,0},u_{0,-1}), \\ \label{upm} u_{1,-1}=\hat{F}(u,u_{1,0},u_{0,-1}), \\ \label{ump} u_{-1,1}=\tilde{F}(u,u_{-1,0},u_{0,1}). \end{gather} These formulas (and their consequences derived by shifts in~$i$ and~$j$) allow us to express any `mixed shift' $u_{p,q}$, $p q \ne 0$, in terms of $u_{k,0}$, $u_{0,l}$ at least locally (i.e.~in an enough small neighborhood of any arbitrary selected solution of~\eqref{uij}, which is considered in a~f\/inite number of the points $(i, j)$). Therefore, we can formulate our reasonings only in terms of an arbitrary solution~$u$ to~\eqref{uij} and its `canonical shifts' $u_{k,0}$, $u_{0,l}$, $k,l \in \mathbb{Z}$. These `canonical shifts' are called {\it dynamical variables} and can be considered as functionally independent. (The mixed shift elimination procedure and the dynamical variables are described in more details, for example, in~\cite{Mikh}.) We use the notation $g[u]$ to designate that the function~$g$ depends on a~f\/inite number of the dynamical variables. All functions are assumed to be analytical in this paper, and our considerations are local. In general, the right-hand sides of~\eqref{umm}--\eqref{ump} are not uniquely def\/ined (may vary with~$i$,~$j$ and with a~solution in a~neighborhood of which we consider these functions). This does not matter if we use~\eqref{umm}--\eqref{ump} to only estimate what dynamical variables do local expressions for $u_{p,q}$ depend on. But this is important in certain cases, and our statements therefore contain the single-valuedness assumptions for $\hat{F}$ when it is needed (to avoid dif\/f\/iculties discussed, for example, in~\cite{NeA}). Let $T_{\rm i}$ and $T_{\rm j}$ denote the operators of the forward shifts in~$i$ and~$j$ by virtue of the equation~\eqref{uij}. For any function~$f$, they satisfy the rules \begin{gather*} T_{\rm i}(f(a,b,c,\dots))=f(T_{\rm i}(a),T_{\rm i}(b),T_{\rm i}(c),\dots), \\ T_{\rm j} (f(a,b,c,\dots))=f(T_{\rm j} (a),T_{\rm j} (b),T_{\rm j} (c),\dots). \end{gather*} In addition, $T_{\rm i}$ and $T_{\rm j}$ map $u_{p,q}$ into $u_{p+1,q}$ and $u_{p,q+1}$, respectively, and then replace any mixed shift of~$u$ with its expression in terms of the dynamical variables. For example, \begin{gather} \label{u1n} T_{\rm i}(u_{0,n})=T_{\rm j}^{n-1}(F), \qquad T_{\rm i}(u_{0,-n})=T_{\rm j}^{1-n}(\hat{F}), \qquad n \in \mathbb{N}. \end{gather} Here a~shift operator with a~superscript~$k$ designates the~$k$-fold application of this operator (e.g. $T_{\rm j}^3:= T_{\rm j} \circ T_{\rm j} \circ T_{\rm j}$, $T_{\rm i}^{-2}:= T_{\rm i}^{-1} \circ T_{\rm i}^{-1}$ and any operator with the zero superscript is the identity mapping). The inverse (backward) shift operators $T_{\rm i}^{-1}$ and $T_{\rm j}^{-1}$ are def\/ined in a~similar way. \begin{definition} \label{def1} Let functions $I[u]$ and $J[u]$ satisfy the relations $T_{\rm j}(I)=I$ and $T_{\rm i}(J)=J$ for an equation of the form~\eqref{uij}, and let each of the functions essentially depend on at least one of the dynamical variables. Then $I[u]$ and $J[u]$ are respectively called an~$i$-integral and a~$j$-integral of the equation~\eqref{uij}, and this equation is called Darboux integrable. \end{definition} It is easy to prove that the~$i$- and~$j$-integrals have the form $I(u_{k,0},u_{k+1,0},\dots,u_{k+m,0})$ and $J(u_{0,l},u_{0,l+1},\dots,u_{0,l+n})$, respectively (see, for example,~\cite{Stn}). The numbers~$m$ and~$n$ are called {\it order} of the corresponding integral if $I_{u_{k,0}} I_{u_{k+m,0}} \ne 0$ and $J_{u_{0,l}} J_{u_{0,l+n}} \ne 0$. We can set $k=l=0$ without loss of generality because $T^{-k}_{\rm i}$ and $T^{-l}_{\rm j}$ respectively map any~$i$- and~$j$-integrals into~$i$- and~$j$-integrals again. Thus, equations~\eqref{umm}--\eqref{ump} are in fact not needed for the above def\/inition. \begin{definition} An equation $u_{t} = f[u]$ is called a~symmetry of equation~\eqref{uij} if the relation $L(f)=0$ holds true, where \begin{gather*} L = T_{\rm i} T_{\rm j} - \frac{\partial F}{\partial u_{1,0}} T_{\rm i} - \frac{\partial F}{\partial u_{0,1}} T_{\rm j} - \frac{\partial F}{\partial u}. \end{gather*} \end{definition} According to~\cite{AdS,Stn}, if equation~\eqref{uij} is Darboux integrable and uniquely solvable for $u_{1,0}$ (i.e.~the right-hand side $\hat{F}$ of~\eqref{upm} is uniquely def\/ined), and $J[u]$ denotes its~$j$-integral, then there exists an operator \begin{gather} \label{rop} R=\sum\limits_{q=0}^r \lambda_q (u_{0,\varrho},u_{0,\varrho +1},\dots,u_{0,s}) T^q_{\rm j}, \qquad \lambda_r \ne 0, \end{gather} such that \begin{gather} \label{ed} u_t=R\big(\eta(T_{\rm j}^{p}(J), T_{\rm j}^{p+1}(J), T_{\rm j}^{p+2}(J), \dots)\big) \end{gather} is a~symmetry of this equation for any integer~$p$ and any function~$\eta$ depending on a~f\/inite number of the arguments. Both the present paper and the article~\cite{Stn} (results of which we use below) in fact describe equations that admit~$j$-integrals and symmetries of the form~\eqref{ed} (the existence of~$i$-integrals is not really used in the most part of the reasonings). We therefore make additional assumptions on a~$j$-integral in Proposition~\ref{d2p} and $\hat{F}$ in Corollaries~\ref{dc1},~\ref{dc2} to guarantee that the transformed and the original equations are uniquely solvable for $v_{1,0}$ and $u_{1,0}$, respectively. Proposition~\ref{d2p} and Corollaries~\ref{dc1},~\ref{dc2} will remain valid without these assumptions if the existence of symmetries~\eqref{ed} for the Darboux integrable equations is proved without employing the single-valuedness of $\hat{F}$. \section{Transformation of integrals} \label{s2} It is convenient for further reasoning to prove the following proposition f\/irst. \begin{lemma} \label{l0} All functions in the set $\{T_{\rm i}^p(\varphi), T_{\rm j}^q(\varphi) \,|\, p,q \in \mathbb{Z}, q \ne 0 \}$ are functionally independent for any equation of the form~\eqref{invt}. \end{lemma} \begin{proof} The function~$\psi$ can be rewritten as $\eta_1(\varphi,u)$, where $\eta_1$ must depend on its second argument due to the functional independence of~$\varphi$ and~$\psi$. Using the formula $T_{\rm i}(\varphi)=\psi=\eta_1(\varphi,u)$ and induction on~$p$, we obtain that $T_{\rm i}^p(\varphi)=\eta_p(\varphi,u,u_{1,0},\dots,u_{p-1,0})$ depends on $u_{p-1,0}$ for any $p>1$. Also,~$\varphi$ can be represented as a~function $\eta_{-1}(\psi,u)$ that depends on its second argument. Therefore, $T_{\rm i}^{-1}(\varphi)=\eta_{-1}(\varphi,u_{-1,0})$ and $T_{\rm i}^p(\varphi)=\eta_p(\varphi,u_{-1,0},\dots,u_{p,0})$ depends on $u_{p,0}$ for $p<0$. Thus, the functions $T_{\rm i}^p(\varphi)$ are functionally independent because~$\varphi$ and $T_{\rm i}(\varphi)=\psi$ are functionally independent and any other $T_{\rm i}^p(\varphi)$ depends on the variable that is absent in either all previous (if $p>1$) or all next (if $p<0$) members of the sequence $T_{\rm i}^s(\varphi)$. If $q>0$, then the function $T_{\rm j}^q(\varphi)=\varphi(u_{0,q},u_{0,q+1})$ can not be expressed in terms of $T_{\rm i}^p(\varphi)$, $p \in \mathbb{Z}$, and $T_{\rm j}^s(\varphi)$, $s<q$, because they do not depend on $u_{0,q+1}$. If $q<0$, then the function $T_{\rm j}^q(\varphi)=\varphi(u_{0,q},u_{0,q+1})$ can not be expressed in terms of $T_{\rm i}^p(\varphi)$, $p \in \mathbb{Z}$, and $T_{\rm j}^s(\varphi)$, $s>q$, because they do not depend on $u_{0,q}$. Thus, $\{T_{\rm i}^p(\varphi), T_{\rm j}^q(\varphi) \,|\, p,q \in \mathbb{Z}, q \ne 0 \}$ is a~set of functionally independent functions. \end{proof} \begin{lemma} \label{intf} Up to shifts in~$j$, any~$n$-th order~$j$-integral of equation~\eqref{invt} can be represented in the form $\Phi(\varphi(u,u_{0,1}),\varphi(u_{0,1},u_{0,2}),\dots,\varphi(u_{0,n-1},u_{0,n}))$. This representation is well-defined on solutions of~\eqref{uij}, and the function~$\Phi$ essentially depends on its first and last arguments. \end{lemma} \begin{proof} As it is demonstrated in the comments to Def\/inition~\ref{def1}, we can assume without loss of generality that the~$j$-integral has the form $J(u,u_{0,1},\dots,u_{0,n})$. Equation~\eqref{qua} implies that the right-hand side of~\eqref{uij} has the form $g(\psi(u,u_{0,1}),u_{1,0})$, where~$g$ is single-valued. The backward shift in~$i$ gives us $u_{0,1}=g(\varphi(u,u_{0,1}),u)$. Using this expression and its consequences derived by shifts in~$j$, we rewrite the~$j$-integral as $J=\Phi(u,\varphi(u,u_{0,1}),\varphi(u_{0,1},u_{0,2}),\dots,\varphi(u_{0,n-1},u_{0,n}))$. Since \begin{gather*} T_{\rm i} \big(\Phi(u, \varphi,\dots,T_{\rm j}^{n-1}(\varphi))\big) = \Phi\big(u_{1,0}, \psi,\dots,T_{\rm j}^{n-1}(\psi)\big) \end{gather*} by virtue of~\eqref{invt}, the relation $T_{\rm i}(\Phi)=\Phi$ can hold true only if~$\Phi$ does not depend on its f\/irst argument ($\Phi_u=0$). The function~$\Phi$ also must depends on~$\varphi$ and $T_{\rm j}^{n-1}(\varphi)$ because $J_{u}=0$ and $J_{u_{0,n}}=0$ otherwise. \end{proof} \begin{theorem} \label{t1} Let the map $(w,z) \rightarrow (\varphi(w,z),\psi(w,z))$ be injective, and the corresponding equation~\eqref{pqd} be uniquely solvable for $v_{1,1}$. Then the equation~\eqref{invt} possesses an~$i$-integral of order~$m$ and a~$j$-integral of order~$n$ if and only if the equation~\eqref{pqd} possesses an~$i$-integral of order $m+1$ and a~$j$-integral of order $n-1$. If $n=2$ and~\eqref{pqd} is not uniquely solvable for~$v_{1,1}$, then the Darboux integrability of~\eqref{invt} implies the existence of an $(m+1)$th order~$i$-integral and a~first-order~$j$-integral for an equation $v_{1,1}=Q(v,v_{1,0},v_{0,1})$ that satisfies the relationship $\Omega(v_{0,1},Q(v,v_{1,0},v_{0,1}))=\Upsilon(v,v_{1,0})$. \end{theorem} \begin{proof} We can assume without loss of generality that the integrals of~\eqref{invt} have the form $I(u,u_{1,0},\dots,u_{m,0})$ and $J(u,u_{0,1},\dots,u_{0,n})$. Substituting $\Omega(\varphi,T_{\rm i}(\varphi))$ instead of~$u$ into the~$i$-integral, we obtain $I(\Omega,\Omega_1,\dots,\Omega_m)$, where $\Omega_k=\Omega (T_{\rm i}^k(\varphi),T_{\rm i}^{k+1}(\varphi))$. The relationship $T_{\rm j}(I)=I$ takes the form \begin{gather} \label{ipq} I(\Upsilon,\Upsilon_1,\dots,\Upsilon_m)=I(\Omega,\Omega_1,\dots,\Omega_m), \qquad \Upsilon_k=\Upsilon\big(T_{\rm i}^k(\varphi),T_{\rm i}^{k+1}(\varphi)\big), \end{gather} because $v=\varphi(u,u_{1,0})$ satisf\/ies equation~\eqref{pqd} and hence $T_{\rm j}$ maps $\Omega(\varphi,T_{\rm i}(\varphi))$ into $\Upsilon(\varphi,T_{\rm i}(\varphi))$. But~$\varphi$, $T_{\rm i}(\varphi)$, $\dots$, $T_{\rm i}^{m+1}(\varphi)$ are functionally independent by Lemma~\ref{l0}, and~\eqref{ipq} can be valid only if the relation \begin{gather*} I(\Upsilon(v,v_{1,0}),\Upsilon(v_{1,0},v_{2,0}),\dots,\Upsilon(v_{m,0},v_{m+1,0})) \\ \qquad =I(\Omega(v,v_{1,0}),\Omega(v_{1,0},v_{2,0}),\dots,\Omega(v_{m,0},v_{m+1,0})) \end{gather*} holds true identically for arbitrary~$v, v_{1,0}, \dots, v_{m+1,0}$. And this means that the right-hand side of the last relationship is an~$i$-integral of equation~\eqref{pqd}. According to Lemma~\ref{intf}, $J=\Phi(\varphi,\dots,T_{\rm j}^{n-1}(\varphi))$. Let us rewrite~\eqref{pqd} as $v_{1,1}=Q(v,v_{1,0},v_{0,1})$. Since $T_{\rm i}(\Phi)=T_{\rm i}(J)=J=\Phi$ and equation~\eqref{pqd} holds true for $v=\varphi(u,u_{0,1})$, the function $\Phi(v,v_{0,1},\dots,v_{0,n-1})$ satisf\/ies the def\/ining relation \begin{gather} \label{defe} \Phi\big(v_{1,0},Q,\dots,T_{\rm j}^{n-2}(Q)\big)=\Phi(v,v_{0,1},\dots,v_{0,n-1}) \end{gather} for~$j$-integral of~\eqref{pqd} when $v_{1,0}=T_{\rm i}(\varphi)$ and $v_{0,k}=T_{\rm j}^k(\varphi)$, $k=\overline{0,n-1}$ (other variables are absent in the def\/ining relation). But $T_{\rm i}(\varphi)=\psi$ and $T_{\rm j}^k(\varphi)$ are functionally independent and, therefore,~\eqref{defe} is valid only if it holds true identically for arbitrary $v_{1,0}$ and $v_{0,k}$, $k=\overline{0,n-1}$. Thus, $\Phi(v,v_{0,1},\dots,v_{0,n-1})$ is a~$j$-integral of equation~\eqref{pqd}. Conversely, let $I(\Omega(v,v_{1,0}),\Omega(v_{1,0},v_{2,0}),\dots,\Omega(v_{m,0},v_{m+1,0}))$ and $\Phi(v,v_{0,1},\dots,v_{0,n-1})$ be inte\-grals of~\eqref{pqd}. Since $v=\varphi(u,u_{1,0})$ satisf\/ies~\eqref{pqd} for any solution~$u$ of the equation~\eqref{invt} and $u=\Omega(\varphi,T_{\rm i}(\varphi))$ identically holds true, the relationships \begin{gather*} T_{\rm i}\big(\Phi(\varphi,T_{\rm j}(\varphi),\dots,T_{\rm j}^{n-1}(\varphi))\big) =\Phi\big(\varphi,T_{\rm j}(\varphi)\dots,T_{\rm j}^{n-1}(\varphi)\big), \\ T_{\rm j}(I(u,u_{1,0},\dots,u_{m,0}))=I(u,u_{1,0},\dots,u_{m,0}) \end{gather*} follow from the def\/ining relation for integrals of~\eqref{pqd}. \end{proof} In particular, this theorem implies that any Darboux integrable equation~\eqref{invt} admitting a~second-order $j$-integral can be derived from a~Darboux integrable equation possessing a~f\/irst-order~$j$-integral. But the equations of the latter kind were described (under an additional assumption) in the recent work~\cite{Stn} and we only need to select the equations of the form~\eqref{pqd} among them. \begin{lemma} \label{ll3} Let an equation \begin{gather} \label{fpr} \tilde{v}_{1,1}=Q(\tilde{v},\tilde{v}_{1,0},\tilde{v}_{0,1}), \qquad \frac{\partial Q}{\partial \tilde{v}} \frac{\partial Q}{\partial \tilde{v}_{1,0}} \frac{\partial Q}{\partial \tilde{v}_{0,1}} \ne 0 \end{gather} be Darboux integrable, satisfy a~relationships of the form $\tilde{\Omega}(\tilde{v}_{0,1},Q(\tilde{v},\tilde{v}_{1,0},\tilde{v}_{0,1}))=\tilde{\Upsilon}(\tilde{v},\tilde{v}_{1,0})$ and possess a~first-order~$j$-integral $\Phi(\tilde{v},\tilde{v}_{0,1})$ such that the equation $\Phi(\tilde{v},\tilde{v}_{0,1})=\tilde{w}$ is uniquely solvable for $\tilde{v}$. Then a~point transformation $\tilde{v}=\zeta(v)$ relates~\eqref{fpr} to an equation of the form \begin{gather} \label{strt} \frac{\delta D - v_{1,1} (A v_{0,1} + C - \delta B)}{v_{1,1}-v_{0,1}} = \frac {(v+B)(\delta v_{1,0} +C)-AD}{v_{1,0}-v}, \end{gather} where the constants~$A$,~$B$,~$C$,~$D$ and~$\delta$ satisfy the inequalities $|\delta A| + |C - \delta B| + |D| \ne 0$, $|\delta|+|A|+|C| \ne 0$. \end{lemma} \begin{proof} The def\/ining relation $\Phi(\tilde{v}_{1,0}, Q) =\Phi(\tilde{v},\tilde{v}_{0,1})$ for the~$j$-integral~$\Phi$ is uniquely solvable for the f\/irst argument of the left-hand side by the theorem assumptions. Hence, the equation~\eqref{fpr} is uniquely solvable for $\tilde{v}_{1,0}$. According to~\cite{Stn}, an equation of the form~\eqref{fpr} is Darboux integrable, uniquely solvable for $\tilde{v}_{1,0}$ and admits an autonomous f\/irst-order~$j$-integral only if a~point change of variables $\tilde{v}=\zeta(v)$ relates this equation to an equation of the form \begin{gather} \label{veq} v_{1,1}=\alpha (\phi) + \frac{\beta(\phi)}{\gamma(\phi)-v_{1,0}}, \qquad \beta \ne 0, \end{gather} where~$\phi$ is the function of~$v$ and $v_{0,1}$ that satisf\/ies the relationship \begin{gather} \label{veq2} v_{0,1}=\alpha (\phi) + \frac{\beta(\phi)}{\gamma(\phi)-v}. \end{gather} It is obvious that the last relationship can hold true only if $|\alpha'|+|\beta'|+|\gamma'| \ne 0$. The function~$\phi$ is a~$j$-integral, and \begin{gather} \label{gin} I[v] = \frac{v_{3,0}-v_{1,0}}{v_{3,0}-v_{2,0}} \cdot \frac{v_{2,0}-v}{v_{1,0}-v} \end{gather} is an~$i$-integral of~\eqref{veq}. It should be noted that some equations of the form~\eqref{veq}, \eqref{veq2} admit~$i$-integrals of order less than 3, but all such equations possess the integral~\eqref{gin} too (see~\cite{Stn} for more details). An equation $v_{1,1}=F(v,v_{1,0},v_{0,1})$ can be written in the form~\eqref{pqd} only if the condition $(F_v/F_{v_{1,0}})_{v_{0,1}} =0$ holds true. Substituting the right-hand side of~\eqref{veq} into this condition, we obtain \begin{gather} \label{cond} \frac{\partial}{\partial v_{0,1}} \left(\frac{\alpha' (\gamma - v_{1,0})^2 + \beta' (\gamma - v_{1,0}) - \beta \gamma'}{\beta} \frac{\partial \phi}{\partial v} \right) = 0. \end{gather} On the other hand, dif\/ferentiation of~\eqref{veq2} with respect to~$v$ gives rise to \begin{gather*} \beta= \big(\beta \gamma' - \alpha' (\gamma - v)^2 - \beta' (\gamma - v)\big) \frac{\partial \phi}{\partial v}. \end{gather*} The condition $\beta \ne 0$ implies that the both factors in the right-hand side do not equal zero. Therefore,~\eqref{cond} takes the form \begin{gather*} \frac{\partial}{\partial v_{0,1}} \left(\frac{\alpha' (\gamma - v_{1,0})^2 + \beta' (\gamma - v_{1,0}) - \beta \gamma'}{\alpha' (\gamma - v)^2 + \beta' (\gamma - v) - \beta \gamma'} \right) = 0. \end{gather*} Performing the dif\/ferentiation in the left-hand side of the last relationship, we obtain \begin{gather*} \big[(\alpha' \gamma'' -\alpha'' \gamma ') \big(v_{1,0}^2 - 2 \gamma v_{1,0}\big) + \big(\gamma' \beta'' - \gamma'' \beta' + 2 \alpha' (\gamma')^2\big) v_{1,0}\big] \beta \\ \qquad {}+ \big(\alpha'' \beta' - \alpha' \beta'' - 2 (\alpha')^2 \gamma'\big)\big(v_{1,0}^2(\gamma - v) + v_{1,0} \big(v^2 -\gamma^2\big)\big) + \dots=0, \end{gather*} where the dots signify terms without $v_{1,0}$. The left-hand side is a~polynomial in~$v$ and $v_{1,0}$. The coef\/f\/icients of it depend on~$\phi$ only and must be equal to zero because $\phi(v,v_{0,1})$,~$v$ and $v_{1,0}$ are functionally independent. Thus, we have \begin{gather} \label{abg} \alpha' \gamma'' -\alpha'' \gamma' = \gamma' \beta'' - \gamma'' \beta' + 2 \alpha' (\gamma')^2 = \alpha' \beta'' - \alpha'' \beta' + 2 (\alpha')^2 \gamma' = 0. \end{gather} If~$\phi$ satisf\/ies~\eqref{veq2}, then any function of~$\phi$ also satisf\/ies~\eqref{veq2} (with another~$\alpha$,~$\beta$ and~$\gamma$). This is why we can assume without loss of generality that $\alpha=\delta \phi$ if $\alpha' \ne 0$. Under this assumption,~\eqref{abg} gives rise to $\gamma=A \phi - B$ and $\beta=D - C \phi -\delta \phi (A \phi - B)$. It is easy to check that an appropriate change $v \rightarrow v+\lambda$ in~\eqref{veq2} reduces the two other cases $\alpha'=0$, $\gamma = \phi$ and $\alpha' = \gamma' =0$, $\beta = \phi$ to the same formulas for~$\alpha$,~$\beta$ and~$\gamma$ with $\delta=0$, $A=1$ and $\delta=A=D=0$, $C=-1$, respectively. Substituting these formulas into~\eqref{veq2} and solving it for~$\phi$, we obtain \begin{gather} \label{phi} \phi = \frac{v_{0,1}(v + B)+D}{A v_{0,1} + \delta v+C}. \end{gather} The corresponding equation~\eqref{veq} is \begin{gather*} v_{1,1}=\frac{((v+B)(\delta v_{1,0} +C) - AD)v_{0,1} + \delta D (v_{1,0}- v)}{A(v_{1,0}-v) v_{0,1} + (v_{1,0}+B)(\delta v +C) - AD} \end{gather*} and can be rewritten as~\eqref{strt}. \end{proof} If we perform a~point transformation $v=\eta(\tilde{v})$ in~\eqref{pqd}, then the corresponding system \begin{gather*} u=\Omega(\eta(\tilde{v}),\eta(\tilde{v}_{1,0})), \qquad u_{0,1}=\Upsilon(\eta(\tilde{v}),\eta(\tilde{v}_{1,0})) \end{gather*} has the solution $\eta(\tilde{v})=\varphi(u,u_{0,1})$, $\eta(\tilde{v}_{1,0})=\psi(u,u_{0,1})$ that generates the unchanged equation~\eqref{invt}. Thus, we can restore the original equation~\eqref{invt} after a~point transformation in~\eqref{pqd}. But the equation $\tilde{\Omega}(\tilde{v}_{0,1},\tilde{v}_{1,1})=\tilde{\Upsilon}(\tilde{v},\tilde{v}_{1,0})$ in the formulation of Lemma~\ref{ll3} may also dif\/fer from~\eqref{pqd} via a~transformation $\tilde{\Omega}=\xi(\Omega)$, $\tilde{\Upsilon}=\xi(\Upsilon)$, that corresponds to the point change $\tilde{u}=\xi(u)$ in~\eqref{invt}. Theorem~\ref{t1}, the above lemmas and the reasonings of the previous paragraph mean that the transformation from~\eqref{pqd} to~\eqref{invt} for equation~\eqref{strt} gives us all, up to point transformations, Darboux integrable equations~\eqref{invt} admitting second-order~$j$-integrals of the form $\Phi(\varphi(u,u_{0,1}),\varphi(u_{0,1},u_{0,2}))$, where the equation $\Phi(v,v_{0,1})=w$ is uniquely solvable for~$v$. Let us perform this transformation. The system~\eqref{pqsd} for~\eqref{strt} takes the form \begin{gather} \label{u10} u=\frac{\delta D - v_{1,0} (A v + C - \delta B)}{v_{1,0}-v}, \\ \label{u11} u_{0,1} = \frac {(v+B)(\delta v_{1,0} +C) - AD}{v_{1,0}-v}. \end{gather} Equation~\eqref{u10} can be rewritten in the form \begin{gather} \label{uint} v_{1,0}=\frac{u v + \delta D}{u + A v + C - \delta B}. \end{gather} Replacing $v_{1,0}$ with the right-hand side of~\eqref{uint}, we transform~\eqref{u11} into the algebraic equation $P_2 v^2 + P_1 v + P_0 =0$, where \begin{gather} \label{p2} P_2 = \delta u + A u_{0,1} + A C, \\ \label{p1} P_1 = (C + \delta B) u + (C - \delta B) u_{0,1} + (B C - A D - \delta D) (A - \delta) + C^2, \\ \label{p0} P_0 = (B C - A D) u - \delta D u_{0,1} + B C (C - \delta B) + D \big(\delta A B - A C + \delta^2 B\big). \end{gather} Solving this equation for~$v$ and substituting a~solution $\theta (u,u_{0,1})$ into~\eqref{uint}, we obtain the following classif\/ication result. \begin{proposition} \label{d2p} Let the map $(w,z) \rightarrow (\varphi(w,z),\psi(w,z))$ be injective, and let the equation~\eqref{invt} possess a~second-order autonomous~$j$-integral~$J$ such that $\varphi(u,u_{0,1})$ is uniquely expressed in terms of~$J$ and $\varphi(u_{0,1},u_{0,2})$. Then the equation~\eqref{invt} is Darboux integrable if and only if a~point change of variables $\rho(u) \rightarrow u$ reduces this equation to the form \begin{gather} \label{ggen} T_{\rm i}(\theta (u,u_{0,1}))=\frac{u \theta (u,u_{0,1}) + \delta D}{A \theta(u,u_{0,1}) + u + C - \delta B}, \end{gather} where~$\theta$ is a~solution of the equation $P_2 \theta^2 + P_1 \theta + P_0=0$ with the coefficients defined by formulas~\eqref{p2}--\eqref{p0}, and the constants~$A$,~$B$,~$C$,~$D$,~$\delta$ satisfy the conditions $|\delta|+|A|+|C| \ne 0$, $|\delta A| + |C - \delta B| + |D| \ne 0$. \end{proposition} Theorem~\ref{t1} guaranties that~\eqref{ggen} is Darboux integrable, and the proof of Theorem~\ref{t1} gives us the way to construct integrals of this equation. Substituting~$\theta$ instead of~$v$ into~\eqref{phi}, we obtain a~$j$-integral \begin{gather*} J[u]=\phi(\theta,T_{\rm j}(\theta))=\frac{T_{\rm j}(\theta)(\theta + B)+D}{A T_{\rm j}(\theta) + \delta \theta+C}. \end{gather*} Equation~\eqref{uint} (and its shifts in~$i$) allows us to represent $v_{1,0}$, $v_{2,0}$, $v_{3,0}$ as functions of~$v$,~$u$, $u_{1,0}$, $u_{2,0}$. Replacing $v_{k,0}$ with these functions in~\eqref{gin}, we derive the formula \begin{gather*} I[u]= \frac{(u_{2,0}+u_{1,0} + C - \delta B) (u_{1,0}+u + C - \delta B)}{u_{1,0}^2 + u_{1,0} (C - \delta B) - \delta A D} \end{gather*} for an~$i$-integral of the equation~\eqref{ggen}. As it is shown in Section~\ref{sec1}, the discrete Liouville equation~\eqref{dle} satisf\/ies all assumptions of Proposition~\ref{d2p}. Therefore, it lies in the equation family~\eqref{ggen} as a~particular case $C = -1$, $B \ne 0$, $A=\delta=0$. Another special case $C=D=0$, $B=-1$, $A \ne -1$, $\delta =1$ gives us the equation $u_{1,1} (u+1) = u_{1,0} (u_{0,1} +1)$, which is contained in a~list of Darboux integrable equations in~\cite{GY}. \section{Transformation of symmetries} \label{s3} Let $h_*$ designate the Frech\'et derivative (linearization operator) \begin{gather*} h_* = \sum\limits_{q=-\infty}^{+\infty} \frac{\partial h}{\partial u_{q,0}} T^q_{\rm i} + \sum\limits_{\genfrac{}{}{0pt}{}{q=-\infty}{q \ne 0}}^{+\infty} \frac{\partial h}{\partial u_{0,q}} T^q_{\rm j} \end{gather*} of the function $h[u]$, and let $\partial_f$ denote the dif\/ferentiation with respect to~$t$ by virtue of the equation $u_t=f[u]$. The formula ${\partial_f (h[u]) = h_* (f)}$ def\/ines $\partial_f$ on the functions of the dynamical variables. Let us consider a~dif\/ferential-dif\/ference equation of the form \begin{gather} \label{eg} u_t=g(u_{0,n},u_{0,n+1},\dots,u_{0,k}). \end{gather} \begin{definition} \label{drp} If a~function $\phi (u_{0,l},u_{0,l+1},\dots,u_{0,m})$, $m > l$, satisf\/ies the inequality $\phi_{u_{0,l}} \phi_{u_{0,m}} \ne 0$ and a~relationship\footnote{This relationship means that $v=\phi[u]$ maps solutions of~\eqref{eg} into solutions of the equation $v_t=\hat{g}$.} of the form \begin{gather} \label{dpd} \partial_g (\phi) = \hat{g} \big(T^n_{\rm j}(\phi), T^{n+1}_{\rm j}(\phi), \dots, T^k_{\rm j}(\phi)\big), \end{gather} then we say that equation~\eqref{eg} admits the dif\/ference substitution \begin{gather} \label{rp} v=\phi (u_{0,l},u_{0,l+1},\dots,u_{0,m}) \end{gather} into the equation $v_t=\hat{g}(v_{0,n},v_{0,n+1},\dots,v_{0,k})$. We call~\eqref{rp} a~Miura-type substitution if there exist operators~\eqref{rop} and \begin{gather*} \hat{R}=\sum\limits_{q=l}^{r+m} \hat{\lambda}_q (v_{0,\hat{\varrho}},v_{0,\hat{\varrho}+1},\dots,v_{0,\hat{s}}) T^q_{\rm j} \end{gather*} such that the equation $u_t=R(\xi (T^{p}_{\rm j} (\phi), T^{p+1}_{\rm j} (\phi), \dots))$ admits the substitution~\eqref{rp} into the equation $v_t=\hat{R} (\xi (v_{0,p},v_{0,p+1},\dots))$ for any integer~$p$ and any function $\xi$ depending on a~f\/inite number of the arguments. \end{definition} Since mixed shifts $u_{q,p}$, $q p \ne 0$, do not appear in the def\/ining relation~\eqref{dpd}, the above def\/inition in no way uses equation~\eqref{uij}. However, equations~\eqref{uij} can generate dif\/ference substitutions, for example, in the way that was used in~\cite{Yam90} for dif\/ferential-dif\/ference analogues of equations~\eqref{invt}. \begin{theorem} \label{t2} If~\eqref{eg} is a~symmetry of~\eqref{invt}, then the equation~\eqref{eg} admits the difference substitutions $v=\varphi(u,u_{0,1})$ and $v=\psi(u,u_{0,1})$ into an equation $v_t=\hat{g}(v_{0,n},v_{0,n+1},\dots,v_{0,k})$. If the map $(w,z) \rightarrow (\varphi(w,z),\psi(w,z))$ is injective and the corresponding equation~\eqref{pqd} is uniquely solvable for $v_{1,0}$, $v_{1,1}$ and $v_{0,1}$, then the following proposition is also true for any symmetry $u_t=f[u]$ of equation~\eqref{invt} such that $f\ne 0$. Let $\hat{f}[v]$ designate the function $\varphi_*(f)$ after substituting $u=\Omega(v,v_{1,0})$ into it and excluding the variables of the form $v_{p,q}$, $p q \ne 0$, by virtue of~\eqref{pqd}. Then $\hat{f}[v] \ne 0$ and $v_t = \hat{f}[v]$ is a~symmetry of equation~\eqref{pqd}. If, in addition, $f_{u_{\delta,0}} \ne 0$ for some $\delta>0$ $(\delta<0)$ or $f_{u_{0,\sigma}} \ne 0$ for some $\sigma>0$ ($\sigma < 0$), then $\hat{f}[v]$ essentially depends on $v_{d,0}$ for some $d \ge \delta$ ($d \le \delta$) or on $v_{0,b}$ for some $b \ge \sigma$ ($b \le \sigma$), respectively. \end{theorem} Although the proof is almost identical to the proof of the analogous proposition for semi-discrete equations in~\cite{Yam90}, we include it for the sake of completeness. It should also be noted that for the rest part of the present article we need only the f\/irst sentence of the theorem. \begin{proof} The application of the Frech\'et derivative to both sides of~\eqref{qua} gives rise to \begin{gather*} T_{\rm i} \left(\frac{\partial \varphi}{\partial u_{0,1}} \right)\! F_{*} + T_{\rm i} \left(\frac{\partial \varphi}{\partial u} \right) = \psi_{*} \ \Rightarrow \ T_{\rm i} \left(\frac{\partial \varphi}{\partial u_{0,1}} \right)\! L = T_{\rm i} \left(\frac{\partial \varphi}{\partial u_{0,1}} \right)\! (T_{\rm i} T_{\rm j} - F_{*}) = T_{\rm i} \circ \varphi_{*} - \psi_{*}, \end{gather*} where $\circ$ denotes the composition of operators. Thus, the def\/ining relation $L(f)=0$ for the symmetry $u_t=f[u]$ takes the form $T_{\rm i} (\varphi_{*} (f))= \psi_{*}(f)$. Let~\eqref{eg} be a~symmetry of~\eqref{invt}. The function $\varphi_*(g)$ can be rewritten as \begin{gather} \label{dgf} \varphi_*(g)=\hat{g}\big(u_{0,n},T^n_{\rm j}(\varphi),T^{n+1}_{\rm j}(\varphi),\dots,T^k_{\rm j}(\varphi)\big). \end{gather} The substitution of~\eqref{dgf} into the def\/ining relation $\psi_{*} (g) = T_{\rm i} (\varphi_{*} (g))$ gives rise to \begin{gather} \label{dgp} \psi_{*} (g) = \hat{g}\big(T_{\rm i}(u_{0,n}),T^n_{\rm j}(\psi),T^{n+1}_{\rm j}(\psi),\dots,T^k_{\rm j}(\psi)\big). \end{gather} But~\eqref{hipc} and~\eqref{u1n} imply that $T_{\rm i}(u_{0,n})$ depends on $u_{1,0}$ whereas $\psi_{*} (g)$ and $T^n_{\rm j}(\psi), \dots, T^k_{\rm j}(\psi)$ do not. Therefore, $\hat{g}$ does not depend on its f\/irst argument and~\eqref{dgf}, \eqref{dgp} take the form \begin{gather*} \varphi_*(g)=\hat{g}\big(T^n_{\rm j}(\varphi),T^{n+1}_{\rm j}(\varphi),\dots,T^k_{\rm j}(\varphi)\big), \\ \psi_{*} (g) = \hat{g}\big(T^n_{\rm j}(\psi),T^{n+1}_{\rm j}(\psi),\dots,T^k_{\rm j}(\psi)\big), \end{gather*} respectively. For the symmetries of the general form $u_t=f[u]$, equation~\eqref{pqsd} and the relationship $T_{\rm i} (\varphi_{*} (f))= \psi_{*} (f)$ imply \begin{gather} \label{invs} \partial_f (\Omega(\varphi,\psi))= \frac{\partial \Omega}{\partial v} \varphi_*(f) + \frac{\partial \Omega}{\partial v_{1,0}} \psi_*(f) = \frac{\partial \Omega}{\partial v} \varphi_*(f) + \frac{\partial \Omega}{\partial v_{1,0}} T_{\rm i} (\varphi_*(f))= \partial_f (u)= f, \\ \partial_f (\Upsilon(\varphi,\psi))= \frac{\partial \Upsilon}{\partial v} \varphi_*(f) + \frac{\partial \Upsilon}{\partial v_{1,0}} \psi_*(f) = \frac{\partial \Upsilon}{\partial v} \varphi_*(f) + \frac{\partial \Upsilon}{\partial v_{1,0}} T_{\rm i} (\varphi_*(f))= \partial_f (u_{0,1})= T_{\rm j}(f). \nonumber \end{gather} The last two formulas, together with $\psi=T_{\rm i}(\varphi)$, mean that the function $\breve{f}[u]:=\varphi_*(f)$ satisf\/ies the relation\-ship \begin{gather} T_{\rm j} \left(\frac{\partial \Omega(\varphi,T_{\rm i}(\varphi))}{\partial v} \breve{f} + \frac{\partial \Omega(\varphi,T_{\rm i}(\varphi))}{\partial v_{1,0}} T_{\rm i} (\breve{f}) \right) = T_{\rm j}(f)\nonumber \\ \qquad =\frac{\partial\Upsilon(\varphi,T_{\rm i}(\varphi))}{\partial v} \breve{f} + \frac{\partial \Upsilon(\varphi,T_{\rm i}(\varphi))}{\partial v_{1,0}} T_{\rm i} (\breve{f}).\label{defsym} \end{gather} Using the identities $u=\Omega(\varphi,T_{\rm i}(\varphi))$, $\Omega(T_{\rm j}(\varphi),T_{\rm i} T_{\rm j} (\varphi))=\Upsilon(\varphi,T_{\rm i}(\varphi))$ and their consequences derived by shifts in~$i$ and~$j$, we can represent $\breve{f}[u]$ as a~function $\hat{f}[\varphi]$ of only $T_{\rm i}^p(\varphi)$, $T_{\rm j}^q(\varphi)$, $p,q \in \mathbb{Z}$, $q \ne 0$, and exclude all mixed shifts $T_{\rm i}^s T_{\rm j}^r(\varphi)$, $sr \ne 0$, from~\eqref{defsym}. As a~result of this, in~\eqref{defsym} the operators $T_{\rm i}$ and $T_{\rm j}$ act on $v_{p,0}=T_{\rm i}^p(\varphi)$, $v_{0,q}=T_{\rm j}^q(\varphi)$ by virtue of~\eqref{pqd}, i.e.~\eqref{defsym} becomes the def\/ining relation for symmetries of~\eqref{pqd}. But the functions $T_{\rm i}^p(\varphi)$, $T_{\rm j}^q(\varphi)$ are functionally independent by Lemma~\ref{l0}, and $\hat{f}[v]$ therefore satisf\/ies this def\/ining relation not only if $v_{p,0}=T_{\rm i}^p(\varphi)$, $v_{0,q}=T_{\rm j}^q(\varphi)$ but for arbitrary values of $v_{p,0}$, $v_{0,q}$. If $\varphi_*(f)=0$, then~\eqref{invs} implies $f=0$. This is why $\hat{f}[v] \ne 0$ if $f \ne 0$. The same logic proves the dependence of $\hat{f}[v]$ on $v_{d,0}$ if we take the relationship $T_{\rm i} T_{\rm j}^q(\varphi)= \psi(u_{0,q},u_{0,q+1})$ and the formulas for $T_{\rm i}^p(\varphi)$, $T_{\rm j}^q(\varphi)$ from the proof of Lemma~\ref{l0} into account. The dependence of $\hat{f}[v]$ on $v_{0,b}$ follows directly form the relation $\varphi_{u_{0,1}} T_{\rm j} (f) + \varphi_u f=\hat{f}[v]\big|_{v=\varphi}$ and the fact that~$f$ can not depend on $u_{0,\sigma}$ for positive or negative~$\sigma$ if $\hat{f}[v]$ does not depend on $v_{0,b}$ for all $b \ge \sigma$ or for all $b \le \sigma$, respectively. \end{proof} \begin{corollary} \label{dc1} If an equation of the form~\eqref{invt} is uniquely solvable for $u_{1,0}$ and Darboux integrable, then $v=\varphi(u,u_{0,1})$ and $w=\psi(u,u_{0,1})$ are Miura-type substitutions. \end{corollary} Recall that in the present paper we understand~\eqref{invt} as an equation of the form~\eqref{uij} satisfying the relationship~\eqref{qua}. According to this, in Corollaries~\ref{dc1},~\ref{dc2} we assume that the equation~\eqref{uij} is uniquely solvable for $u_{1,0}$ (i.e.~the right-hand side of the corresponding equation~\eqref{upm} is well-def\/ined). We can replace this assumption with the more strong assumption that the formula $v=\varphi(u,u_{0,1})$ is uniquely solvable for~$u$. The last assumption guarantees that any equation~\eqref{uij} satisfying~\eqref{qua} is uniquely solvable for $u_{1,0}$. \begin{proof} It is easy to see that $\hat{R}$ in Def\/inition~\ref{drp} coincides with the operator $\phi_* \circ R = \sum\limits_{q=l}^{r+m} \breve{\lambda}_q[u] T_{\rm j}^q$ and $v=\phi[u]$ is a~Miura-type substitution if all $\breve{\lambda}_q[u]$ can be expressed in terms of $T_{\rm j}^b (\phi)$ only. According to~\cite{AdS,Stn}, if equation~\eqref{uij} is Darboux integrable and uniquely solvable for $u_{1,0}$, then it possesses symmetries of the form~\eqref{rop}, \eqref{ed}. Theorem~\ref{t2} implies that the symmetries~\eqref{ed} admit the substitutions $v=\varphi(u,u_{0,1})$ and $w=\psi(u,u_{0,1})$, i.e.~$\varphi_*(R(\eta))$ and $\psi_* (R(\eta))$ can be respectively expressed in terms of $v_{0,b}= T_{\rm j}^b (\varphi)$ and $w_{0,b}:=T_{\rm j}^b (\psi)$, $b \in \mathbb{Z}$. But in the proof of Lemma~\ref{intf} we demonstrate that any~$j$-integral $J[u]$ of~\eqref{invt} can be represented both as a~function of $T_{\rm j}^b (\varphi)$ and as a~function of $T_{\rm j}^b (\psi)$. Hence, the same is true for any function~$\eta$ of the~$j$-integrals. Due to the arbitrariness of~$\eta$ in~\eqref{ed}, this implies that the coef\/f\/icients of the operators $\varphi_* \circ R$ and $\psi_* \circ R$ can be expressed in terms of $T_{\rm j}^b (\varphi)$ and $T_{\rm j}^b (\psi)$, respectively. Thus, the equation $u_t=R(\xi)$ admits the substitutions $v=\varphi(u,u_{0,1})$ and $w=\psi(u,u_{0,1})$ for any function $\xi$ depending on arguments of the forms $T_{\rm j}^b (\varphi)$ and $T_{\rm j}^b (\psi)$, respectively. \end{proof} It is proved in~\cite{foi} that $v=\phi(u,u_{0,1})$ is a~Miura-type substitution only if~$\phi$ satisf\/ies the relationship \begin{gather*} \zeta(u_{0,1})=\alpha (\phi) + \frac{\beta(\phi)}{\gamma(\phi)-\zeta(u)}, \end{gather*} for some functions~$\alpha$,~$\beta$,~$\gamma$ and~$\zeta$ such that $\beta \zeta' \ne 0$. This implies the following proposition. \begin{corollary} \label{dc2} If an equation of the form~\eqref{invt} is uniquely solvable for $u_{1,0}$ and Darboux integrable, then there exist functions~$\alpha$,~$\beta$,~$\gamma$,~$\zeta$ and $\hat{\alpha}$, $\hat{\beta}$, $\hat{\gamma}$, $\hat{\zeta}$ such that \begin{gather*} \zeta(u_{0,1})=\alpha (\varphi) + \frac{\beta(\varphi)}{\gamma(\varphi)-\zeta(u)}, \qquad \beta \zeta' \ne 0, \\ \hat{\zeta}(u_{0,1})=\hat{\alpha}(\psi) + \frac{\hat{\beta}(\psi)}{\hat{\gamma}(\psi)-\hat{\zeta}(u)}, \qquad \hat{\beta} \hat{\zeta}' \ne 0. \end{gather*} \end{corollary} Applying $T_{\rm i}$ and $T_{\rm i}^{-1}$ to the both sides of the former and the latter relationships, respectively, we obtain that any Darboux integrable equation~\eqref{invt} can be represented in the form \begin{gather*} \zeta(u_{1,1})=\alpha (\psi) + \frac{\beta(\psi)}{\gamma(\psi)-\zeta(u_{1,0})} \end{gather*} as well as in the form \begin{gather*} \hat{\zeta}(u_{-1,1})=\hat{\alpha}(\varphi) + \frac{\hat{\beta}(\varphi)}{\hat{\gamma}(\varphi)-\hat{\zeta}(u_{-1,0})}. \end{gather*} We can assume either $\zeta(u)=u$ or $\hat{\zeta}(u)=u$ without loss of generality because we can perform either the point change of variables $\zeta(u)\rightarrow u$ or the point change $\hat{\zeta}(u) \rightarrow u$. \subsection*{Acknowledgments} The author thanks the referees for useful suggestions. This work is partially supported by the Russian Foundation for Basic Research (grant number 13-01-00070-a). \pdfbookmark[1]{References}{ref}
\section{Introduction}\label{sec1} \subsection*{Model and motivation} We propose to study the following system of infinitely-many interacting diffusions: {\renewcommand{\theequation}{SHE} \begin{equation}\label{SHE} \frac{\d u_t(x)}{\d t} = (\mathscr{L}u_t) (x) + \sigma \bigl(u_t(x)\bigr) \frac{\d B_t(x)}{\d t}, \end{equation}} \hspace*{-3pt}where $t>0$ denotes the \emph{time} variable and $x\in\Z^d$ is the \emph{space} variable. In parts of the literature, \eqref{SHE} is thought of as a \emph{stochastic heat equation} on $(0,\infty)\times \Z^d$, viewed as a semi-discrete stochastic partial differential equation. We interpret \eqref{SHE} as an infinite-dimensional system of It\^o stochastic differential equations, where $\{B(x)\}_{x\in\Z^d}$ denotes a field of independent standard linear Brownian motions and $\sigma\dvtx\R\to\R$ is a Lipschitz-continuous nonrandom function with \setcounter{equation}{0} \begin{equation} \label{eq:sigma(0)=0} \sigma(0)=0. \end{equation} The drift operator $\mathscr{L}$ acts on the variable $x$ only, and is the generator of a continuous-time random walk $X:=\{X_t\}_{t\ge0}:=\{\sum_{j=1}^{N_t}Z_j\}_{t\ge0}$ on $\Z^d$ where $N_t$ is a Poisson process with jump-rate one and the $Z_j$'s are i.i.d. random variables with values in $\Z^d$. We consider only initial values $u_0\dvtx\Z^d\to\R$ that are nonrandom and satisfy \begin{equation} \label{init} u_0(x)\ge0\qquad\mbox{for all }x\in\Z^d \quad\mbox{and}\quad 0<\sup_{x\in\Z^d}u_0(x)<\infty, \end{equation} though some of the theory developed here applies to more general initial profiles. According to Shiga and Shimizu \cite{ShigaShimizu}, condition \eqref{init} ensures that the system \eqref{SHE} has an a.s.-unique solution. Such systems have been studied extensively \cite{CKM,CM94,CFG,CoxGreven,CranstonMolchanov,CranstonMountfordShiga,CGM,Funaki,GrevenDenHollander,Mueller,Shiga,ShigaShimizu}, most commonly in the context of well-established models of statistical mechanics, population genetics and related models of infinitely-many interacting diffusion processes. One of the central examples of this literature is the \emph{parabolic Anderson model} \cite{CM94}---also known as \emph{diffusion in random potential}---which is \eqref{SHE} when the function $\sigma$ is linear. It is not hard to prove that, for the parabolic Anderson model, the \emph{$k$th moment Lyapunov exponent} $\gamma_k(u)$ exists and is positive and finite for all real numbers $k\ge2$, where \begin{equation} \gamma_k(u):= \lim_{t\to\infty} t^{-1}\log\E \bigl( \bigl|u_t(x)\bigr|^k \bigr). \end{equation} [One can prove that, because of \eqref{init}, $\gamma_k(u)$ does not depend on $x$.] Jensen's inequality readily implies that $k\mapsto\gamma_k(u)$ is nondecreasing on $[2 ,\infty)$. In the case that $\mathscr{L}$ denotes the discrete Laplacian on $\Z$, for example, the theory of Carmona and Molchanov \cite{CM94} implies that \begin{equation} \label{CM-I} k\mapsto\gamma_k(u)\qquad\mbox{is strictly increasing on $[2 ,\infty)$}. \end{equation} This property is referred to as \emph{intermittency} and suggests that the random function $u$ develops very tall peaks that are distributed over small space--time ``islands.'' Section~2.4 of Bertini and Cancrini \cite{BC} and Section~7.1 of Khoshnevisan \cite{K-CBMS} describe two heuristic derivations of this ``peaking behavior'' from intermittency condition~\eqref{CM-I}. In the present nonlinear setup, the Lyapunov exponents do not generally exist. Therefore, one considers instead the (maximal) \emph{bottom} and \emph{top Lyapunov exponents} of the solution $u$ to \eqref{SHE}; those are, respectively, defined as \begin{eqnarray} \underline{\gamma}_k(u) &:= &\liminf_{t\to\infty} \sup_{x\in\Z^d} t^{-1} \log\E \bigl( \bigl|u_t(x)\bigr|^k \bigr), \nonumber \\[-8pt] \\[-8pt] \nonumber \overline{\gamma}_k(u) &:= &\limsup_{t\to\infty} \sup _{x\in\Z^d} t^{-1} \log\E \bigl( \bigl|u_t(x)\bigr|^k \bigr). \end{eqnarray} Whenever\vspace*{1pt} $\underline{\gamma}_k(u)=\overline{\gamma}_k(u)$, we write $\gamma_k(u)$ for their common value and think of $\gamma_k(u)$ as the \emph{$k$th moment Lyapunov exponent} of $u$. In the present nonlinear setting, intermittency is defined as the property that the functions $k \mapsto k^{-1}\underline{\gamma}_k(u)$ and $k\mapsto k^{-1}\overline{\gamma}_k(u)$ are both strictly increasing on $[2 ,\infty)$. Because of convexity, one can establish intermittency when $\underline{\gamma}_2(u)>0$ and $\overline{\gamma}_k(u)<\infty$ for all $k\ge2$, \cite{K-CBMS}, Proposition~7.2. We will prove that $\overline{\gamma}_k(u)$ is generically finite for all $k\in[2 ,\infty)$; see Theorem~\ref{th:exist:unique}. Therefore, as far as matters of intermittency are concerned, it remains to establish the positivity of the bottom Lyapunov exponent. This is a nongeneric property. To wit, when $u_0$ is a constant and $\mathscr{L}$ is the discrete Laplacian on $\Z^d$, Carmona and Molchanov \cite{CM94} have shown that $\gamma_2(u)>0$ if and only if $d\in\{1 ,2\}$. This paper is concerned with various results that surround this general topic. As a first example of the type of result that we will establish, let us point out the following, which is related to the mentioned peaking property of the parabolic Anderson model: One can apply Theorem~\ref{th:exist:unique} below to the parabolic Anderson model in order to see that for all $t>0$ there exist finite and positive constants $a(t)$ and $A(t)$ such that \begin{equation} \label{CM-U} a(t) k^2 \le\log\E \bigl( \bigl|u_t(x)\bigr|^k \bigr) \le A(t) k^2, \end{equation} uniformly for all $x\in\Z^d$ and $k\in[2 ,\infty)$. It is then possible to combine this bound, together with the method of Conus et al. \cite{CJK13}, in order to estimate the size of the peaks of the solution relative to the spatial variable $x$. For example, if $\mathscr{L}$ is finite range, then it is possible to prove that the tall peaks grow at all times as $\exp\{\operatorname{const}\,\cdot\, \sqrt{\log\|x\|}\}$ for large values of $\|x\|$ and more precisely, that \begin{equation} \label{peaks} 0<\limsup_{\|x\|\to\infty} \frac{\log|u_t(x)|}{\sqrt{\log\|x\|}} <\infty \qquad\mbox{a.s.} \end{equation} We will not establish this fact since it follows fairly readily from \eqref{CM-U} and the methods of \cite{CJK13}. Instead let us return to \eqref{SHE} in its nonlinear form. In general, properties such as \eqref{CM-U}, whence \eqref{peaks}, can be shown to fail. This is so, for example, when $\sigma$ is bounded; see Conus et al. \cite{CJK13} for analogous results. Therefore, in order to prove \eqref{CM-U} for \eqref{SHE} we need to impose some growth conditions on the nonlinearity $\sigma$. Define \begin{equation} \lip:= \sup_{-\infty<a\neq b<\infty} \biggl\llvert \frac{\sigma(a)-\sigma(b)}{a-b}\biggr \rrvert , \qquad \ell_\sigma:=\inf_{z\in\R} \biggl\llvert \frac{\sigma(z)}{z} \biggr\rrvert . \label{def:Lipell} \end{equation} Because of \eqref{eq:sigma(0)=0}, \begin{equation} \ell_\sigma\le\biggl\llvert \frac{\sigma(z)}{z}\biggr\rrvert \le\lip\qquad \mbox{for all $z\in\R\setminus\{0\}$}. \end{equation} Suppose that $\ell_{\sigma}>0$ and $\sigma(z)>0$ for all $z>0$. Then we will prove that \eqref{CM-U} holds, and in particular the Lyapunov exponents are always positive and finite; see Theorem~\ref{th:exist:unique}. Property \eqref{CM-U} contrasts sharply with the continuous-space analogues of \eqref{SHE} wherein typically the $k$th moment Lyapunov exponents are of sharp order $k^3$ as $k\to\infty$ \cite{ACQ,BorCor,BQS,BC,FK} and implies that, although the intermittency peaks of semi-discrete stochastic PDEs grow rapidly with time, they grow far less rapidly than those of fully-continuous stochastic PDEs. Not much else seems to be known about the detailed behavior of the Lyapunov exponents of the solution to \eqref{SHE}: Ours seem to be the only methods that thus far have succeeded in analyzing the asymptotics of the Lyapunov exponents of the solution to \eqref{SHE} when $\sigma$ is nonlinear and/or $\mathscr{L}$ is nonlocal; Borodin and Corwin have found a few remarkable instances of \eqref{SHE} where all integer-moment Lyapunov exponents can be computed precisely. The second nontrivial contribution of this paper concerns the local behavior of the solution to \eqref{SHE}. The statement is that there frequently exists a monotone function $S$ such that for all $t>0$ fixed, $\eta_\tau$ converges to white noise on $\Z^d$ as $\tau\downarrow 0$, where \begin{equation} \eta_\tau(x) := \frac{ S(u_{t+\tau}(x)) - S(u_t(x)) }{\sqrt\tau} \qquad\mbox{for all $x\in \Z^d$}. \end{equation} See Theorem~\ref{th:local:RadonNikodym}, and especially Theorem~\ref{th:CLT}, for more details. One can interpret our result as saying that the function $S$ is the infinite-dimensional analogue of the scale function for a finite-dimensional diffusion. The function $S$ is in fact also an abstract Hopf--Cole transformation for many nonlinear systems of the form \eqref{SHE}; see \cite{ACQ,BC} for the role of the latter transformation in the continuous-space parabolic Anderson model. Finally, we state and prove perhaps the most interesting contribution of this paper, namely that the solution to \eqref{SHE} is strongly dissipative when \eqref{SHE} is weakly disordered. See Theorem~\ref{th:as} for a precise statement. The latter theorem implies that \eqref{SHE} is a model with hysterisis; see Remark~\ref{rem:Hysteresis} for more details. Theorem~\ref{th:as} is also significant because it rules out the possibility of the existence of an Anderson mobility edge for the present model \eqref{SHE}, when \eqref{SHE} is weakly disordered. We are aware only of one such nonexistence theorem, the original stationary Anderson model (but on ``tree graphs''); see the recent paper by Aizenman and Warzel \cite{AizenmanWarzel}. Among other things, the proof of Theorem~\ref{th:as} relies on a comparison theorem for renewal processes, which we state and prove in the \hyperref[app]{Appendix} [see Lemma~\ref{lem:RE:comparison}]. It is likely that our comparison theorem has other uses in applied probability as well. \section{Results} In this section we present the main results of this paper. Let us begin by making the assertions in the \hyperref[sec1]{Introduction} more precise. \begin{theorem} \label{th:exist:unique} The nonlinear stochastic heat equation \mbox{\eqref{SHE}} has a solution $u$ that is continuous in the variable $t$, and is unique among all predictable random fields that satisfy $\sup_{t\in[0,T]}\sup_{x\in\Z^d}\E(|u_t(x)|^2)<\infty$ for all $T>0$. Moreover, \begin{equation} \label{eq:Upper:LE} \overline{\gamma}_k(u) \le 8\lip^2 k^2 \qquad\mbox{for all integers $k\ge2$}. \end{equation} Furthermore, $u_t(x)\ge0$ for all $t\ge0$ and $x\in\Z^d$ a.s., provided that $u_0(x)\ge0$ for all $x\in\Z^d$. Finally, if $\ell_\sigma>0$ and $\sigma(x)> 0$ for all $x>0$, then for all $\varepsilon\in(0 ,1)$, \begin{equation} \label{eq:Lower:LE} \underline{\gamma}_k(u) \ge(1-\varepsilon) \ell_\sigma^2 k^2 \qquad\mbox{for every integer $k\ge \varepsilon^{-1}+\bigl(\varepsilon\ell _\sigma^2 \bigr)^{-1}$}. \end{equation} \end{theorem} Our next result shows that, at each point $x\in\Z^d$, the solution behaves locally in time like a Brownian motion. Standard moment methods---which we will have to reproduce here as well---show that $t\mapsto u_t(x)$ is almost surely a H\"older-continuous random function for every H\"older exponent $<\nicefrac12$. The following proves that the H\"older exponent $\nicefrac12$ is sharp. \begin{theorem}[(A Radon--Nikod\'ym property)]\label{th:local:RadonNikodym} For every $t\ge0$ and $x\in\Z^d$, \begin{equation} \label{eq:local:RN} \lim_{\tau\downarrow0}\frac{u_{t+\tau}(x) - u_t(x)}{B_{t+\tau}(x) - B_t(x)} = \sigma \bigl( u_t(x) \bigr)\qquad \mbox{in probability}. \end{equation} In addition, \begin{equation}\quad \label{eq:local:LIL} \limsup_{\tau\downarrow0}\frac{u_{t+\tau}(x) - u_t(x)}{ \sqrt{2\tau\log\log(1/\tau)}} =-\liminf _{\tau\downarrow0}\frac{u_{t+\tau}(x) - u_t(x)}{ \sqrt{2\tau\log\log(1/\tau)}} = \bigl\llvert \sigma \bigl( u_t(x) \bigr)\bigr\rrvert , \end{equation} almost surely. \end{theorem} Local iterated logarithm laws, such as \eqref{eq:local:LIL}, are well known in the context of finite-dimensional diffusions; see, for instance, Anderson \cite{Anderson}, Theorem~4.1. The time-change methods employed in the finite-dimensional setting will, however, not work effectively in the present infinite-dimensional context. Here, we obtain~\eqref{eq:local:LIL} as a ready consequence of the proof of the ``random Radon--Nikod\'ym property''~\eqref{eq:local:RN}. \begin{remark} Fix an $x\in\Z^d$ and a $t>0$, and consider the ratio $R(\tau):=[u_{t+\tau}(x)-u_t(x)]/[B_{t+\tau }(x)-B_t(x)]$; this is a well-defined random variable for every $\tau>0$, since $B_{t+\tau}(x)-B_t(x)\neq0$ with probability one for every $\tau>0$. However, $\{R(\tau)\}_{\tau>0}$ is not a well-defined stochastic process since there exist random times $\tau>0$ such that $B_{t+\tau}(x)-B_t(x)=0$ a.s. Thus one does not expect that the mode of convergence in \eqref{eq:local:RN} can be improved to almost-sure convergence. This statement can be strengthened further still, but we will not do so here \end{remark} \begin{remark} According to \eqref{eq:local:RN}, the solution to the \eqref{SHE} behaves as the noninteracting system ``$\d u_t(x)\approx\sigma (u_t(x))\,\d B_t(x)$'' of diffusions, locally to first order. This might seem to suggest the [false] assertion that $x\mapsto u_t(x)$ ought to be a sequence of independent random variables. This is not true, as can be seen by looking more closely at the time increments of $t\mapsto u_t(x)$. In fact, our arguments can be extended to show that the spatial correlation structure of $u$ appears at second-order approximation levels in the sense of the following three-term stochastic Taylor expansion: in the scale $\tau^{1/2}$: \begin{equation} \label{eq:Taylor} u_{t+\tau}(x) \simeq u_t(x) + \tau^{1/2} \sigma\bigl(u_t(x)\bigr) Z_1 +\tau Z_2 +\tau^{3/2} U(\tau) \qquad\mbox{as $\tau\downarrow0$}, \end{equation} where: (i) ``$\simeq$'' denotes approximation of distributions; (ii) $Z_1$ is a standard normal variable independent of $u_t(x)$; (iii) $Z_2$ is a nontrivial random variable that depends on the entire random field $\{u_s(y)\}_{s\in[0,t],y\in\Z^d}$; (iv) $U(\tau )=O_\P(1)$ as $\tau\downarrow0$ means that $\lim_{m\uparrow\infty}\limsup_{\tau\downarrow0}\P\{|U(\tau )|\ge m\}=0$. In particular, \eqref{eq:Taylor} tells us that the temporally-local interactions in the random field $x\mapsto u_t(x)$ are second order in nature. \end{remark} Rather than prove these refined assertions, we next turn our attention to a different local property of the solution to \eqref{SHE} and show that, after a scale change, the local-in-time behavior of the solution to \eqref{SHE} is that of spatial white noise. \begin{theorem}\label{th:CLT} Suppose $\sigma(z)>0$ for all $z\in\R\setminus\{0\}$, and define \begin{equation} S(z) := \int_{z_0}^z \frac{\d w}{\sigma(w)} \qquad(z\ge0), \end{equation} where $z_0\in\R\setminus\{0\}$ is a fixed number. Then, $S(u_t(x))<\infty$ a.s. for all $t>0$ and $x\in\Z^d$. Furthermore, if we choose and fix $m$ distinct points $x_1,\ldots,x_m\in\Z^d$, then for all $t> 0$ and $q_1,\ldots,q_m\in\R$, \begin{equation} \lim_{\tau\downarrow0} \P \Biggl(\bigcap_{j=1}^m \bigl\{ S \bigl( u_{t+\tau}(x_j) \bigr) - S \bigl( u_t(x_j) \bigr) \le q_j\sqrt\tau \bigr\} \Biggr) = \prod_{j=1}^ m \Phi(q_j), \end{equation} where $\Phi(q) :=(2\pi)^{-1/2}\int_{-\infty}^q \exp(-w^2/2) \,\d w$ denotes the standard Gaussian cumulative distribution function. \end{theorem} The preceding manifests itself in interesting ways for different choices of the nonlinearity coefficient $\sigma$. Let us mention the following parabolic Anderson model, which has been a motivating example for us. \begin{example} Consider the semi-discrete parabolic Anderson model, which is \eqref{SHE} with $\sigma(x)\equiv qx$ [for some fixed constant $q> 0$]. In that case, the solution to \eqref{SHE} is positive and the ``scale function'' $S$ is $S(z) = q^{-1}\ln(z/z_0)$ for $z,z_0>0$. As such, $\sigma(u_t(x))=qu_t(x)$ in \eqref{SHE}, and we find the following log-normal limit law: For every $t>0$ and $x_1,\ldots,x_m\in\Z^d$ fixed, \begin{equation}\qquad \biggl( \biggl[\frac{u_{t+\tau}(x_1)}{u_t(x_1)} \biggr]^{1/\sqrt{\tau}},\ldots, \biggl[ \frac{u_{t+\tau}(x_m)}{u_t(x_m)} \biggr]^{1/\sqrt{\tau}} \biggr) \Rightarrow\bigl({\mathrm{e}}^{qN_1},\ldots,{\mathrm{e}}^{qN_m}\bigr) \qquad\tau \downarrow0, \end{equation} where $N_1 ,\ldots,N_m$ are i.i.d. standard normal variables, and ``$\Rightarrow$'' denotes convergence in distribution. \end{example} Our final main result is a statement about the large-time behavior of the solution $u$ to \eqref{SHE}. We prove a rigorous version of the following assertion: ``\emph{If the random walk $X$ is transient and $\lip$ is sufficiently small---so that \textup{\eqref{SHE}} is not very noisy---then a decay condition such as $u_0\in\ell^1(\Z^d)$ on the initial profile is enough to ensure that $\sup_{x\in\Z^d}|u_t(x)|\to0$ almost surely as $t\to\infty$.}'' This is new even for the parabolic Anderson model, where $\sigma(x)\propto x$ and $\mathscr{L}:=$ the generator of the simple walk on $\Z^d$. In fact, this result gives a partial [though strong] negative answer to an open problem of Carmona and Molchanov \cite{CM94}, page 122, and rules out the existence of [the analogue of] a nontrivial ``Anderson mobility edge'' in the present nonstationary setting, when $u_0\in\ell^1(\Z^d)$. Recall that $X:=\{X_t\}_{t\ge0}$ is a continuous-time random walk on $\Z^d$ with generator $\mathscr{L}$. Let $X'$ denote an independent copy of $X$, and define \begin{equation} \label{Upsilon(0)} \Upsilon(0) := \int_0^\infty\P \bigl\{X_t=X_t'\bigr\} \,\d t = \E\int _0^\infty \1 _{\{0\}} \bigl(X_t-X_t'\bigr) \,\d t. \end{equation} We can think of $\Upsilon(0)$ as the expected value of the total occupation time of $\{0\}$, as viewed by the symmetrized random walk $X-X'$. Although $\Upsilon(0)$ is always well defined, it is finite if and only if the symmetrized random walk $X-X'$ is transient~\cite{ChungFuchs}. We are ready to state our final result. \begin{theorem}[(Dissipation)]\label{th:as} Suppose that \begin{equation} \label{cond:zk} \lip< \bigl[\Upsilon(0) \bigr]^{-1/2}, \end{equation} and that there exists $\alpha\in(1 ,\infty)$ such that \begin{equation} \label{cond:alpha} \P\bigl\{X_t=X'_t\bigr\}= O\bigl(t^{-\alpha}\bigr) \qquad(t\to\infty), \end{equation} where $X'$ denotes an independent copy of $X$. If, in addition, $u_0\in\ell^1(\Z^d)$ and the underlying probability space is complete, then \begin{equation} \lim_{t\to\infty}\sup_{x\in\Z^d}\bigl|u_t(x)\bigr|= \lim_{t\to\infty}\sum_{x\in\Z^d}\bigl|u_t(x)\bigr|^2 = 0\qquad \mbox{almost surely}. \end{equation} \end{theorem} \begin{remark}[(Hysteresis)]\label{rem:Hysteresis} Consider the parabolic Anderson model\break [$\sigma(x)\propto x$], where the underlying symmetrized walk $X-X'$ is transient, the noise level is small and $u_0$ is a constant. It is well known that under these conditions $u_t(x)$ converges weakly as $t\to\infty$ to a nondegenerate random variable $u_\infty(x)$ for every $x\in\Z^d$. See, for example, Greven and den Hollander \cite{GrevenDenHollander}, Theorem~1.4, Cox and Greven \cite{CoxGreven}, and Shiga \cite{Shiga}. These results provide a partial affirmative answer to a question of Carmona and Molchanov \cite{CM94}, page 122, about the existence of long-term invariant laws in the low-noise regime of the transient parabolic Anderson model, in particular. By contrast, Theorem~\ref{th:as} shows that if $u_0$ is far from stationary (here, it decays at infinity), then the system is very strongly dissipative in the low-noise regime. This result implies that \emph{the parabolic Anderson model remembers its initial state forever} \end{remark} \begin{example} Continuous-time walks that have property \eqref{cond:alpha} include all transient finite-variance centered random walks on $\Z^d$ [$d>2$, necessarily]. For those walks, $\alpha:=d/2$, thanks to the local central limit theorem. There are more interesting examples as well. For instance, suppose $t^{-1/p}X_t$ converges in distribution to a stable random variable $S$ as $t\to\infty$; see Gnedenko and Kolmogorov~\cite{GK}, Section~35, for necessary and sufficient conditions. Then $S$ is necessarily stable with index $p$, $p\in(0 ,2]$, and $t^{-1/p}(X_t-X_t')$ converges in law to a symmetric stable random variable $S$ with stability index $p$. If, in addition, the group of all possible values of $X_t-X_t'$ generates all of $\Z^d$, then a theorem of Gnedenko \cite{GK}, page 236, ensures that $t^{1/p}\P\{X_t=X_t'\}$ converges to $f(0)<\infty$, where $f$ denotes the probability density function of $S$, as long as $p\in(0 ,1)$. \end{example} \textit{Organization of the paper}. In Section~\ref{sec:prelim} we introduce the mild solution to \eqref{SHE} and state the version of Burkholder--Davis--Gundy inequality that we will use throughout the paper. In Section~\ref{sec:thm1.1} we prove Theorem~\ref{th:exist:unique}: Section~\ref{subsec:upper lyap} includes proof of upper bounds for the Lyapunov exponents, from which existence of the solution follows; Section~\ref{subsec:lower lyap} contains a comparison principle for \eqref{SHE}, together with proof of lower bounds for the Lyapunov exponents. Section~\ref{sec:approx} has some results on the local (in time) behavior of the solution. These results are used in Section~\ref{sec:thm1.2} in order to prove Theorem~\ref {th:local:RadonNikodym}. Section~\ref{sec:thm1.4} contains a proof of Theorem~\ref{th:CLT}. That proof hinges on Theorem~\ref{th:local:RadonNikodym} and the fact that the solution to \eqref{SHE} immediately becomes strictly positive everywhere (Proposition~\ref {prop:pos}). Section~\ref{sec:prelim thm1.6} explains how $\|u_t\|^2_{\ell^{2}(\Z^d)}$ is connected to the intersection local times of two independent continuous time random walks with common generator $\mathscr{L}$. The results of Section~\ref{sec:prelim thm1.6} are then used in Section~9 in order to prove Theorem~\ref{th:as}. \section{Preliminaries} \label{sec:prelim} In this small section we collect some preliminary facts about SPDEs interacting diffusion processes and BDG-type martingale inequalities. These facts are used throughout the rest of the paper. \subsection{The mild solution} \label{subsec:mild} Recall that the convolution on $\Z^d$ is defined by \begin{equation} (f*g) (x) := \sum_{y\in\Z^d} f(x-y)g(y) \qquad\bigl(x\in \Z^d\bigr). \end{equation} For every function $h\dvtx\Z^d\to\R$ we define a new function $\tilde{h}$, \begin{equation} \tilde{h}(x) := h(-x)\qquad \bigl(x\in\Z^d\bigr), \end{equation} as the \emph{reflection} of $h$. By a ``solution'' to \eqref{SHE} we mean a solution in integrated-- or ``mild''---form. That is, a predictable process $t\mapsto u_t$, with values in $\R^{\Z^d}$, that solves the following infinite system of It\^o SDEs: \begin{equation} \label{mild} u_t(x) = (\tilde{p}_t*u_0) (x) + \sum_{y\in\Z^d}\int_0^t p_{t-s}(y-x)\sigma\bigl(u_s(y)\bigr) \,\d B_s(y), \end{equation} where $p_t(x) := \P\{X_t=x\}$. It might be helpful to note also that $(P_t\phi)(x) := (\tilde{p}_t*\phi )(x)$ defines the semigroup of the random walk $X$ via the identity $(P_t\phi)(x) = \E\phi(x+X_t)$. Thus we can write \eqref{mild} in the following, perhaps more familar, form: \begin{equation} u_t(x) = (P_tu_0) (x) + \sum _{y\in\Z^d}\int_0^t p_{t-s}(y-x)\sigma\bigl(u_s(y)\bigr) \,\d B_s(y). \end{equation} \subsection{A BDG inequality} \label{subsec:BDG} We begin this subsection with some background on Burkholder's constants which will give us the best constants in Lemma~\ref{lem:BDG}. According to the Burkholder--Davis--Gundy inequality \cite{Burkholder,BDG,BG}, \begin{equation} z_p := \sup_x \sup_{t>0} \biggl[\frac{\E (|x_t|^p )}{ \E (\langle x\rangle_t^{p/2} )} \biggr]^{1/p}<\infty, \end{equation} where the supremum ``$\sup_x$'' is taken over all nonzero martinagles $x:=\{x_t\}_{t\ge0}$ that have continuous trajectories and are in $L^2(\P)$ at all times, $\langle x\rangle_t$ denotes the quadratic variation of $x$ at time $t$ and $0/0:=\infty/\infty:=0$. Davis \cite{Davis} has computed the numerical value of $z_p$ in terms of zeroes of special functions. When $p\ge2$ an integer, Davis's theorem implies that $z_p$ is equal to the largest positive root of the modified Hermite polynomial $\mbox{\sl He}_p$. Thus, for example, we obtain the following from direct evaluation of the zeros: \begin{eqnarray} z_2 &= &1, \qquad z_3=\sqrt3, \qquad z_4= \sqrt{3+\sqrt{6}} \approx2.334, \nonumber \\[-8pt] \\[-8pt] \nonumber z_5&=&\sqrt{5+\sqrt{10}}\approx2.857,\qquad z_6 \approx3.324,\ldots. \end{eqnarray} It is known that $ z_p\sim2\sqrt p$ as $p\to\infty$, and $\sup_{p\ge2} ( z_p/\sqrt p)=2$; see Carlen and Kr{\'e}e \cite{CK}, Appendix Suppose $Z:=\{Z_t(x)\}_{t\ge0,x\in\Z^d}$ is a predictable random field, with respect to the infinite-dimensional Brownian motion $\{B_t(\bullet)\}_{t\ge0}$, that satisfies the moment bound $\E\int_0^t\|Z_s\|_{\ell^2(\Z^d)}^2 \,\d s<\infty$. Then the It\^o integral process defined by \begin{equation} \int_0^t Z_s\cdot\,\d B_s:=\sum_{y\in\Z^d}\int _0^t Z_s(y) \,\d B_s(y)\qquad (t\ge0) \end{equation} exists and defines a continuous $L^2(\P)$ martingale. See, for example, Pr{\'e}v{\^o}t and R{\"o}ckner \cite{PrevotRoeckner}. The following variation of the Burkholder--Davis--Gundy inequality yields moment bounds for this martingale. \begin{lemma}[(BDG lemma)]\label{lem:BDG} For all finite real numbers $k\ge2$ and $t\ge0$, \begin{equation} \label{BDG} \E \biggl(\biggl\llvert \int_0^t Z_s\cdot\,\d B_s \biggr\rrvert ^k \biggr) \le \biggl|4k\sum_{y\in\Z^d}\int_0^t \bigl\{ \E \bigl( \bigl|Z_s(y)\bigr|^k \bigr) \bigr \}^{2/k}\,\d s \biggr|^{k/2}. \end{equation} \end{lemma} \begin{pf} We follow a method of Foondun and Khoshnevisan \cite{FK}. A standard approximation argument tells us that it suffices to consider the case where $y\mapsto Z_s(y)$ has finite support. Let $F\subset\Z^d$ be a finite set of cardinality \mbox{$m\ge1$}, and suppose $Z_s(y)=0$ for all $y\notin F$. Consider the (standard, finite-dimensional) It\^o integral process $\int_0^t Z_s\cdot\,\d B_s := \sum_{y\in F}\int_0^t Z_s(y) \,\d B_s(y)$. According to Davis's \cite{Davis} form of the Burkholder--Davis--Gundy inequality $m$-dimensional Brownian motion \cite{Burkholder,BDG,BG}, \begin{equation} \label{eq:Davis} \E \biggl( \biggl\llvert \int_0^t Z_s\cdot\,\d B_s\biggr\rrvert ^k \biggr) \le z_k^k \E \biggl( \biggl| \sum_{y\in F} \int_0^t \bigl[Z_s(y) \bigr]^2 \,\d s \biggr|^{k/2} \biggr). \end{equation} Finally, we use the Carlen--Kr{\'e}e bound $z_k\le2\sqrt k$ \cite{CK} together with the Minkowski inequality to finish the proof in the case where $F$ is finite. A standard finite-dimensional approximation completes the proof. \end{pf} \section{Proof of Theorem \texorpdfstring{\protect\ref{th:exist:unique}}{2.1}} \label{sec:thm1.1} \subsection{Bounds for the upper Lyapunov exponents} \label {subsec:upper lyap} Existence and uniqueness, and also continuity, of the solution are dealt with extensively in the literature and are well known; see, for example, Shiga and Shimizu \cite{ShigaShimizu} and the general theory of Pr{\'e}v{\^o}t and R{\"o}ckner \cite{PrevotRoeckner} for some of the latest developments. However, in order to derive our estimates of the Lyapunov exponents we will need a priori estimates which will also yield existence and uniqueness. Therefore, in this section, we hash out some---though not all---of the details. Let us proceed by applying Picard iteration. Let $u^{(0)}_t(x):=u_0(x)$, and then define iteratively for all $n\ge0$, \begin{equation} \label{eq:Picard} u^{(n+1)}_t(x) := (\tilde{p}_t*u_0) (x) + \sum_{y\in\Z^d}\int_0^t p_{t-s}(y-x)\sigma \bigl( u^{(n)}_s(y) \bigr) \,\d B_s(y). \end{equation} It follows from the properties of the It\^o integral that \begin{equation} \label{eq:Mn} M_t^{(n+1)}:=\sup_{x\in\Z^d} \E \bigl(\bigl| u_t^{(n+1)}(x) \bigr|^k \bigr) \le2^{k-1}\sup_{x\in\Z^d}(I_x+J_x), \end{equation} where \begin{eqnarray} \label{eq:IJ} I_x&:=& \bigl\llvert (\tilde p_t*u_0) (x) \bigr\rrvert ^k, \nonumber \\[-8pt] \\[-8pt] \nonumber J_x&:=&\E \biggl( \biggl| \sum_{y\in\Z^d}\int _0^t p_{t-s}(y-x) \sigma \bigl( u_s^{(n)}(y) \bigr) \,\d B_s(y) \biggr|^k \biggr). \end{eqnarray} The first term is easy to bound: \begin{equation} \label{eq:Ix} \sup_{x\in\Z^d} I_x \le \|u_0\|_{\ell^\infty(\Z^d)}^k, \end{equation} since $\sum_xp_t(x)= 1$. Next we bound $J_x$. Because $\sigma$ is Lipschitz continuous and $\sigma(0)=0$, we can see that $|\sigma(z)|\le\lip|z|$ for all $z\in\R$. Thus we may use the BDG lemma (Lemma~\ref{lem:BDG}) in order to see that \begin{equation} \label{eq:Jx:prec} J_x^{2/k} \le4k\lip^2\sum _{y\in\Z^d}\int_0^t \bigl[p_{t-s}(y-x)\bigr]^2 \bigl\{ \E \bigl( \bigl\llvert u^{(n)}_s(y)\bigr\rrvert ^k \bigr) \bigr \}^{2/k} \,\d s. \end{equation} Therefore, we may recall the inductive definition \eqref{eq:Mn} of $M$ to see that \begin{eqnarray} \label{eq:Jx} J_x^{2/k} &\le&4k \lip^2\sum_{y\in\Z^d}\int_0^t \bigl[p_{t-s}(y-x)\bigr]^2 \bigl( M^{(n)}_s \bigr)^{2/k} \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&4k\lip^2\int_0^t \bigl( M^{(n)}_s \bigr)^{2/k} \,\d s, \end{eqnarray} since \begin{equation} \label{eq:replica} \sum_{z\in\Z^d}\bigl[p_r(z) \bigr]^2=\P\bigl\{X_r=X_r' \bigr\}\le1, \end{equation} where $X'$ denotes an independent copy of $X$. (This last bound might appear to be quite crude, and it is when $r$ is large. However, it turns out that the behavior of $r$ near zero matters more to us. Therefore, the inequality is tight in the regime $r\approx0$ of interest to us.) We may combine \eqref{eq:Mn}, \eqref{eq:Ix} and \eqref{eq:Jx} in order to see that for all $\beta,t>0$, \begin{eqnarray}\quad &&{\mathrm{e}}^{-\beta t}M^{(n+1)}_t \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le2^{k-1}\|u_0\|_{\ell^\infty(\Z^d)}^k +\bigl(16k\lip^2\bigr)^{k/2}\biggl\llvert \int _0^t{\mathrm{e}}^{-2\beta(t-s)/k} \bigl( {\mathrm{e}}^{-\beta s}M^{(n)}_s \bigr)^{2/k} \,\d s \biggr\rrvert ^{k/2}. \end{eqnarray} Consequently, the sequence defined by \begin{equation} N_\beta^{(m)}:= \sup_{t\ge0} \bigl( {\mathrm{e}}^{-\beta t}M^{(m)}_t \bigr)\qquad (m\ge0) \end{equation} satisfies the recursive inequality \begin{eqnarray} N_\beta^{(n+1)} &\le&2^{k-1} \|u_0\|_{\ell^\infty(\Z^d)}^k + \bigl(16k\lip^2 \bigr)^{k/2}\biggl\llvert \int_0^t{\mathrm{e}}^{-2\beta s/k} \,\d s \biggr\rrvert ^{k/2}N_\beta^{(n)} \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&2^{k-1}\|u_0\|_{\ell^\infty(\Z^d)}^k + \biggl(\frac{8 k^2\lip^2}{\beta} \biggr)^{k/2}N_\beta^{(n)}. \end{eqnarray} In particular, if we denote (temporarily for this proof) \begin{equation} \alpha:= 8(1+\delta)\lip^2, \end{equation} where $\delta>0$ is fixed but arbitrary, then \begin{equation} N_{\alpha k^2}^{(n+1)} \le2^{k-1}\|u_0 \|_{\ell^\infty(\Z^d)}^k + (1+\delta)^{-k/2}N_{\alpha k^2}^{(n)}. \end{equation} We may apply induction on $n$ now in order to see that $\sup_{n\ge0}N_{\alpha k^2}^{(n)}<\infty$; equivalently, for all $k\ge2$ there exists $c_k\in(0 ,\infty)$ such that \begin{equation} \label{eq:u:UB} \sup_{x\in\Z^d}\E \bigl( \bigl\llvert u^{(n)}_t(x) \bigr\rrvert ^k \bigr) \le c_k {\mathrm{e}}^{8(1+\delta)\lip^2 k^2t} \qquad\mbox{for all $t\ge0$}. \end{equation} Similarly, \begin{eqnarray} &&\E \bigl( \bigl| u^{(n+1)}_t(x) - u^{(n)}_t(x) \bigr|^k \bigr)\nonumber \\ &&\qquad=\E \biggl( \biggl| \sum_{y\in\Z^d}\int _0^t p_{t-s}(y-x) \bigl\{ \sigma \bigl( u^{(n)}_s(y) \bigr) - \sigma \bigl( u^{(n-1)}_s(y) \bigr) \bigr\} \,\d B_s(y) \biggr|^k \biggr) \\ \nonumber &&\qquad\le\bigl(4k\lip^2\bigr)^{k/2} \E \biggl( \biggl| \sum _{y\in\Z^d}\int_0^t \bigl[ p_{t-s}(y-x) \bigr]^2 \bigl\{ u^{(n)}_s(y) - u^{(n-1)}_s(y) \bigr\}^2 \,\d s\biggr |^{k/2} \biggr).\hspace*{-12pt} \end{eqnarray} Define \begin{equation} L_t^{(n+1)}:= \sup_{x\in\Z^d} \E \bigl( \bigl \llvert u^{(n+1)}_t(x) - u^{(n)}_t(x) \bigr\rrvert ^k \bigr) \end{equation} to deduce from the preceding, \eqref{eq:replica} and Minkowski's inequality that \begin{eqnarray} L^{(n+1)}_t &\le&\bigl(4k\lip^2 \bigr)^{k/2} \biggl( \sum_{y\in\Z^d}\int _0^t \bigl[ p_{t-s}(y-x) \bigr]^2 \bigl( L_s^{(n)} \bigr)^{2/k} \,\d s \biggr)^{k/2} \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&\bigl(4k\lip^2\bigr)^{k/2} \biggl( \int _0^t \bigl(L_s^{(n)} \bigr)^{2/k} \,\d s \biggr)^{k/2}. \end{eqnarray} Therefore, \begin{equation} K_{\alpha k^2}^{(m)}= \sup_{t\ge0} \bigl( {\mathrm{e}}^{-\alpha k^2 t} L_t^{(m)} \bigr) \end{equation} satisfies \begin{eqnarray} \nonumber K_{\alpha k^2}^{(n+1)} &\le&\bigl(4k\lip^2 \bigr)^{k/2} \biggl( \int_0^t{\mathrm{e}} ^{-2\alpha k(t-s)} \,\d s \biggr)^{k/2} K_{\alpha k^2}^{(n)}\\ &\le& \biggl( \frac{4\lip^2} {2\alpha} \biggr)^{k/2} K_{\alpha k^2}^{(n)} \\ &\le&2^{-k}K_{\alpha k^2}^{(n)}.\nonumber \end{eqnarray} From this we conclude that $\sum_{n=0}^\infty K_{\alpha k^2}^{(n)}< \infty$. Therefore, there exists a random field $u_t(x)$ such that $\lim_{n\to\infty}u_t^{(n)}(x)=u_t(x)$ in $L^k(\P)$. It follows readily that $u$ solves \eqref{SHE}, and $u$ satisfies \eqref {eq:Upper:LE} by \eqref{eq:u:UB} and Fatou's lemma. Uniqueness is proved by similar means, and we skip the details. \subsection{Bounds for the lower Lyapunov exponents} \label {subsec:lower lyap} We start the section with a truncation error estimate for the nonlinearity $\sigma$. This will be needed to use the results of Cox, Fleischmann and Greven \cite{CFG} on comparison of moments for interacting diffusions. We can then reduce our problem to the case of $\sigma(x)=\ell_\sigma x$. \begin{lemma}\label{lem:truncate} Define $\sigma^{(N)}$ by $\sigma^{(N)}:=\sigma$ on $(-N ,N)$, $\sigma^{(N)}:=0$ on $[-N-1 ,N+1]^c$, and defined by linear interpolation on $[-N-1 ,-N]\cup[N ,N+1]$. Let $U^{(N)}_t(x)$ denote the a.s.-unique solution to \mbox{\eqref{SHE}} where $\sigma$ is replaced by $\sigma^{(N)}$. Then, $\lim_{N\to\infty} U^{(N)}_t(x)=u_t(x)$ a.s. and in $L^k(\P)$ for all $k\ge2$, $t\ge0$ and $x\in\Z^d$. \end{lemma} \begin{pf} Since $\sigma^{(N)}$ is Lipschitz continuous, Theorem~\ref{th:exist:unique} ensures the existence and uniqueness of $U^{(N)}$ for every $N\ge1$. Then by \eqref{mild} \begin{equation} u_t(x) - U^{(N)}_t(x) = T_1 + T_2, \end{equation} where \begin{eqnarray} T_1 &:=&\sum_{y\in\Z^d}\int _0^t p_{t-s}(y-x) \bigl\lbrace \sigma\bigl( u_s(y)\bigr)-\sigma^{(N)}(u_s(y) \bigr\rbrace\,\d B_s(y); \nonumber \\[-8pt] \\[-8pt] \nonumber T_2 &:=& \sum_{y\in\Z^d}\int _0^t p_{t-s}(y-x) \bigl\{ \sigma^{(N)} \bigl( u_s(y)\bigr) - \sigma^{(N)} \bigl( U^{(N)}_s(y) \bigr) \bigr\} \,\d B_s(y). \end{eqnarray} Because $|\sigma(z)|\le\lip|z|$, Lemma~\ref{lem:BDG} implies that $ \{\E(|T_1|^k) \}^{2/k}$ is at most \begin{equation} 4k\lip^2\sum_{y\in\Z^d}\int _0^t \bigl[p_{t-s}(y-x) \bigr]^2 \bigl\{ \E \bigl(\bigl\llvert u_s(y)\bigr \rrvert ^k; \bigl|u_s(y)\bigr|\ge N \bigr) \bigr\}^{2/k} \,\d s. \end{equation} We have $\E( |Y|^k; |Y|\ge N ) \le N^{-k}\E(Y^{2k} )$, valid for all $Y\in L^{2k}(\Omega)$. Therefore, \begin{equation}\quad \bigl\{\E\bigl(|T_1|^k\bigr) \bigr\}^{2/k}\le \frac{4k\lip^2}{N^2} \sum_{y\in\Z^d}\int _0^t \bigl[p_{t-s}(y-x) \bigr]^2 \bigl\{ \E \bigl(\bigl\llvert u_s(y)\bigr \rrvert ^{2k} \bigr) \bigr\}^{2/k} \,\d s. \end{equation} Because $\sum_{y\in\Z^d}[p_{t-s}(y-x)]^2\le1$---see \eqref {eq:replica}-- the already-proved bound \eqref{eq:Upper:LE} tells us that \begin{equation} \label{eq:T_1} \bigl\{\E\bigl(|T_1|^k\bigr) \bigr \}^{2/k} \le\frac{a_k}{N^2} \int_0^t{\mathrm{e}}^{128\lip^2 k s} \,\d s \le\frac{Aa_k{\mathrm{e}}^{A kt}}{N^2}, \end{equation} where $a_k$ and $A$ are uninteresting finite and positive constants; moreover, $a_k$ depends only on $k$. This estimates the norm of $T_1$. As for $T_2$, we use the simple inequality $|\sigma^{(N)}(r)-\sigma^{(N)}(\rho)|\le C |r-\rho|$, together with the BDG Lemma~\ref{lem:BDG} in order to find that \begin{equation} \bigl\{ \E \bigl(|T_2|^k \bigr) \bigr \}^{2/k} \le b_k\int_0^t \sup_{y\in\Z^d} \bigl\{\E \bigl(\bigl\llvert u_s(y)- U^{(N)}_s(y)\bigr\rrvert ^k \bigr) \bigr \}^{2/k} \,\d s, \end{equation} where $b_k$ is a constant dependent on $\sigma$ and $k$. Together, the preceding moment bounds for $T_1$ and $T_2$ imply that \begin{equation} D^{(N)}_t := \sup_{x\in\Z^d} \bigl\{\E \bigl(\bigl\llvert u_t(x)-U^{(N)}_t(x)\bigr \rrvert ^k \bigr) \bigr\}^{2/k} \end{equation} satisfies the recursion \begin{equation} D^{(N)}_t \le\frac{\tilde{a}_k{\mathrm{e}}^{\tilde{A}k^2t}}{N^2} + \tilde{b}_k \int_0^t D^{(N)}_s \,\d s, \end{equation} where $\tilde{a}_k$, $\tilde{b}_k$ and $\tilde{A}$ are positive and finite constants, and the first two depend only on $k$ (whereas the latter is universal). An application of the Gronwall inequality shows that $\sup_{t\in[0,T]}D^{(N)}_t = O(N^{-2})$ as $N\to\infty$, for every fixed value $T\in(0 ,\infty)$. This is enough to yield the lemma. \end{pf} We complete the proof of Theorem~\ref{th:exist:unique} by verifying the two remaining assertions of that theorem: (i) The solution is nonnegative because $u_0(x)\ge0$ and $\sigma(0)=0$; and (ii) The lower bound \eqref{eq:Lower:LE} for the lower Lyapunov exponent holds. We keep the two parts separate, as they use different ideas. \begin{theorem}[(Comparison principle)]\label{th:comparison} Suppose $u$ and $v$ are the solutions to \mbox{\eqref{SHE}} with respective initial functions $u_0$ and $v_0$. If $u_0(x)\ge v_0(x)$ for all $x\in\Z^d$, then $u_t(x)\ge v_t(x)$ for all $t\ge0$ and $x\in\Z^d$ a.s. \end{theorem} The nonnegativity assertion of Theorem~\ref{th:exist:unique} is well known \cite{Shiga}, but also follows from the preceding comparison principle. This is because condition \eqref{eq:sigma(0)=0} implies that $v_t(x)\equiv0$ is the unique solution to \eqref{SHE} with initial condition $v_0(x)\equiv0$. Therefore, the comparison principle yields $u_t(x)\ge v_t(x)=0$ a.s. \begin{pf*}{Proof of Theorem~\ref{th:comparison}} Consider the following infinite dimensional SDE: \begin{eqnarray}\label{eq:infinitedimSHE} \qquad w_t(x)=w_0(x)+\int_{0}^{t}( \mathscr{L}w_s) (x) \,\d s + \int_{0}^{t} \sigma\bigl(w_s(x)\bigr)\,\d B_s(x) \nonumber \\[-8pt] \\[-8pt] \eqntext{\bigl(x\in \Z^d\bigr).} \end{eqnarray} It is a well-known fact that the mild solution to \eqref{SHE} is also a solution in the weak sense. See, for example, Theorem~3.1 of Iwata \cite{Iwata} and its proof. Therefore, $u_t(x)$ and $v_t(x)$, respectively, solve \eqref{eq:infinitedimSHE} with initial conditions $u_0(x)$ and $v_0(x)$. Let $\{S_n\}_{n=1}^\infty$ denote a growing sequence of finite subsets of $\Z^d$ that exhaust all of $\Z^d$. Consider, for every $n\ge1$, the stochastic integral equation, \begin{eqnarray} \cases{\displaystyle u_t^{(n)}(x)=u_0(x)+\int _{0}^{t}\bigl(\mathscr{L}u_s^{(n)} \bigr) (x) \,\d s\vspace*{2pt}\cr {}\hspace*{42pt} +\displaystyle \int_{0}^{t}\sigma \bigl(u_s^{(n)}(x)\bigr)\,\d B_s(x),&\quad $\mbox{if $x\in S_n$}$;\vspace*{2pt} \cr u_t^{(n)}(x)=u_0(x),&\quad $ \mbox{if $x\notin S_n$}$.} \end{eqnarray} Similarly, we let $v^{(n)}$ solve the same equation, but start it from $v_0(x)$. Each of these equations is in fact a finite-dimensional SDE, and has a unique strong solution, by It\^o's theory. Moreover, Shiga and Shimizu's proof of their Theorem~2.1 \cite{ShigaShimizu} shows that, for every $x\in\Z^d$ and $t>0$, there exists a subsequence $\{n_k\}_{k=1}^\infty$ of increasing integers such that \begin{equation} u^{(n_k)}_t(x) \stackrel{\P} {\longrightarrow}u_t(x) \quad\mbox{and}\quad v^{(n_k)}_t(x) \stackrel{\P} { \longrightarrow}v_t(x), \end{equation} as $k\to\infty$. Therefore, we may appeal to a comparison principle for finite-dimensional SDEs, such as that of Geiss and Manthey \cite{GM}, Theorem~1.2, in order to conclude the result; the quasi-monotonicity condition of \cite{GM} is met simply because $\mathscr{L}$ is the generator of a Markov chain. The verification of that detail is left to the interested reader. \end{pf*} We are now in position to establish the lower bound \eqref{eq:Lower:LE} on the lower Lyapunov exponent of the solution to \eqref{SHE}. \begin{pf*}{Proof of Theorem~\ref{th:exist:unique}: Verification of \eqref{eq:Lower:LE}} Let $v$ solve the stochastic heat equation \begin{equation} \d v_t(x) = (\mathscr{L}v_t) (x) \,\d t + \ell_\sigma v_t(x) \,\d B_t(x), \end{equation} subject to $v_0(x) := u_0(x)$. Also define $V^{(N)}$ to be the solution to \begin{equation} \d V^{(N)}_t(x) = (\mathscr{L}v_t) (x) \,\d t + \zeta^{(N)} \bigl(V^{(N)}_t(x) \bigr) \,\d B_t(x), \end{equation} where $\zeta^{(N)}(x) :=\ell_\sigma x$ on $(-N ,N)$, $\zeta^{(N)}(x):=0$ when $|x|\ge N+1$, and $\zeta^{(N)}$ is defined by linear interpolation everywhere else. Define $\sigma^{(N)}$ and $U^{(N)}$ as in Lemma~\ref{lem:truncate}. Because $\sigma^{(N)}\ge\zeta^{(N)}$ everywhere on $\R_+$, and since both $U^{(N)}$ and $V^{(N)}$ are $\ge0$ a.s. and pointwise, the comparison theorem of Cox, Fleischmann and Greven \cite{CFG}, Theorem~1, shows us that \begin{equation} \E \bigl(\bigl\llvert V^{(N)}_t(x)\bigr\rrvert ^k \bigr) \le\E \bigl( \bigl\llvert U^{(N)}_t(x) \bigr\rrvert ^k \bigr), \end{equation} for all $t\ge0$, $x\in\Z^d$, $k\ge2$ and $N\ge1$. Let $N\to\infty$, and apply Lemma~\ref{lem:truncate} to find that $V^{(N)}_t(x)\to v_t(x)$ and $U^{(N)}_t(x)\to u_t(x)$ in $L^k(\P)$ for all $k\ge2$. As a result, one can let $N\to\infty$ in the preceding display in order to deduce the following: \begin{equation} \E \bigl(\bigl\llvert v_t(x)\bigr\rrvert ^k \bigr) \le \E \bigl( \bigl\llvert u_t(x) \bigr\rrvert ^k \bigr). \end{equation} Therefore, it remains to bound $\underline{\gamma}_k(v)$ from below. Let $\{X^{(i)}\}_{i=1}^k$ denote $k$ independent copies of the random walk $X$. It is possible to prove that \begin{equation} \label{eq:conus} \E \bigl( \bigl|v_t(x)\bigr|^k \bigr) = \E \Biggl( \prod_{j=1}^k u_0 \bigl( X^{(j)}_t+x \bigr) \,\cdot\,{\mathrm{e}}^{M_k(t)} \Biggr), \end{equation} where $M_k(t)$ denotes the ``multiple collision local time,'' \begin{equation} M_k(t) := 2\ell_\sigma^2 \mathop{\sum \sum}_{1\le i<j\le k} \int_0^t \1_{\{0\}} \bigl( X^{(i)}_s-X^{(j)}_s \bigr) \,\d s. \end{equation} When $X$ is the continuous-time simple random walk on $\Z^d$, this is a well-known consequence of a Feynman--Kac formula; see, for instance, Carmona and Molchanov \cite{CM94}, page 19. When $X$ is replaced by a L\'evy process, Conus \cite{Conus} has found an elegant derivation of this formula. The class of all L\'evy processes includes that of continuous-time random walks, whence follows \eqref{eq:conus}. Note that a.s. on the event that none of the walks $X^{(1)},\ldots,X^{(k)}$ jump in the time interval $[0 ,t]$, \begin{equation} \prod_{j=1}^k u_0 \bigl( X^{(j)}+x \bigr) {\mathrm{e}}^ M_k(t)} \ge\bigl[u_0(x) \bigr]^k {\mathrm{e}}^{k(k-1)\ell_\sigma^2 t}. \end{equation} Since the probability is $\exp(-t)$ that $X^{(j)}$ does not jump in $[0 ,t]$, it follows from the independence of $X^{(1)},\ldots,X^{(k)}$ that \begin{equation} \E \bigl(\bigl|v_t(x)\bigr|^k \bigr) \ge\bigl[u_0(x) \bigr]^k\exp \bigl\{ \bigl[ k(k-1) \ell_\sigma^2 - k \bigr]t \bigr\}. \end{equation} Because $u_0$ is not identically zero, it follows that \begin{equation} \underline{\gamma}_k(u) \ge\underline{\gamma}_k(v) \ge k(k-1) \ell_\sigma^2 - k. \end{equation} The preceding is $\ge(1-\varepsilon) k^2\ell_\sigma^2$ when $k\ge\varepsilon^{-1}+(\varepsilon\ell_\sigma^2)^{-1}$. This completes the proof of the theorem. \end{pf*} \section{A local approximation theorem} \label{sec:approx} In this section we develop a description of the local dynamics of the random field $t\mapsto u_t(\bullet)$ in the form of several approximation results. Our first approximation lemma is a standard sample-function continuity result; it states basically that outside a single null set, \begin{equation}\qquad \label{eq:local:modulus} u_{t+\tau}(x)= u_t(x) + O \bigl( \tau^{(1+o(1))/2} \bigr)\qquad \mbox{as $\tau\to0$, for all $t\ge0$ and $x\in \Z^d$}. \end{equation} The result is well known, but we need to be cautious with various constants that crop up in the proof. Therefore, we include the details to account for the dependencies of the implied constants. \begin{lemma}\label{lem:modulus} There exists a version of $u$ that is a.s. continuous in $t$ with critical H\"older exponent $\ge\nicefrac12$. In fact, for every $T\ge1$, $\varepsilon\in(0 ,1)$ and $k\ge2$, \begin{equation} \sup_{x\in\Z^d}\sup_I \E \biggl( \mathop{ \sup_{s,t\in I}}_{ s\neq t} \biggl[ \frac{|u_t(x) - u_s(x)|}{|t-s|^{(1-\varepsilon)/2}} \biggr]^k \biggr)<\infty, \end{equation} where ``$\sup_I$'' denotes the supremum over all closed subintervals $I$ of $[0 ,T]$ that have length $\le1$. \end{lemma} \begin{pf} Minkowski's inequality gives \begin{equation} \label{QQQ} \bigl[ \E \bigl( \bigl\llvert u_{t+\tau}(x) - u_t(x)\bigr\rrvert ^k \bigr) \bigr]^{1/k} \le|Q_1| + Q_2 + Q_3, \end{equation} where \begin{eqnarray} \label{eq:Q_1-Q_3} \nonumber Q_1 &:=& (\tilde{p}_{t+\tau}*u_0) (x) - (\tilde{p}_t*u_0) (x), \\ \qquad Q_2 &:=& \biggl[ \E \biggl(\biggl | \sum _{y\in\Z^d}\int_0^t \bigl[ p_{t+\tau-s}(y-x) \nonumber \\[-10pt] \\[-10pt] \nonumber &&\hspace*{58pt}{} - p_{t-s}(y-x) \bigr] \sigma \bigl(u_s(y)\bigr) \,\d B_s(y) \biggr|^k \biggr) \biggr]^{1/k} , \\ Q_3 &:=& \biggl[ \E \biggl( \biggl| \sum_{y\in\Z^d} \int_t^{t+\tau} p_{t+\tau-s}(y-x) \sigma \bigl(u_s(y)\bigr) \,\d B_s(y)\biggr |^k \biggr) \biggr]^{1/k}.\nonumber \end{eqnarray} We estimate each item in turn. Let $J_{t,t+\tau}$ denote the event that the random walk $X$ jumps some time during the time interval $(t ,t+\tau)$. Because \begin{eqnarray} \label{eq:p-p} \sum_{x\in\Z^d}\bigl\llvert p_{t+\tau}(x) - p_t(x) \bigr\rrvert &=&\sum _{x\in\Z^d}\bigl\llvert \E ( \1_{\{X_{t+\tau}=x\}} - \1_{\{ X_t=x\} }; J_{t,t+\tau} )\bigr\rrvert \nonumber \\[-8pt] \\[-8pt] \nonumber &\le& 2\P(J_{t,t+\tau}) = 2 \bigl(1-{\mathrm{e}}^{-\tau} \bigr) \le2\tau, \end{eqnarray} we obtain the following estimate for $|Q_1|$: \begin{equation} \label{eq:Q_1} |Q_1| \le2\|u_0\|_{\ell^\infty(\Z^d)}\tau. \end{equation} By BDG Lemma~\ref{lem:BDG}, \begin{eqnarray} \label{eq:Q2:prec} Q_2^2 &\le&4k\sum _{y\in\Z^d}\int_0^t \bigl[ p_{t+\tau-s}(y-x) - p_{t-s}(y-x) \bigr]^2 \bigl\{ \E \bigl( \bigl\llvert \sigma\bigl(u_s(y)\bigr)\bigr\rrvert ^k \bigr) \bigr\}^{2/k} \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&4k \int_0^t \mathscr{Q}(s)\sup _{y\in\Z^d} \bigl\{ \E \bigl( \bigl\llvert \sigma \bigl(u_s(y)\bigr)\bigr\rrvert ^k \bigr) \bigr \}^{2/k} \,\d s, \end{eqnarray} where \begin{equation} \mathscr{Q}(s) := \sum_{z\in\Z^d}\bigl|p_{t+\tau-s}(z)-p_{t-s}(z)\bigr|^2 \qquad (0<s<t).\vadjust{\goodbreak} \end{equation} Note that $\mathscr{Q}(s) \le [\sum_z | p_{t+\tau-s}(z)-p_{t-s} | ]^2 \le4\tau^2$ uniformly for $s\in(0,t)$ from~\eqref{eq:p-p}. This shows that \[ Q_2^2 \le16k\tau^2 \int _0^t\sup_{y\in\Z^d} \bigl\{ \E \bigl( \bigl\llvert \sigma\bigl(u_s(y)\bigr)\bigr\rrvert ^k \bigr) \bigr\}^{2/k} \,\d s. \] Because $|\sigma(z)|\le\lip|z|$ for all $z\in\R$, the already-proved bound \eqref{eq:Upper:LE} tells us that there exist constants $c,c_k\in(0 ,\infty)$ $[k\ge2]$ such that \begin{equation} \label{eq:LE:k} \sup_{y\in\Z^d}\E \bigl( \bigl\llvert \sigma \bigl(u_s(y)\bigr)\bigr\rrvert ^k \bigr) \le c_k^k {\mathrm{e}}^{ck^2s}\qquad \mbox{for all integers $k \ge2$ and $s\ge0$}. \end{equation} Therefore, \begin{equation} \label{eq:Q_2} Q_2^2 \le8c^{-1}c_k^2{\mathrm{e}}^{2ckt}\tau^2. \end{equation} Finally, we apply BDG Lemma~\ref{lem:BDG} to see that \begin{eqnarray} \label{eq:Q_3:prec} Q_3^2 &\le&4k\sum _{y\in\Z^d} \int_t^{t+\tau} \bigl[ p_{t+\tau-s}(y-x) \bigr]^2 \bigl\{ \E \bigl( \bigl\llvert \sigma \bigl(u_s(y)\bigr)\bigr\rrvert ^k \bigr) \bigr \}^{2/k} \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&4kc_k^2\sum_{y\in\Z^d} \int _t^{t+\tau} \bigl[ p_{t+\tau-s}(y-x) \bigr]^2 {\mathrm{e}}^{2cks} \,\d s, \end{eqnarray} owing to \eqref{eq:LE:k}. Because $\sum_{y\in\Z^d}[p_h(y-x)]^2\le1$ for all $h\ge0$, we find that \begin{equation} \label{eq:Q_3} Q_3^2 \le4c^{-1}c_k^2{\mathrm{e}}^{2ck(t+\tau)}\tau. \end{equation} We combine \eqref{eq:Q_1}, \eqref{eq:Q_2} and \eqref{eq:Q_3} and find that for all integers $k\ge2$, there exists a finite and positive constant $\tilde{a}:= \tilde{a}(T,k)$ such that for every $\tau\in(0 ,1)$, \begin{equation} \label{eq:mod:u} \sup_{x\in\Z^d} \sup_{t\in(0,T)} \E \bigl(\bigl\llvert u_{t+\tau}(x) - u_t(x)\bigr\rrvert ^k \bigr) \le \tilde{a} {\mathrm{e}}^{\tilde{a}T}\tau^{k/2}. \end{equation} The lemma follows from this bound, and an application of a quantitative form of the Kolmogorov continuity theorem \cite{RevuzYor}, Theorem~2.1, page 25. We omit the remaining details, as they are nowadays standard. \end{pf} Our next approximation result is the highlight of this section, and refines \eqref{eq:local:modulus} by inspecting more closely the main contribution to the $O(\tau^{(1+o(1))/2})$ error term in \eqref{eq:local:modulus}. In order to describe the next approximation result, we first define for every fixed $t\ge0$ an infinite-dimensional Brownian motion $B^{(t)}$ as follows: \begin{equation} B^{(t)}_\tau(x) := B_{\tau+t}(x)-B_t(x)\qquad \bigl(x\in\Z^d, \tau\ge0\bigr). \end{equation} If we continue to hold $t$ fixed, then it is easy to see that $\{B^{(t)}_\bullet(x)\}_{x\in\Z^d}$ is a collection of independent $d$-dimensional Brownian motions. Furthermore, the entire process $B^{(t)}$ is independent of the infinite-dimensional random variable $u_t(\bullet)$, since it is easy to see from the proof of the first part of Theorem~\ref{th:exist:unique} that $u_t$ is a measurable function of $\{B_s(y)\}_{s\in[0,t],y\in\Z^d}$, which is therefore independent of $B^{(t)}$ by the Markov property of $B$. Now for every fixed $t\ge0$ and $x\in\Z^d$, consider the solution $u^{(t)}_\bullet(x)$ to the following (autonomous/noninteracting) It\^o stochastic differential equation: \begin{equation} \label{eq:Feller \cases{\displaystyle\frac{\d u^{(t)}_\tau(x)}{\d\tau} = \frac{\d(\tilde{p}_\tau*u_t)(x)}{\d\tau} + \sigma \bigl( u^{(t)}_\tau(x) \bigr)\frac{\d B^{(t)}_\tau(x)}{\d\tau},\vspace *{2pt} \cr \mbox{subject to\qquad$u^{(t)}_0(x) = u_t(x)$}.} \end{equation} Note, once again, that $B^{(t)}$ is independent of $u_t(\bullet)$. Moreover, \begin{equation} \sup_{\tau>0}\E \bigl(\bigl\llvert (\tilde{p}_\tau*u_t ) (x)\bigr\rrvert ^2 \bigr) \le\sup_{y\in\Z^d}\E \bigl(\bigl|u_t(y)\bigr|^2 \bigr)<\infty, \end{equation} thanks to the already-proved bound \eqref{eq:Upper:LE} and the Cauchy--Schwarz inequality. Therefore, \eqref{eq:Feller} is a standard It\^o-type SDE and hence has a unique strong solution. \begin{theorem}[(The local-diffusion property)]\label{th:local:OU} For every $t\ge0$, the following holds a.s. for all $x\in\Z^d$: \begin{equation} \label{eq:local:OU} u_{t+\tau}(x) = u^{(t)}_\tau(x) + O \bigl( \tau^{(3/2)+o(1)} \bigr)\qquad \mbox{as $\tau\downarrow0$}. \end{equation} \end{theorem} The proof of Theorem~\ref{th:local:OU} hinges on three technical lemmas that we state next. \begin{lemma}\label{lem:local:OU} Choose and fix $t\ge0$, $\tau\in[0 ,1]$, and $x\in\Z^d$, and define \begin{eqnarray} \mathscr{A} &:=&\sum_{y\in\Z^d} \int _t^{t+\tau} p_{t+\tau-s}(y-x) \sigma \bigl(u_s(y)\bigr) \,\d B_s(y), \nonumber \\[-8pt] \\[-8pt] \nonumber \mathscr{B} &:=& \int_t^{t+\tau} \sigma \bigl(u_s(x)\bigr) \,\d B_s(x). \end{eqnarray} Then, for all real numbers $k\ge2$ there exist a finite constant $C_k>0$---depending on $k$ but not on $(t ,\tau,x)$---and a finite constant $C>0$---not depending on $(t ,\tau,x ,k)$---such that \begin{equation} \E \bigl( \llvert \mathscr{A}-\mathscr{B}\rrvert ^k \bigr) \le C_k{\mathrm{e}}^{Ck^2(t+1)}\tau^{3k/2}. \end{equation} \end{lemma} \begin{lemma}\label{lem:local:OU1} For every $k\ge2$ and $T\ge1$, there exists a finite constant $C(k ,T)$ such that for every $\tau\in(0 ,1]$, \begin{equation} \sup_{t\in[0,T]}\sup_{x\in\Z^d} \E \bigl( \bigl \llvert u_{t+\tau}(x) - u^{(t)}_\tau(x) \bigr\rrvert ^k \bigr) \le C(k ,T)\tau^{3k/2}. \end{equation} \end{lemma} \begin{lemma}\label{lem:modulus:OU} There exists a version of $u^{(\bullet)}$ that is a.s. continuous in $(t ,\tau)$. Moreover, for every $T\ge1$, $\varepsilon\in(0 ,1)$ and $k\ge2$, \begin{equation} \sup_{t\in[0,T]}\sup_{x\in\Z^d}\sup _I \E \biggl( \mathop{\sup_{\nu,\mu\in I}}_{ \nu\neq\mu} \biggl[ \frac{|u^{(t)}_\nu(x) - u^{(t)}_\mu(x)|}{ |\nu-\mu|^{(1-\varepsilon)/2}} \biggr]^k \biggr)<\infty, \end{equation} where ``$\sup_I$'' denotes the supremum over all closed subintervals $I$ of $[0 ,T]$ that have length $\le1$. \end{lemma} In order to maintain the flow of the discussion, we prove Theorem~\ref{th:local:OU} first. Then we conclude this section by establishing the three supporting lemmas mentioned above. \begin{pf*}{Proof of Theorem~\ref{th:local:OU}} Throughout the proof we choose and fix some $t\in[0 ,T]$ and $x\in\Z^d$. Our plan is to prove that for all $\delta\in(0 ,\nicefrac12)$, \begin{equation} u_{t+\tau}(x) - u^{(t)}_\tau(x) = O \bigl( \tau^{(3/2)-\delta} \bigr)\qquad \mbox{as $\tau \downarrow0$, a.s.} \end{equation} Henceforth, we choose and fix some $\delta\in(0 ,\nicefrac12)$, and denote by $A_k,A_k',A_k''$, etc. finite constants that depend only on a parameter $k\ge2$ that will be selected later, during the course of the proof. Thanks to Lemma~\ref{lem:local:OU1}, for all $k\ge2$ and $\tau\in[0 ,1]$, \begin{equation} \P \bigl\{\bigl\llvert u_{t+\tau}(x) - u^{(t)}_\tau(x) \bigr\rrvert \ge\tfrac{1}3\tau^{(3/2)-\delta} \bigr\} \le C(k ,T)\tau ^{\delta k}. \end{equation} We can choose $k$ large enough and then apply the Borel--Cantelli lemma in order to deduce that with probability one, \begin{equation}\qquad \bigl\llvert u_{t+\tau_n}(x) - u^{(t)}_{\tau_n}(x) \bigr \rrvert < \tau_n^{(3/2)-\delta}\qquad \mbox{for all but a finite number of $n$'s,} \end{equation} where $\tau_n := n^{\delta-(1/2)}$. Because $\tau_n-\tau_{n+1} \sim\operatorname{const} \times n^{-1}\tau_n$ as $n\to\infty$, H\"older continuity ensures the following (Lemmas \ref{lem:modulus} and \ref{lem:modulus:OU}): Uniformly for all $\tau\in[\tau_{n+1},\tau_n]$, \begin{eqnarray} \bigl\llvert u_{t+\tau}(x) - u_{t+\tau_n}(x)\bigr\rrvert + \bigl\llvert u^{(t)}_{\tau_n}(x) - u^{(t)}_\tau(x) \bigr\rrvert &=& O \bigl( [\tau_n/n ]^{(1/2)-\delta} \bigr)\qquad \mbox{a.s.} \nonumber \\ &=&O \bigl(\tau_n^{(3/2)-\delta} \bigr), \end{eqnarray} by the particular choice of the sequence $\{\tau_n\}_{n=1}^\infty$. The preceding two displays can now be combined to imply \eqref{eq:local:OU}. \end{pf*} \begin{pf*}{Proof of Lemma~\ref{lem:local:OU}} We may rewrite $\mathscr{B}$ as follows: \begin{equation} \mathscr{B} =\sum_{y\in\Z^d} \int_t^{t+\tau} \1_{\{0\}}(y-x) \sigma\bigl(u_s(y)\bigr) \,\d B_s(y). \end{equation} Therefore, BDG Lemma~\ref{lem:BDG} can be used to show that \begin{eqnarray} \nonumber && \bigl\{ \E \bigl(\llvert \mathscr{A}-\mathscr{B}\rrvert ^k \bigr) \bigr\}^{2/k} \\ &&\qquad\le4k\sum_{y\in\Z^d}\int_t^{t+\tau} \bigl[ p_{t+\tau-s}(y-x) - \1_{\{0\}}(y-x) \bigr]^2 \bigl\{ \E \bigl( \bigl\llvert \sigma\bigl(u_s(y)\bigr)\bigr\rrvert ^k \bigr) \bigr\}^{2/k} \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le4kc_k^2{\mathrm{e}}^{2ck(t+1)} \biggl(\sum _{y\in\Z^d\setminus\{0\} }\int_0^\tau \bigl[ p_{s}(y) \bigr]^2 \,\d s+\int_{0}^{\tau} \bigl[ 1- p_s(0) \bigr]^2\,\mathrm{d}s \biggr) \\ \nonumber &&\qquad\le4kc_k^2{\mathrm{e}}^{2ck(t+1)} 2\int _{0}^{\tau} \bigl[ 1-p_s(0) \bigr]^2\,\mathrm{d}s, \end{eqnarray} where $c,c_k$ appear in \eqref{eq:LE:k}. Observe that $p_s(0)= \P\{X_s=0\}\geq\P\{N_s=0\}={\mathrm{e}}^{-s}$, where $\{N_s\} _{s\ge0}$ denotes the underlying Poisson clock. Therefore, we obtain $\int_0^{\tau} [ 1-p_s(0) ]^2 \,\d s \le(1/3)\tau^3$, and hence \begin{equation} \E \bigl(\llvert \mathscr{A}-\mathscr{B}\rrvert ^k \bigr) \le(8/3)^{k/2} k^{k/2}c_k^k{\mathrm{e}}^{ck^2(t+1)}\tau^{3k/2}. \end{equation} This implies the lemma. \end{pf*} \begin{pf*}{Proof of Lemma~\ref{lem:local:OU1}} In accord with \eqref{mild}, we may write $u_{t+\tau}(x)$ as \begin{equation} ( \tilde{p}_{t+\tau}*u_0 ) (x) + \sum _{y\in\Z^d} \int_0^t p_{t+\tau-s}(y-x)\sigma\bigl(u_s(y)\bigr) \,\d B_s(y) + \mathscr{A}, \label{eq:post:t:u} \end{equation} where $\mathscr{A}$ was defined in Lemma~\ref{lem:local:OU}. By the Chapman--Kolmogorov property of the transition functions $\{p_t\}_{t\ge0}$, \begin{eqnarray} &&(\tilde{p}_\tau*u_t) (x) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad= (\tilde{p}_{t+\tau}*u_0) (x) + \sum_{y\in\Z^d}\int_0^t p_{t+\tau-s}(y-x)\sigma \bigl(u_s(y) \bigr) \,\d B_s(y). \end{eqnarray} The exchange of summation with stochastic integration can be justified, using the already-proved moment bound \eqref{eq:Upper:LE} of Theorem~\ref{th:exist:unique}; we omit the details. Instead, let us apply this in \eqref{eq:post:t:u} to see that \begin{eqnarray} u_{t+\tau}(x) &=& (\tilde{p}_\tau*u_t) (x) + \int_t^{t+\tau} \sigma\bigl(u_s(x) \bigr) \,\d B_s(x) + ( \mathscr{A}-\mathscr{B} ) \nonumber \\[-8pt] \\[-8pt] \nonumber &=& (\tilde{p}_\tau*u_t) (x) + \int_0^\tau \sigma\bigl(u_{t+s}(x)\bigr) \,\d_s B^{(t)}_s(x) + ( \mathscr{A}-\mathscr{B} ). \end{eqnarray} Lemma~\ref{lem:local:OU} implies that for all $k\ge2$, $t,\tau\ge0$ and $x\in\Z^d$, \begin{eqnarray} \label{eq:local:approx1} &&\E \biggl( \biggl| u_{t+\tau}(x) - ( \tilde{p}_\tau*u_t) (x) -\int_0^\tau \sigma\bigl(u_{t+s}(x)\bigr) \,\d_s B^{(t)}_s(x) \biggr|^k \biggr) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad \le a_k {\mathrm{e}}^{ak^2(t+1)}\tau^{3k/2}, \end{eqnarray} where $a\in(0 ,\infty)$ is universal and $a_k\in(0 ,\infty)$ depends only on $k$. On the other hand, \begin{equation} u^{(t)}_\tau(x) - (\tilde{p}_\tau*u_t ) (x) - \int_0^\tau\sigma \bigl(u^{(t)}_s(x) \bigr) \,\d B^{(t)}_s(x)=0\qquad \mbox{a.s.,} \end{equation} by the very definition of $u^{(t)}$, and thanks to the fact that $u^{(t)}_0(y)=u_t(y)$. The preceding two displays and Minkowski's inequality that \begin{equation} \psi(\tau) := \bigl\{\E \bigl( \bigl\llvert u_{t+\tau}(x) - u^{(t)}_\tau(x) \bigr\rrvert ^k \bigr) \bigr \}^{1/k} \le a_k^{1/k}{\mathrm{e}}^{ak(t+1)} \tau^{3/2} + Q, \end{equation} where \begin{equation} Q := \biggl\{\E \biggl(\biggl\llvert \int_0^\tau \bigl[\sigma\bigl(u_{t+s}(x)\bigr) - \sigma \bigl( u^{(t)}_s(x) \bigr) \bigr] \,\d_s B^{(t)}_s(x)\biggr\rrvert ^k \biggr) \biggr\}^{1/k}. \end{equation} According to BDG Lemma~\ref{lem:BDG} (actually we need a one-dimensional version of that lemma only), and since $|\sigma(r)-\sigma(\rho)| \le\lip|r-\rho|$, \begin{eqnarray} Q^2 &\le&4k\lip^2\int_0^\tau \bigl\{ \E \bigl(\bigl\llvert u_{t+s}(x)-u^{(t)}_s(x) \bigr\rrvert ^k \bigr) \bigr\}^{2/k} \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &=&4k\lip^2 \int_0^\tau\bigl[ \psi(s)\bigr]^2 \,\d s. \end{eqnarray} Thus we find that \begin{eqnarray} \bigl[\psi(\tau)\bigr]^2 \le2a_k^{2/k}{\mathrm{e}}^{2ak(t+1)}\tau^{3} + 8k\lip^2\int _0^\tau\bigl[\psi(s)\bigr]^2 \,\d s \nonumber \\[-8pt] \\[-8pt] \eqntext{\mbox{for all $0\le\tau\le1$}.} \end{eqnarray} The lemma follows from this and an application of Gronwall's lemma. \end{pf*} \begin{pf*}{Proof of Lemma~\ref{lem:modulus:OU}} One can model closely a proof after that of Lemma~\ref{lem:modulus}. However, we omit the details, since this is a result about finite-dimensional diffusions and as such simpler than Lemma~\ref{lem:modulus}. \end{pf*} We conclude this section with a final approximation lemma. The next assertion shows that the solution to \eqref{SHE} depends continuously on its initial function (in a suitable topology). \begin{lemma}\label{lem:flow} Let $u$ and $v$ denote the unique solutions to \mbox{\eqref{SHE}}, corresponding, respectively, to initial functions $u_0$ and $v_0$. Then \begin{equation}\qquad \sup_{x\in\Z^d}\E \bigl(\bigl\llvert u_t(x)-v_t(x) \bigr\rrvert ^2 \bigr) \le \|u_0-v_0 \|_{\ell^\infty(\Z^d)}^2 {\mathrm{e}}^{\lip^2 t} \qquad\mbox{for all $t\ge0$}. \end{equation} \end{lemma} \begin{pf} Choose and fix $t\ge0$. The fact that $\sum_{y\in\Z^d} p_t(y)=1$ alone ensures that \begin{equation} \sup_{x\in\Z^d}\bigl\llvert ( \tilde{p}_t*u_0 ) (x) - ( \tilde{p}_t*v_0 ) (x) \bigr\rrvert \le \|u_0-v_0\|_{\ell^\infty(\Z^d)}. \end{equation} Therefore, \eqref{mild} and It\^o's isometry together imply that \begin{eqnarray} &&\E \bigl(\bigl\llvert u_t(x) - v_t(x)\bigr\rrvert ^2 \bigr) \nonumber \\ &&\qquad\le\|u_0-v_0\|_{\ell^\infty(\Z^d)}^2 \\ &&\qquad\quad{}+ \lip^2 \int_0^t \llVert p_s \rrVert _{\ell^2(\Z^d)}^2 \,\cdot\,\sup _{y\in\Z^d}\E \bigl( \bigl\llvert u_s(y) - v_s(y)\bigr\rrvert ^2 \bigr) \,\d s.\nonumber \end{eqnarray} Since $\|p_s\|_{\ell^2(\Z^d)}^2 = \P\{X_s=X_s'\}\le1$, where $X'$ is an independent copy of $X$, we may conclude that $f(t):=\sup_{x\in\Z^d}\E (|u_t(x)-v_t(x)|^2)$ satisfies \begin{equation} f(t) \le\|u_0-v_0\|_{\ell^\infty(\Z^d)}^2 + \lip^2\int_0^tf(s) \,\d s. \end{equation} Therefore, the lemma follows from Gronwall's inequality. \end{pf} \section{Proof of Theorem \texorpdfstring{\protect\ref{th:local:RadonNikodym}}{2.2}} \label{sec:thm1.2} Theorem~\ref{th:local:RadonNikodym} is a consequence of the following result. \begin{proposition}\label{pr:local} For every $t\ge0$, the following holds a.s. for all $x\in\Z^d$: \begin{eqnarray} \label{eq:local:OU1} u_{t+\tau}(x) - u_t(x) = \sigma \bigl( u_t(x) \bigr) \bigl\{ B_{t+\tau}(x)-B_t(x) \bigr \} + o \bigl( \tau^{1+o(1)} \bigr) \nonumber \\[-8pt] \\[-8pt] \eqntext{\mbox{as $\tau\downarrow0$}.} \end{eqnarray} \end{proposition} Indeed,\vspace*{1pt} we obtain \eqref{eq:local:RN} from this proposition, simply because well-known properties of Brownian motion imply that for all $\varepsilon\in(0 ,\nicefrac12)$ and $t\ge0$, \begin{equation} \lim_{\tau\downarrow0}\frac{\tau^{1-\varepsilon}}{ B_{t+\tau}(x)-B_t(x)}=0 \qquad\mbox{in probability}. \end{equation} Moreover, \eqref{eq:local:LIL} follows from the local law of the iterated logarithm for Brownian motion. It remains to prove Proposition~\ref{pr:local}. \begin{pf} According to \eqref{eq:local:approx1}, for every integer $k\ge2$, and all $t,\tau\ge0$ and $x\in\Z^d$, \begin{eqnarray} &&\E \biggl(\biggl\llvert u_{t+\tau}(x) - u_t(x) - \int_0^\tau\sigma\bigl(u_{t+s}(x) \bigr) \,\d_s B^{(t)}_s(x)\biggr\rrvert ^k \biggr) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le2^{k-1} \bigl[ a_k{\mathrm{e}}^{ak^2(t+1)} \tau^{3k/2} + \E \bigl(\bigl\llvert u_t(x) - ( \tilde{p}_\tau*u_t ) (x) \bigr\rrvert ^k \bigr) \bigr]. \end{eqnarray} We may write \begin{eqnarray} &&\E \bigl(\bigl\llvert u_t(x) - (\tilde{p}_\tau*u_t ) (x) \bigr\rrvert ^k \bigr)\nonumber \\ &&\qquad=\E \biggl( \biggl| u_t(x) - \sum_{y\in\Z^d}p_\tau(y-x)u_t(y) \biggr|^k \biggr) \\ &&\qquad=\E \biggl( \biggl| u_t(x)\P\{X_\tau\neq0\} - \sum _{y\in\Z^d\setminus\{x\}} p_\tau(y-x) u_t(y) \biggr|^k \biggr).\nonumber \end{eqnarray} Because $\P\{X_\tau\neq0\}=1-\exp(-\tau)\le\tau$, Minkowski's inequality shows that \begin{eqnarray} && \bigl\{ \E \bigl(\bigl\llvert u_t(x) - ( \tilde{p}_\tau*u_t ) (x) \bigr\rrvert ^k \bigr) \bigr\}^{1/k} \nonumber \\ &&\qquad\le\tau \bigl\{ \E \bigl(\bigl|u_t(x)\bigr|^k \bigr) \bigr\} ^{1/k} + \sum_{y\in\Z^d\setminus\{x\}} p_\tau(y-x) \bigl\{ \E \bigl(\bigl|u_t(y)\bigr|^k \bigr) \bigr\}^{1/k} \\ &&\qquad\le2\tau\sup_{y\in\Z^d} \bigl\{ \E \bigl(\bigl|u_t(y)\bigr|^k \bigr) \bigr\}^{1/k}. \nonumber \end{eqnarray} We can conclude from this development and from Theorem~\ref{th:exist:unique} that there exists $A_k<\infty $, depending only on $k$, and a universal $A<\infty$ such that \begin{eqnarray} \label{eq:uuu} &&\E \biggl(\biggl\llvert u_{t+\tau}(x) - u_t(x) -\int_0^\tau\sigma \bigl(u_{t+s}(x)\bigr) \,\d_s B^{(t)}_s(x) \biggr\rrvert ^k \biggr) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad \le A_k {\mathrm{e}}^{Ak^2(t+1)} \bigl[ \tau^{3k/2} + \tau^k \bigr] \le A_k {\mathrm{e}}^{Ak^2(t+1)} \tau^k, \end{eqnarray} for all $\tau\in[0 ,1]$. Now, we may apply BDG Lemma~\ref{lem:BDG} in order to see that \begin{eqnarray} &&\biggl[\E \biggl(\biggl |\int_0^\tau\sigma \bigl(u_{t+s}(x)\bigr) \,\d_s B^{(t)}_s(x) -\sigma\bigl(u_t(x)\bigr) \bigl\{ B_{t+\tau}(x)-B_t(x) \bigr\}\biggr |^k \biggr) \biggr]^{2/k} \nonumber \\ &&\qquad= \biggl[\E \biggl(\biggl\llvert \int_0^\tau \bigl[ \sigma\bigl(u_{t+s}(x)\bigr) -\sigma\bigl(u_t(x) \bigr) \bigr] \,\d_s B^{(t)}_s(x) \biggr\rrvert ^k \biggr) \biggr]^{2/k} \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le4k\lip^2\int_0^\tau \bigl[ \E \bigl(\bigl\llvert u_{t+s}(x)- u_t(x)\bigr\rrvert ^k \bigr) \bigr]^{2/k} \,\d s \\ &&\qquad\le\tilde{a}_k{\mathrm{e}}^{\tilde{a}t}\int_0^\tau s \,\d s \le\operatorname{const}\,\cdot\,\tau^2,\nonumber \end{eqnarray} using \eqref{eq:mod:u}. Therefore, we can deduce from \eqref{eq:uuu} that \begin{equation} \E \bigl(\bigl\llvert D(\tau, x) \bigr\rrvert ^k \bigr) \le c_{k,t} \tau^k\qquad (0\le\tau\le1), \end{equation} where we defined \begin{equation} D(\tau, x):= u_{t+\tau}(x) - u_t(x) -\sigma \bigl(u_t(x)\bigr) \bigl\{ B_{t+\tau }(x)-B_t(x) \bigr\}, \end{equation} and $c_{k,t}$ is a finite constant that depends only on $k$ and $t$; in particular, $c_{k,t}$ does not depend on $\tau$. Choose and fix some $\eta>\xi>0$ such that $\eta+\xi<\nicefrac12$, and then apply the Chebyshev inequality, and the preceding with any choice of integer $k>\xi^{-1}$, in order to see that $\sum_{n=1}^\infty\P\{ | D(n^{-\eta}, x) | > n^{-(\eta-\xi)} \} \le c_{k,t}\sum_{n=1}^\infty n^{-\xi k}<\infty$. Thus \begin{equation} \label{eq:D} D \bigl( n^{-\eta}, x \bigr)= O \bigl( n^{-(\eta-\xi)} \bigr) \qquad\mbox{a.s.} \end{equation} by the Borel--Cantelli lemma. Because $n^{-\eta} - (n+1)^{-\eta} = O ( n^{-1-\eta} )$, the modulus of continuity of Brownian motion, together with Lemma~\ref{lem:modulus}, imply that \begin{equation}\qquad \sup_{(n+1)^{-\eta}\le\tau\le n^{-\eta}} \bigl\llvert D \bigl( n^{-\eta}, x \bigr)- D(\tau, x) \bigr\rrvert = O \bigl( n^{-1/2} \bigr)=o \bigl( n^{-(\eta+\xi)} \bigr) \qquad\mbox{a.s.} \end{equation} Therefore a standard monotonicity argument and \eqref{eq:D} together reveal that $D(t, x) = O ( t^{(\eta-\xi)/\eta} )$ as $t\downarrow0$, a.s. Since $\eta>\xi$ are arbitrary positive numbers, it follows that $\limsup_{t\downarrow0}(\log D(t, x)/\log t )\le1$ a.s. This is another way to state the result. \end{pf} \section{Proof of Theorem \texorpdfstring{\protect\ref{th:CLT}}{2.5}} \label{sec:thm1.4} First we prove a preliminary lemma that guarantees strict positivity of the solution to the \eqref{SHE}. We follow the method described in Conus, Joseph and Khoshnevisan \cite{CJK11}, Theorem~5.1, which in turn borrowed heavily from ideas of Mueller \cite{Mueller} and Mueller and Nualart \cite{MN}. \begin{lemma}\label{lem:strict:pos} $\inf_{0\le t \le T} u_t(x) > 0$ a.s. for every $T\in(0 ,\infty)$ and all $x\in\Z^d$ that satisfy $u_0(x) > 0$. \end{lemma} \begin{pf} We are going to prove that if $u_0(x_0)>0$ for a fixed $x_0\in\Z^d$, then there exist finite and positive constants $A$ and $C$ such that \begin{equation} \P \Bigl\{\inf_{0<s<t}u_s(x_0) \le \varepsilon \Bigr\} \le A\varepsilon ^{C\log\vert \log\varepsilon\vert}, \end{equation} for that same point $x_0$, uniformly for all $\varepsilon\in(0 ,1)$. It turns out to be convenient to prove the following equivalent formulation of the preceding: \begin{equation} \label{eq:goal:pos} \P \Bigl\{\inf_{0<s<t}u_s(x_0) \le{\mathrm{e}}^{-n} \Bigr\} \le An^{-Cn}, \end{equation} simultaneously for all $n\ge1$, after a possible relabeling of the constants $A,C\in(0 ,\infty)$. If so, then we can simply let $n\to\infty$ and deduce the lemma. Without loss of generality we assume that $u_0(0)>0$, and we aim to prove \eqref{eq:goal:pos} with $x_0=0$. In fact, we will simplify the exposition further and establish \eqref {eq:goal:pos} when $u_0(0)=1$; the general case follows from this one and scaling. Finally, we appeal to a comparison principle (Theorem~\ref{th:comparison}) in order to reduce our problem further to the following special case: \begin{equation} u_0(x) = \delta_0(x) \qquad\mbox{for all $x\in \Z^d$.} \end{equation} Thus we consider this case only from now on. Let $\mathcal{F}_t:=\sigma\{B_s(x)\dvtx x \in\Z^d, 0<s \le t\}$ describe the filtration generated by time $t$ by all the Brownian motions, enlarged so that $t\mapsto u_t$ is a $C(\R)$-valued (strong) Markov process. Set $T_0 := 0$, and define iteratively for $k\ge0$ the sequence of $\{ \mathcal{F}_t \}_{t>0}$-stopping times \begin{equation} T_{k+1}:= \inf \bigl\{s>T_k\dvtx u_s(0) \le{\mathrm{e}}^{-k-1} \bigr\}, \end{equation} using the usual convention that $\inf\varnothing:= \infty$. We may observe that the preceding definitions imply that, almost surely on $\{T_k<\infty\}$, \begin{equation} u_{T_k}(x) \ge{\mathrm{e}}^{-k}\delta_0(x)\qquad \mbox{for all } x\in\Z^d. \label{eq:restart:sol} \end{equation} We plan to apply the strong Markov property. In order to do that, we first define $u^{(k+1)}$ to be the unique continuous solution to the \eqref{SHE} (for same Brownian motions, pathwise), with initial data $u^{(k+1)}_0(x):={\mathrm{e}}^{-k}\delta_0(x)$. Next we note that, for every $k\ge0$, the random field \begin{equation} w^{(k+1)}_t(x) := {\mathrm{e}}^ku^{(k+1)}_t(x) \end{equation} solves the system \begin{equation} \cases{ \displaystyle\frac{\d w^{(k+1)}_t(x)}{\d t}=\bigl(\mathscr{L}w^{(k+1)}_t \bigr) (x) + \sigma_k \bigl( w^{(k+1)}_t(x) \bigr)\frac{\d B_t(x)}{\d t}, \vspace*{2pt} \cr w_0^{(k+1)}(x)=\delta_0(x),} \label{eq:auxil} \end{equation} where $\sigma_k(y):= {\mathrm{e}}^k\sigma({\mathrm{e}}^{-k}y)$. Because $\sigma(0)=0$, we have $\operatorname{Lip}_{\sigma_k} = \lip$, uniformly for all $k\ge1$. Thus we can keep track of the constants in the proof of Lemma~\ref{lem:modulus}, in order to deduce the existence of a finite constant $K:=K(\varepsilon)$ so that for all $t,s$ with $|t-s|<1$, \begin{equation} \E \biggl(\sup_{0<|t-s|<1}\frac{\llvert w^{(k+1)}_{t}(0) -w^{(k+1)}_s(0)\rrvert ^m}{|t-s|^{m(1-\varepsilon)/2}} \biggr) \le Km^2{\mathrm{e}}^{Km^2}, \label{eq:sup:const} \end{equation} for all real numbers $m\ge2$. For each $k\ge0$ let us define \begin{equation} T^{(k+1)}_1 = \inf \bigl\{ t>0\dvtx w_t^{(k+1)}(0) \le{\mathrm{e}}^{-1} \bigr\}. \end{equation} Equation \eqref{eq:restart:sol}, the strong Markov property and the comparison principle (Theorem~\ref{th:comparison}) together imply that outside of a null set, the solution to the revised SPDE \eqref{eq:auxil} satisfies \begin{equation} {\mathrm{e}}^{-k}w^{(k+1)}_t(x) \le u_{T_k+t}(x). \end{equation} Therefore, in particular, \begin{equation} T^{(k+1)}_1 \le T_{k+1} - T_k, \label{eq:time:order} \end{equation} and the stopping times $T^{(k+1)}_1$ and $T^{(\ell+1)}_1 $ are independent if $k \neq\ell$. For all real numbers $t\in(0 ,1)$ and $m\ge2$, \begin{eqnarray} \label{eq:time:bound} \P \bigl\{ T^{(k+1)}_1 \leq t \bigr\} &\le&\P \Bigl\{ \sup_{0<s< t}\bigl\llvert w_t^{(k+1)}(0)-w_s^{(k+1)}(0) \bigr\rrvert \ge1- {\mathrm{e}}^{-1} \Bigr\} \nonumber \\[-8pt] \\[-8pt] \nonumber &\le& K m^2{\mathrm{e}}^{Km^2} \bigl(1-{\mathrm{e}}^{-1} \bigr)^{-m} t^{(1-\varepsilon)m/2} , \end{eqnarray} where the last inequality follows by Chebyshev's inequality and \eqref{eq:sup:const} and is valid for all $0<\varepsilon<1$. Let us emphasize that the constant of the bound in \eqref{eq:time:bound} does not depend on the parameter $k$ which appears in the superscript of the random variable $T_1^{(k+1)}$. Now we compute \begin{eqnarray} \label{eq:iid:sum} \P \Bigl\{\inf_{0< s \le t} u_s(0)\le{{\mathrm{e}} }^{-n} \Bigr\} &=& \P\{ T_n \le t\} \nonumber\\ &=&\P\bigl\{(T_n -T_{n-1})+\cdots+(T_1 - T_0) \le t\bigr\} \\ \nonumber &\le&\P \bigl\{T^{(n)}_1+T^{(n-1)}_1+ \cdots+T^{(1)}_1 \le t \bigr\}, \end{eqnarray} owing to \eqref{eq:time:order}. The terms $T^{(n)}_1,\ldots,T^{(1)}_1$ that appear in the ultimate line of \eqref{eq:iid:sum} are independent nonnegative random variables. By the triangle inequality, if the sum of those terms is at most $t$, then certainly it must be that at least $n/2$ of those terms are at most $2t/n$. (This application of the triangle inequality is also known as the pigeon-hole principle.) If $n$ is an even integer, larger than $t>2$, then a simple union bound on \eqref{eq:iid:sum} and \eqref{eq:time:bound} yields \begin{eqnarray} \nonumber &&\P \Bigl\{\inf_{0< s \le t}u_s(0)\le{\mathrm{e}}^{-n} \Bigr\} \label{eq:n-m:rel} \\ &&\qquad\le\pmatrix{n \cr n/2} K^{n/2} m^n{\mathrm{e}}^{Km^2n/2} \bigl(1-{\mathrm{e}}^{-1} \bigr)^{-mn/2} (2t/n)^{(1-\varepsilon)mn/4} \\ \nonumber &&\qquad\le\widetilde{K}^n m^n{\mathrm{e}}^{Km^2n/2} \bigl(1-{\mathrm{e}}^{-1} \bigr)^{-mn/2} t^{(1-\varepsilon)mn/4} n^{-(1-\varepsilon)mn/4}2^{n(1+m(1-\varepsilon)/4)}, \end{eqnarray} uniformly for all real numbers $m\ge2$. Now we set $m := \log n / \log\log n$ in \eqref{eq:n-m:rel} in order to deduce \eqref{eq:goal:pos} for $x_0=0$ and every $n\ge1$ sufficiently large. This readily yields \eqref{eq:goal:pos}. \end{pf} Next we show that if we start with an initial profile $u_0$ such that $u_0(x)>0$ \emph{for at least one point} $x\in\Z^d$, then $u_t(z)>0$ \emph{for all} $z \in\Z^d$ and $t>0$ a.s. Because we are interested in establishing a lower bound, we may apply scaling and a comparison theorem (Theorem~\ref{th:comparison}) in order to reduce our problem to the following special case: \begin{equation} u_0=\delta_0. \end{equation} In this way, we are led to the following representation of the solution: \begin{equation} \label{eq:udelta} u_t(x)=p_t(x)+\int_0^t \sum_{y\in\Z^d} p_{t-s}(y-x) \sigma \bigl(u_s(y) \bigr) \,\d B_s(y). \end{equation} \begin{proposition}\label{prop:pos} If $u_0=\delta_0$, then $u_t(x)>0$ for all $x \in\Z^d$ and $t>0$ a.s. \end{proposition} Proposition~\ref{prop:pos} follows from a few preparatory lemmas. \begin{lemma}\label{lem:pos1} If $u_0=\delta_0$, then \begin{equation}\quad \E \bigl(\bigl|u_t(x)\bigr|^2 \bigr) \le\exp \bigl( \lip^2 t \bigr)\,\cdot\, \bigl[p_t(x)\bigr]^2\qquad \mbox{for all $t>0$ and $x\in\Z^d$}. \end{equation} \end{lemma} \begin{pf} We begin with representation \eqref{eq:udelta} of the solution $u$, in integral form, and appeal to Picard's iteration in order to prove the lemma. Let $u^{(0)}_t(x) := 1$ for all $t\ge0, x\in\Z^d$, and then let $\{u^{(n+1)}\}_{n\ge0}$ be defined iteratively by \begin{equation} u_t^{(n+1)}(x) := p_t(x)+ \int _0^t\sum_{y\in\Z^d} p_{t-s}(y-x) \sigma \bigl(u_s^{(n)}(y) \bigr) \,\d B_s(y). \end{equation} Let us define \begin{equation} M^{(k)}_t := \sup_{x\in\Z^d} \E \biggl( \biggl\llvert \frac{u^{(n+1)}_t(x)}{p_t(x)}\biggr\rrvert ^2 \biggr), \end{equation} and apply It\^o's isometry in order to deduce the recursive inequality for the $M^{(k)}$'s, \begin{equation} \label{eq:recur} M^{(n+1)}_t\le1+ \operatorname{Lip}_{\sigma}^2 \,\cdot\, \sup_{x\in\Z^d}\int_0^t \sum_y \biggl[\frac{p_{t-s}(y-x)p_s(y)}{p_t(x)} \biggr]^2 M^{(n)}_s \,\d s. \end{equation} Because $\sum_{y\in\Z^d}[f(y)]^2\le[\sum_{y\in\Z^d}f(y)]^2$ for all $f\dvtx\Z^d\to\R_+$, the semigroup property of $\{p_t\}_{t>0}$ yields the bound \begin{equation} \label{eq:3p} \sum_{y\in\Z^d} \bigl[p_{t-s}(y-x)p_s(y) \bigr]^2 \le\bigl[p_t(x)\bigr]^2, \end{equation} whence $M^{(n+1)}_t \le1 + \lip^2\,\cdot\,\int_0^t M^{(n)}_s \,\d s$ for all $t>0$ and $n\ge0$. It follows readily from this that $M^{(n)}_t \le\exp(\lip^2 t)$, uniformly for all $n\ge0$ and $t>0$; equivalently, \begin{equation} \E \bigl(\bigl|u^{(n)}_t(x)\bigr|^2 \bigr)\le{\mathrm{e}}^{\lip^2 t}\bigl[p_t(x)\bigr]^2, \end{equation} uniformly for all $n\ge0$, $x\in\Z^d$ and $t>0$. The lemma follows from this and Fatou's lemma, since $u^{(n)}_t(x)\to u_t(x)$ in $L^2(\P)$ as $n\to\infty$. \end{pf} Our next lemma shows that the random term on the right-hand side of \eqref{eq:udelta} is small, for small time, as compared with the nonrandom term in \eqref{eq:udelta}. \begin{lemma}\label{lem:pos2} Assume the conditions of Proposition~\ref{prop:pos}. Then there exists a finite constant $C>0$ such that for all $t\in(0 ,1)$, \begin{equation} \label{eq:randombd} \sup_{x\in\Z^d}\P \biggl\{\biggl\llvert \int _0^t \sum_{y\in\Z^d} p_{t-s}(y-x) \sigma \bigl(u_s(y) \bigr)\,\d B_s(y) \biggr\rrvert >\frac{p_t(x)}{2} \biggr\} \le Ct. \end{equation} \end{lemma} \begin{pf} By Lemma~\ref{lem:pos1} and It\^o's isometry, \begin{eqnarray} &&\E \biggl(\biggl\llvert \int_0^t \sum _{y\in\Z^d} p_{t-s}(y-x) \sigma \bigl(u_s(y) \bigr)\,\d B_s(y)\biggr\rrvert ^2 \biggr) \nonumber \\ &&\qquad\le\lip^2\,\cdot\,\int_0^t \sum_{y\in\Z^d} \bigl[p_{t-s}(y-x) p_s(y) \bigr]^2{\mathrm{e}}^{\lip^2 s} \,\d s \\ &&\qquad\le \lip^2 \bigl[p_t(x)\bigr]^2\,\cdot\,\int _0^t{\mathrm{e}}^{\lip^2 s} \,\d s,\nonumber \end{eqnarray} where we have used \eqref{eq:3p} in the last inequality. Because $\int_0^t \exp(\lip^2 s) \,\d s\le ct $ for all $t\in(0 ,1)$ with $c:=\exp(\lip^2)$, the lemma follows from Chebyshev's inequality. \end{pf} Now we can establish Proposition~\ref{prop:pos}. \begin{pf*}{Proof of Proposition~\ref{prop:pos}} Let us choose and fix an arbitrary $x\in\Z^d$. By the strong Markov property of the solution, and thanks to Lemma~\ref{lem:strict:pos}, we know that once the solution becomes positive at a point, it remains positive at that point at all future times, almost surely. Thus it suffices to show that $u_t(x)>0$ for all times of the form $t=2^{-k}$, when $k$ is a large enough integer. This is immediate from \eqref{eq:udelta} and \eqref{eq:randombd}, thanks to the Borel--Cantelli lemma. \end{pf*} The preceding lemmas lay the groundwork for the proof of Theorem~\ref{th:CLT}. We now proceed with the main proof. \begin{pf*}{Proof of Theorem~\ref{th:CLT}} Let us first consider the case $m=1$ and, without loss of generality, $x_1 = 0$. In this case, we may write \begin{eqnarray} &&\lim_{\tau\downarrow0} \P \bigl\{ S\bigl(u_{t+\tau}(0) \bigr) - S\bigl(u_t(0)\bigr) \le q\sqrt{\tau} \bigr\} \nonumber\\ &&\qquad=\lim_{\tau\downarrow0} \P \biggl\{ \int_{u_t(0)}^{u_{t+\tau}(0)} \frac{\d y}{\sigma(y)}\le q\sqrt{\tau} \biggr\} \\ \nonumber &&\qquad= \lim_{\tau\downarrow0} \P \biggl\{ \int_{u_t(0)}^{u_{t+\tau}(0)} \biggl(\frac{1}{\sigma(y)}-\frac{1}{\sigma(u_t(0))} \biggr)\,\d y +\frac{u_{t+\tau}(0)-u_t(0)}{\sigma(u_t(0))} \le q\sqrt\tau \biggr\}. \end{eqnarray} Lemma~\ref{lem:strict:pos} and the positivity condition on $\sigma$ ensure that $\sigma(u_t(0))>0$ a.s. Therefore, the theorem follows from Theorem~\ref{th:local:RadonNikodym} if we were to show that \begin{equation} \label{eq:goal:tau} \qquad\frac{1}{\sqrt\tau}\int_{u_t(0)}^{u_{t+\tau}(0)} \biggl(\frac{1}{\sigma(y)}-\frac{1}{\sigma(u_t(0))} \biggr)\,\d y \to0 \qquad\mbox{almost surely, as $\tau\downarrow0$}. \end{equation} Let $\mathcal{I}(t,t+\tau)$ denote the random closed interval with endpoints $u_t(0)$ and $u_{t+\tau}(0)$. Our strict positivity result (Lemma~\ref{lem:strict:pos}) implies that \begin{equation} \label{eq:I} \mathcal{I}(t ,t+\tau)\subset(0 ,\infty) \qquad\mbox{for all $t,\tau>0$ a.s.,} \end{equation} and thus paves way for the a.s. bounds \begin{eqnarray*} \label{eq:tau:bound} \biggl\llvert \int_{u_t(0)}^{u_{t+\tau}(0)} \biggl(\frac{1}{\sigma(y)}-\frac{1}{\sigma(u_t(0))} \biggr) \,\d y\biggr\rrvert &\le&\lip \,\cdot\,\frac{|u_t(0) - u_{t+\tau}(0)|^2}{ \inf_{y\in\mathcal{I}(t,t+\tau)}|\sigma(y)|^2} \\ &=&O \bigl(\tau\log\vert\log\tau\vert \bigr) \qquad(\tau\downarrow0); \end{eqnarray*} see \eqref{eq:local:LIL} for the last part. This implies \eqref{eq:goal:tau} and thus completes our proof for $m=1$. The proof for general $m$ is an easy adaption since $\{B(x_j)\}_{j=1}^m$ are i.i.d. Brownian motions. \end{pf*} \section{Preliminaries for the proof of Theorem \texorpdfstring{\protect\ref{th:as}}{2.7}} \label {sec:prelim thm1.6} The following function will play a prominent role in the ensuing analysis: \begin{equation} \label{eq:P:bar} \bar{P}(\tau) := \llVert p_\tau\rrVert _{\ell^2(\Z^d)}^2 =\sum_{x\in\Z^d} \bigl[p_\tau(x)\bigr]^2\qquad \mbox{for all $\tau\ge0$}. \end{equation} Because of the Chapman--Kolmogorov property, we can also think of $\bar{P}$ as \begin{equation} \label{eq:P:bar:replica} \bar{P}(\tau) := \P\bigl\{X_\tau- X_\tau' =0\bigr\}, \end{equation} where $X'$ is an independent copy of $X$. There is another useful way to think of $\bar{P}$ as well. Using the fact that \begin{equation} \label{eq:CHF} \E{\mathrm{e}}^{i\xi\,\cdot\, X_t} ={\mathrm{e}}^{-t(1-\varphi(\xi))} \qquad\mbox{for all $ \xi\in\R^d$ and $t\ge0$} \end{equation} and the Plancherel theorem, we see that \begin{eqnarray} \label{eq:P:bar:FT} \bar{P}(\tau) &=& (2\pi)^{-d}\int _{(-\pi,\pi)^d}\bigl\llvert \E\exp (i\xi\,\cdot\, X_\tau) \bigr \rrvert ^2 \,\d\xi \nonumber \\[-8pt] \\[-8pt] \nonumber &=&(2\pi)^{-d}\int_{(-\pi,\pi)^d}{\mathrm{e}}^{-2\tau(1-\Re\varphi(\xi ))} \,\d \xi, \end{eqnarray} where $\varphi(\xi)=\E[\exp(i\xi\,\cdot\, Z_1)]$; recall that $Z_1$ is the distribution of jump size. Therefore, in particular, the Laplace transform of $\bar{P}$ is \begin{eqnarray} \label{eq:Upsilon} \Upsilon(\beta) &:= &\int_0^\infty{\mathrm{e}}^{-\beta\tau}\bar {P}(\tau) \,\d\tau\qquad (\beta\ge0) \nonumber \\[-8pt] \\[-8pt] \nonumber &=&(2\pi)^{-d}\int_{(-\pi,\pi)^d}\frac{\d\xi}{\beta+2(1-\Re \varphi(\xi))}. \end{eqnarray} The interchange of the integrals is justified by Tonelli's theorem, since $1-\Re\varphi(\xi)\ge0$. Note that $\Upsilon(0)$ agrees with \eqref{Upsilon(0)}. Also, the classical theory of random walks tells us that $X-X'$ is transient if and only if $\Upsilon(0)=\int_0^\infty\bar{P}(\tau) \,\d\tau<\infty$, which is in turn equivalent to the condition \begin{equation} \int_{(-\pi,\pi)^d}\frac{\d\xi}{1-\Re\varphi(\xi)}<\infty; \end{equation} this is the Chung--Fuchs theorem \cite{ChungFuchs}, transliterated to the setting of\break continuous-time symmetric random walks, thanks to a standard Poissonization argument which we feel free to omit. \begin{lemma}\label{lem:L2:preserved} If $u_0\in\ell^2(\Z^d)$, then $u_t\in\ell^2(\Z^d)$ a.s. for all $t\ge0$. Moreover, for every $\beta \ge0$ such that $\lip^2\Upsilon(\beta)<1$, \begin{equation} \label{eq:E(E_t)} \E \bigl(\|u_t\|_{\ell^2(\Z^d)}^2 \bigr) \le\frac{\|u_0\|_{\ell ^2(\Z^d)}^2 {\mathrm{e}}^{\beta t}}{1-\lip^2\Upsilon(\beta)}\qquad \mbox{for all $t\ge0$}. \end{equation} \end{lemma} \begin{pf} Let $u^{(0)}_t(x) := u_0(x)$ for all $t\ge0$ and $x\in\Z^d$, and define $u^{(k)}$ to be the resulting $k$th-step approximation to $u$ via Picard iteration. It follows that \begin{eqnarray} \label{eq:mild:L2} &&\E \bigl(\bigl|u_t^{(n+1)}(x)\bigr|^2 \bigr)\nonumber \\ &&\qquad= \bigl\llvert (\tilde{p}_t*u_0 ) (x)\bigr\rrvert ^2 +\sum_{y\in\Z^d}\int_0^t \bigl[p_{t-s}(y-x)\bigr]^2\E \bigl(\bigl\llvert \sigma \bigl(u_s^{(n)}(y) \bigr)\bigr\rrvert ^2 \bigr) \,\d s \\ &&\qquad\le\bigl\llvert (\tilde{p}_t*u_0 ) (x)\bigr\rrvert ^2 + \lip^2\sum_{y\in\Z^d}\int _0^t\bigl[p_{t-s}(y-x) \bigr]^2\E \bigl( \bigl\llvert u_s^{(n)}(y) \bigr\rrvert ^2 \bigr) \,\d s.\nonumber \end{eqnarray} We may add over all $x\in\Z^d$ to deduce from this and Young's inequality that \begin{equation}\quad \label{eq:L2L2} \E \bigl( \bigl\llVert u^{(n+1)}_t\bigr \rrVert _{\ell^2(\Z^d)}^2 \bigr) \le\|u_0 \|_{\ell^2(\Z^d)}^2+\lip^2\int_0^t \bar{P}(t-s)\E \bigl(\bigl\llVert u^{(n)}_s\bigr\rrVert _{\ell^2(\Z^d)}^2 \bigr) \,\d s. \end{equation} Since $\Upsilon(\beta)=\beta^{-1}\int_0^\infty\exp(-s)\bar {P}(s/\beta ) \,\d s \le\beta^{-1}<\infty$, we can find $\beta>0$ large enough to guarantee that $\lip^2\Upsilon(\beta)<1$. We multiply both sides of \eqref{eq:L2L2} by $\exp(-\beta t)$---for this choice of $\beta$---and notice from \eqref{eq:L2L2} that \begin{equation} A_k := \sup_{t\ge0} \bigl[ {\mathrm{e}}^{-\beta t} \E \bigl(\bigl\llVert u^{(k)}_t\bigr\rrVert _{\ell^2(\Z^d)}^2 \bigr) \bigr] \qquad(k\ge0) \end{equation} satisfies \begin{equation} A_{n+1}\le\|u_0\|_{\ell^2(\Z^d)}^2 + \lip^2\Upsilon(\beta) A_n\qquad \mbox{for all $n\ge0$. } \end{equation} Since $A_0=\|u_0\|_{\ell^2(\Z^d)}^2$, the preceding shows that $\sup_{n\ge0}A_n$ is bounded above by $(1-\lip^2\Upsilon(\beta))^{-1}\|u_0\|_{\ell^2(\Z^d)}^2$. \end{pf} \begin{proposition}\label{pr:L2:int} If $u_0\in\ell^1(\Z^d)$, then for every $\beta\ge0$ such that\break $\lip^2\Upsilon(\beta)<1$, \begin{equation} \int_0^\infty{\mathrm{e}}^{-\beta t}\E \bigl( \|u_t\|_{\ell^2(\Z ^d)}^2 \bigr)\,\d t\le \frac{\|u_0\|_{\ell^1(\Z^d)}^2\Upsilon(\beta)}{1-\lip^2\Upsilon (\beta)}. \end{equation} Moreover, \begin{equation} \int_0^\infty{\mathrm{e}}^{-\beta t}\E \bigl( \|u_t\|_{\ell^2(\Z ^d)}^2 \bigr)\,\d t =\infty, \end{equation} for all $\beta\ge0$ such that $\ell_\sigma^2\Upsilon(\beta)\ge1$. \end{proposition} \begin{pf} We proceed as we did for Lemma \ref{lem:L2:preserved}. But instead of deducing \eqref{eq:L2L2} from \eqref{eq:mild:L2}, we use a different bound for $\|\tilde{p}_t*u_0\|_{\ell^2(\Z^d)}$ \begin{eqnarray} \label{eq:ell^2} &&\E \bigl( \bigl\llVert u^{(n+1)}_t \bigr\rrVert _{\ell^2(\Z^d)}^2 \bigr)\nonumber \\ &&\qquad \le\|p_t\|_{\ell^2(\Z^d)}^2\|u_0 \|_{\ell^1(\Z^d)}^2 +\lip^2\int_0^t \bar{P}(t-s)\E \bigl(\bigl\llVert u^{(n)}_s\bigr\rrVert _{\ell^2(\Z^d)}^2 \bigr) \,\d s \\ &&\qquad = \bar{P}(t)\|u_0\|_{\ell^1(\Z^d)}^2 + \lip^2\int_0^t \bar{P}(t-s)\E \bigl(\bigl\llVert u^{(n)}_s\bigr\rrVert _{\ell^2(\Z^d)}^2 \bigr) \,\d s, \nonumber \end{eqnarray} thanks to a slightly different application of Young's inequality. If we integrate both sides $[\exp(-\beta t)\,\d t]$, then we find that \begin{equation} I_k:=\int_0^\infty {\mathrm{e}}^{-\beta t}\E \bigl( \bigl\llVert u^{(k)}_t\bigr \rrVert _{\ell^2(\Z ^d)}^2 \bigr) \,\d t\qquad (k\ge0) \end{equation} satisfies \begin{eqnarray} I_{n+1} &\le&\|u_0\|_{\ell^1(\Z^d)}^2 \int_0^\infty{\mathrm{e}}^{-\beta t}\bar {P}(t) \,\d t+ I_n\times\lip^2\int_0^\infty{\mathrm{e}}^{-\beta t}\bar{P}(t) \,\d t \nonumber \\[-8pt] \\[-8pt] \nonumber &=& \|u_0\|_{\ell^1(\Z^d)}^2\Upsilon(\beta) + I_n \lip^2\Upsilon(\beta); \end{eqnarray} see \eqref{eq:Upsilon}. The first portion of the lemma follows from this, induction and Fatou's lemma since $\lip^2\Upsilon(\beta)<1$. Next, let us suppose that $\ell_\sigma^2\Upsilon(\beta)\ge1$. The following complimentary form of~\eqref{eq:ell^2} holds [for the same reasons that \eqref{eq:ell^2} held]: \begin{equation} \label{eq:LB:L2} \quad\E \bigl(\|u_t\|_{\ell^2(\Z^d)}^2 \bigr) \ge\llVert \tilde{p}_t*u_0\rrVert _{\ell^2(\Z^d)}^2 +\ell_\sigma^2 \int _0^t\bar{P}(t-s)\E \bigl(\|u_s \|_{\ell^2(\Z^d)}^2 \bigr)\,\d s. \end{equation} It is not hard to verify directly that \begin{equation} \label{eq:LB:p*u} \|\tilde{p}_t*u_0\|_{\ell^2(\Z^d)}^2 \ge u_0^2(x_0)\|p_t \|_{\ell^2(\Z^d)}^2, \end{equation} whence, by $u_0(x_0)>0$ for some $x_0>0$, it follows that \begin{equation} F(t) := \E \bigl(\|u_t\|_{\ell^2(\Z^d)}^2 \bigr)\qquad (t \ge0) \end{equation} solves the renewal inequality \begin{equation} F(t) \ge u_0^2(x_0)\bar{P}(t) + \ell_\sigma^2 \int_0^t \bar{P}(t-s)F(s) \,\d s. \end{equation} Therefore, $\tilde{F}(\beta):=\int_0^\infty\exp(-\beta t)F(t) \,\d t$ satisfies \begin{equation} \tilde{F}(\beta) \ge u_0^2(x_0)\Upsilon( \beta) + \ell_\sigma^2\Upsilon(\beta)\tilde{F}(\beta). \end{equation} Since $u_0(x_0)>0$ and $\Upsilon(\beta)>0$ for all $\beta\ge0$, it follows that $\tilde{F}(\beta)=\infty$ whenever $\ell_\sigma^2\Upsilon(\beta)\ge1$. \end{pf} \begin{proposition}\label{pr:bounded:preserved} If $u_0\in\ell^1(\Z^d)$, then \begin{equation} \label{eq:bounded:preserved} \sup_{t\ge0}\sup_{x\in\Z^d}u_t(x)< \infty,\qquad \sum_{y\in\Z^d}\int_0^\infty \bigl\llvert \sigma\bigl(u_s(y)\bigr) \bigr\rrvert ^2\,\d s<\infty\qquad\mbox{a.s.} \end{equation} Moreover: \textup{(i)} If, in addition, $q:=\lip^2\Upsilon(0)<1$, then \begin{eqnarray} \E \Bigl(\sup_{t\ge0}\sup_{x\in\Z^d} \bigl|u_t(x)\bigr|^2 \Bigr) &\le&\E \Bigl(\sup_{t\ge0} \|u_t\|_{\ell^1(\Z^d)}^2 \Bigr) \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&2\|u_0 \|_{\ell^1(\Z^d)}^2+\frac{8q \,\cdot\,\|u_0\|_{\ell^1(\Z ^d)}^2}{1-q}. \end{eqnarray} \textup{(ii)} If, in addition, $\ell_\sigma^2\Upsilon(0)\ge1$, then \begin{equation} \E \Bigl(\sup_{t\ge0}\|u_t\|_{\ell^1(\Z^d)}^2 \Bigr) =\int_0^\infty\E \bigl( \|u_s\|_{\ell^2(\Z^d)}^2 \bigr) \,\d s= \infty. \end{equation} \end{proposition} \begin{remark}\label{rem:liminf=0} Clearly, \eqref{eq:bounded:preserved} implies that if $u_0\in\ell^1(\Z^d)$, then \begin{equation} \liminf_{t\to\infty}\sup_{x\in\Z^d}\bigl|\sigma \bigl(u_t(x)\bigr)\bigr|^2 \le\liminf_{t\to\infty} \sum_{x\in\Z^d}\bigl|\sigma\bigl(u_t(x) \bigr)\bigr|^2=0 \qquad\mbox{a.s.} \end{equation} If, in addition, $\ell_\sigma>0$ (say), then we can deduce from the preceding fact that $\liminf_{t\to\infty}\sup_{x\in\Z^d} |u_t(x)|=0$ a.s. \end{remark} Recall that $X-X'$ is transient if and only if $\Upsilon(0)<\infty$. Therefore, in order for the condition $\lip^2\Upsilon(0)<1$ to hold, it is necessary---though not sufficient---that $X-X'$ be transient. \begin{pf*}{Proof of Proposition~\ref{pr:bounded:preserved}} First of all, Theorem~\ref{th:exist:unique} assures us that $u_t(x)\ge0$ a.s., and hence $\|u_t\|_{\ell^1(\Z^d)}=\sum_{x\in\Z^d}u_t(x)$. Therefore, if we add both sides of \eqref{mild}, then we find that \begin{equation} \label{eq:ell^1} \|u_t\|_{\ell^1(\Z^d)} = \|u_0 \|_{\ell^1(\Z^d)} + \sum_{y\in\Z^d}\int _0^t\sigma\bigl(u_s(y)\bigr) \,\d B_s(y). \end{equation} (It is easy to apply the moment bound of Theorem~\ref{th:exist:unique} to justify the interchange of the sum and the stochastic integral.) In particular, it follows that \begin{equation} \label{eq:M} M_t := \|u_t\|_{\ell^1(\Z^d)} \qquad(t\ge0) \end{equation} defines a nonnegative continuous martingale with mean $\|u_0\|_{\ell^1(\Z^d)}$. Its\break quadratic variation satisfies the following relations: \begin{equation} \langle M\rangle_t = \sum_{y\in\Z^d}\int _0^t \bigl|\sigma\bigl(u_s(y) \bigr)\bigr|^2 \,\d s\le \lip^2 \int_0^t \|u_s\|_{\ell^2(\Z^d)}^2 \,\d s. \end{equation} Bound \eqref{eq:Upper:LE} of Theorem~\ref{th:exist:unique} is more than enough to show that $M:=\{M_t\}_{t\ge0}$ is a continuous $L^2(\P)$ martingale. Since $M_t\ge0$ a.s. (Theorem~\ref{th:exist:unique}) it follows from the martingale convergence theorem that $\lim_{t\to\infty}M_t$ exists a.s. and is finite a.s., which proves the first part of (\ref {eq:bounded:preserved}). And therefore, $\langle M\rangle_\infty= \sum_{y\in\Z^d}\int_0^t|\sigma(u_s(y))|^2 \,\d s$ has to be also a.s. finite., since we can realize $M_t$ as $W(\langle M\rangle_t)$ for some Brownian motion $W$, thanks to the Dubins, Dambis-Schwartz representation theorem \cite{RevuzYor}, page 170. (i) If we know also that $\lip^2\Upsilon(0)<1$, then Proposition~\ref{pr:L2:int} guarantees that $\E\langle M\rangle_\infty$ is bounded from above by $(1-\lip^2\Upsilon(0))^{-1} \lip^2\Upsilon(0)\times\break \|u_0\|_{\ell^1(\Z^d)}^2<\infty$, whence it follows that $M:=\{M_t\}_{t\ge0}$ is a continuous $L^2(\P)$-bounded martingale with \begin{equation} \E \Bigl(\sup_{t\ge0}M_t^2 \Bigr) \le2 \|u_0\|_{\ell^1(\Z ^d)}^2+\frac{8 \lip^2\Upsilon(0) \,\cdot\,\|u_0\|_{\ell^1(\Z^d)}^2 }{1-\lip^2\Upsilon(0)}, \end{equation} thanks to Doob's maximal inequality. This proves part (i) because $\|u_t\|_{\ell^\infty(\Z^d)}$ is bounded above by $\|u_t\|_{\ell^1(\Z^d)}$. (ii) Finally consider the case $\ell_\sigma^2\Upsilon(0)\ge1$. Since \begin{eqnarray} \E \bigl(\|u_t\|_{\ell^1(\Z^d)}^2 \bigr)&=&\E \bigl( M_t^2 \bigr) =\| u_0 \|_{\ell^1(\Z^d)}^2 + \sum_{y\in\Z^d}\int _0^t\E \bigl(\bigl\llvert \sigma \bigl(u_s(y)\bigr)\bigr\rrvert ^2 \bigr) \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &\ge&\| u_0\|_{\ell^1(\Z^d)}^2 +\ell_\sigma^2 \int_0^t\E \bigl(\|u_s \|_{\ell^2(\Z^d)}^2 \bigr) \,\d s, \end{eqnarray} it suffices to show that this final integral is unbounded (as a function of $t$), which follows from the second part of Proposition~\ref{pr:L2:int}. \end{pf*} \begin{corollary}\label{co:iff} If $u_0\in\ell^1(\Z^d)$, then the following is a $\P$-null set: \begin{equation} \Bigl\{\omega\dvtx\lim_{t\to\infty}\sup_{x\in\Z^d} \bigl|u_t(x) (\omega)\bigr|=0 \Bigr\} \,\triangle\, \Bigl\{ \omega\dvtx \lim _{t\to\infty}\|u_t\|_{\ell^2(\Z ^d)}(\omega)=0 \Bigr\}. \end{equation} \end{corollary} \begin{pf} Let $E_1$ denote the event that $\lim_{t\to\infty}\sup_{x\in\Z^d} |u_t(x)|=0$ and $E_2$ the event that $\lim_{t\to\infty}\|u_t\|_{\ell^2(\Z^d)}=0$. Because of the real-variable bounds, $\|u_t\|_{\ell^\infty (\Z^d)}^2 \le\|u_t\|_{\ell^2(\Z^d)}^2 \le\|u_t\|_{\ell^\infty(\Z ^d)}\,\cdot\, \|u_t\|_{\ell^1(\Z^d)}$, we have \begin{equation} E_1\,\triangle \, E_2 \subseteq \Bigl\{ \omega\dvtx \limsup _{t\to\infty}\|u_t\|_{\ell^1(\Z ^d)}(\omega) =\infty \Bigr\}. \label{eq:subset:null} \end{equation} We have already noted, however, that $M_t:=\|u_t\|_{\ell^1(\Z^d)}$ defines a nonnegative martingale, under the conditions of this corollary. Therefore, the final event in \eqref{eq:subset:null} is $\P$-null, thanks to Doob's martingale convergence theorem. Thus we find that $E_1\,\triangle\, E_2$ is a measurable subset of a $\P$-null set, and is hence $\P$-null. \end{pf} \begin{proposition}\label{pr:Liap:L2} Suppose $u_0\in\ell^1(\Z^d)$ and the random walk $X$ is transient; that is, $\Upsilon(0)<\infty$. Then \begin{equation} \label{eq:Liap:L2:1} \limsup_{t\to\infty}\frac{1}t \log\E \bigl( \|u_t\|_{\ell^1(\Z ^d)}^2 \bigr) \le\inf \bigl\{\beta>0 \dvtx \lip^2\Upsilon(\beta)<1 \bigr\} <\infty. \end{equation} If, in addition, $\ell_\sigma^2\Upsilon(0)>1$, then \begin{equation} \label{eq:Liap:L2:2} \liminf_{t\to\infty}\frac{1}t \log\E \bigl( \llVert u_t\rrVert _ \ell^2(\Z^d)}^2 \bigr) \ge\inf \bigl\{\beta>0\dvtx \ell_\sigma^2\Upsilon(\beta)<1 \bigr\}>0. \end{equation} \end{proposition} \begin{pf} We have already proved a slightly weaker version of \eqref {eq:Liap:L2:1}. Indeed, since $\ell^1(\Z^d)\subset\ell^2(\Z^d)$, \eqref{eq:E(E_t)} implies that \begin{equation} \label{eq:Liap:L2:3} \limsup_{t\to\infty}\frac{1}t \log\E \bigl( \|u_t\|_{\ell^2(\Z ^d)}^2 \bigr) \le\inf \bigl\{\beta>0 \dvtx \lip^2\Upsilon(\beta)<1 \bigr\}. \end{equation} Then \eqref{eq:ell^1} and \eqref{eq:Liap:L2:3} together tell us that for every $C>\inf\{\beta>0\dvtx \break \lip^2\Upsilon(\beta)<1\}$, there exists $K=K(C)\in(0 ,\infty)$ such that \begin{eqnarray}\qquad \E \bigl(\|u_t\|_{\ell^1(\Z^d)}^2 \bigr) &\le& \|u_0\|_{\ell^1(\Z^d)}^2 +\lip^2\int _0^t \E \bigl(\|u_s \|_{\ell^2(\Z^d)}^2 \bigr) \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&\|u_0\|_{\ell^1(\Z^d)}^2 +K\int _0^t {\mathrm{e}}^{Cs} \,\d s = O \bigl( {\mathrm{e}}^{(C+o(1))t} \bigr) \qquad\mbox{as $t\to\infty$}. \end{eqnarray} Thus follows the first bound of the proposition. Because of \eqref{eq:LB:L2} and \eqref{eq:LB:p*u}, we find that \begin{equation} F(t) := \E \bigl(\llVert u_t\rrVert _{\ell^2(\Z^d)}^2 \bigr)\qquad (t\ge0) \end{equation} solves the renewal inequality \begin{equation} F(t) \ge g(t) + \int_0^t h(t-s)F(s) \,\d s\qquad (t \ge0), \end{equation} where \begin{equation} g(t) := u_0^2(x_0)\bar{P}(t),\qquad h(t):= \ell_\sigma^2\bar{P}(t)\qquad (t\ge0). \end{equation} A comparison result (Lemma~\ref{lem:RE:comparison}) tells us that $F(t)\ge f(t)$ for all $t\ge0$, where $f$ is the solution to the renewal equation \begin{equation} f(t) = g(t) + \int_0^t h(t-s)f(s) \,\d s\qquad (t \ge0). \end{equation} The condition that $\ell_\sigma^2\Upsilon(0)>1$ is equivalent to $\int_0^\infty h(t) \,\d t>1$. Because of transience $[\Upsilon (0)<\infty]$ and the fact that $\Upsilon(\beta)$ is strictly decreasing and continuous, we can find $\beta^{*}>0$ such that $\int_0^\infty\exp(-\beta^{*} t)h(t) \,\d t=1$. Note that $f_{\beta^{*}}(t):=\exp(-\beta^{*} t)f(t)$ solves the renewal equation \begin{equation} f_{\beta^*} (t) = g_{\beta^*}(t) + \int_0^t h_{\beta^*}(t-s) f_{\beta ^*}(s) \,\d s\qquad (t\ge0), \end{equation} where $g_{\beta^*}(t):=\exp(-\beta^* t)g(t)$ and $h_{\beta^*}(t):=\exp (-{\beta^*} t)h(t)$. Since $h_{\beta^*}$ is a probability density function, and $g_{\beta ^*}$ is nonincreasing [see \eqref{eq:P:bar:FT}], Blackwell's key renewal theorem \cite {FellerVolII} implies that \begin{eqnarray}\qquad \liminf_{t\to\infty}{\mathrm{e}}^{-\beta^* t}F(t) &\ge&\lim _{t\to \infty}f_{\beta^*}(t) = \biggl(\int_0^\infty sh_{\beta^*}(s) \,\d s \biggr)^{-1}\,\cdot\,\int_0^\infty g_{\beta^*}(s) \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &=& u_0^2(x_0)\ell_\sigma^{-2} \biggl(\int_0^\infty s{\mathrm{e}}^{-\beta^* s} \bar{P}(s) \,\d s \biggr)^{-1}\,\cdot\,\Upsilon\bigl(\beta^*\bigr). \end{eqnarray} Since $\bar{P}(s)\le1$, the right-most quantity is at least $u_0^2(x_0) \ell_\sigma^{-2} (\beta^{*})^2\Upsilon(\beta^*)>0$. This completes the proof of \eqref{eq:Liap:L2:2}. Note that we have used the fact that $\Upsilon(\beta)$ is continuous in $\beta$ and strictly decreasing, so that $\beta^{*}=\inf\{\beta>0\dvtx\ell_{\sigma }^2\Upsilon (\beta)<1\}>0$. \end{pf} \begin{proposition}\label{pr:L2:WeakDisorder} If $u_0 \in\ell^1(\Z^d)$ and $\lip^2\Upsilon(0)<1$, then\break $\lim_{t\to\infty}\E(\|u_t\|_{\ell^2(\Z^d)}^2)=0$. Furthermore, as $t\to\infty$, \begin{eqnarray} \label{ass:assert} \bar{P}(t) &=&O \bigl(\E \bigl( \|u_t \|_{\ell^2(\Z^d)}^2 \bigr) \bigr)\quad \mbox{and} \nonumber \\[-8pt] \\[-8pt] \nonumber \qquad\E \bigl(\|u_t\|_{\ell^2(\Z^d)}^2 \bigr)& =&O \bigl(t^{-\alpha } \bigr)\qquad \mbox{for all $\alpha\ge0$ such that $\bar{P}(t)=O \bigl(t^{-\alpha}\bigr)$}. \end{eqnarray} \end{proposition} \begin{pf} The first assertion of \eqref{ass:assert} is simple to prove; in fact,\break $\E(\|u_t\|_{\ell^2(\Z^d)}^2) \ge [u_0(x_0)]^2 \bar{P}(t)$ $(t\ge0)$ for any $x_0\in\Z^d$ and all $t>0$; see \eqref{eq:LB:L2} and \eqref{eq:LB:p*u}. We concentrate our efforts on the remaining statements. Thanks to \eqref{eq:ell^2}, \begin{equation}\qquad \E \bigl( \| u_t\|_{\ell^2(\Z^d)}^2 \bigr) \le \bar{P}(t) \|u_0\| _{\ell ^1(\Z^d)}^2 + \lip^2\int_0^t\bar{P}(t-s)\E \bigl( \|u_s\|_{\ell^2(\Z ^d)}^2 \bigr) \,\d s. \end{equation} That is, $F(t):=\E(\|u_t\|_{\ell^2(\Z^d)}^2)$ is a sub solution to a renewal equation; namely, \begin{equation} \label{eq:subcrit:Fgh} F(t) \le g(t) +\int_0^t h(t-s)F(s) \,\d s\qquad (t\ge0), \end{equation} for \begin{equation} g(t) := \bar{P}(t)\|u_0\|_{\ell^1(\Z^d)}^2,\qquad h(t) := \lip^2\bar{P}(t). \end{equation} A comparison lemma (Lemma~\ref{lem:RE:comparison}) shows that $0\le F(t)\le f(t)$ for all $t\ge0$, where \begin{equation} f(t) = g(t) +\int_0^t h(t-s)f(s) \,\d s\qquad (t \ge0). \end{equation} Therefore, it remains to prove that $f(t)\to0$ as $t\to\infty$. It is easy, as well as classical, that we can write $f$ in terms of the renewal function of $h$; that is, \begin{equation} \label{eq:f:g:h} f(t) = g(t) + \sum_{n=0}^\infty \int_0^t h^{*(n)}(s)g(t-s) \,\d s\qquad (t \ge0), \end{equation} where $h^{*(1)}(t):=\int_0^t h(t-s)h(s) \,\d s$ denotes the convolution of $h$ with itself, and $h^{*(k+1)}(t) :=\int_0^t h^{*(k)}(t-s)h(s) \,\d s$ for all $k\ge0$. We might note that $g(t)\le g(0)=\|u_0\|_{\ell^2(\Z^d)}^2$ because $\bar{P}$ is nonincreasing [see \eqref{eq:P:bar:FT}] and one at zero. Therefore, \begin{eqnarray} \label{eq:h*h} 0&\le&\int_0^t h^{*(n)}(s)g(t-s) \,\d s \le\|u_0\|_{\ell^2(\Z^d)}^2 \int_0^\infty h^{*(n)}(s) \,\d s \nonumber \\ &\le&\|u_0\|_{\ell^2(\Z^d)}^2 \biggl(\int _0^\infty h(s) \,\d s \biggr)^{n+1}\qquad \mbox{[Young's inequality]} \\ &=&\|u_0\|_{\ell^2(\Z^d)}^2 \bigl( \lip^2 \Upsilon(0) \bigr)^{n+1}.\nonumber \end{eqnarray} \upqed\end{pf} It is not hard to see that $\lim_{t\to\infty}g(t)= \lim_{t\to\infty}\bar{P}(t)=0$; this follows from \eqref{eq:P:bar:FT} and the monotone convergence theorem. Because $\lip^2\Upsilon(0)<1$, we can deduce from \eqref{eq:h*h} and \eqref {eq:f:g:h}, in conjunction with the dominated convergence theorem, that $f(t)$---hence $F(t)=\E(\|u_t\|_{\ell^2(\Z^d)}^2)$---converges to zero as $t\to \infty$. It remains to prove the second assertion in \eqref{ass:assert}. With this in mind, let us suppose $\bar{P}$ satisfies the following: There exists $c\in(0 ,\infty)$ and $\alpha\in[0 ,\infty)$ such that \begin{equation} \bar{P}(t) \le c(1+t)^{-\alpha}, \end{equation} for there is nothing to consider otherwise. We aim to prove that \begin{equation} \label{eq:UB:powerdecay} \E \bigl(\|u_t\|_{\ell^2(\Z^d)}^2 \bigr) \le\operatorname {const}\cdot\, (1+t)^{-\alpha}, \end{equation} for some finite constant that does not depend on $t$. This proves the proposition. Define $F_k(t):=\E(\|u_t^{(k)}\|_{\ell^2(\Z^d)}^2)$, where $u^{(k)}$ denotes the $k$th approximation to $u$ via Picard's iteration \eqref {eq:Picard}, starting at $u^{(0)}_t(x)\equiv0$. We can write \eqref{eq:ell^2}, in short hand, as follows: \begin{equation} \label{eq:ell^2:recur} F_{n+1}(t) \le\bar{P}(t)\|u_0 \|_{\ell^1(\Z^d)}^2 + \lip^2\int_0^t \bar {P}(t-s) F_n(s) \,\d s. \end{equation} Now let us choose and fix $\varepsilon\in(0 ,1)$ and write \begin{eqnarray} &&\int_0^t \bar{P}(t-s) F_n(s) \,\d s\nonumber\\ &&\qquad= \int_{t\varepsilon}^t \bar{P}(s) F_n(t-s) \,\d s +\int_{t(1-\varepsilon)}^t \bar{P}(t-s) F_n(s) \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le c\int_{t\varepsilon}^t \frac{F_n(t-s)}{(1+s)^\alpha} \,\d s+ \sup_{w\ge0} \bigl[(1+w)^\alpha F_n(w) \bigr]\int_{t(1-\varepsilon)}^t \frac{\bar{P}(t-s)}{(1+s)^\alpha} \,\d s \\ &&\qquad\le\frac{c}{\varepsilon^\alpha(1+t)^\alpha}\int_0^\infty F_n(s) \,\d s+ \sup_{w\ge0} \bigl[(1+w)^\alpha F_n(w) \bigr]\frac{\Upsilon(0)}{ (1-\varepsilon)^\alpha(1+t)^\alpha}.\nonumber \end{eqnarray} The proof of Proposition~\ref{pr:L2:int} shows that \begin{equation} \label{eq:F_n:recur} \sup_{n\ge0} \int_0^\infty F_n(s) \,\d s \le\frac{\|u_0\|_{\ell^1(\Z ^d)}^2\Upsilon(0)}{ 1-\lip^2\Upsilon(0)}. \end{equation} Consequently, \begin{equation} R_k:= \sup_{w\ge0} \bigl[ (1+w)^\alpha F_k(w) \bigr]\qquad (k\ge0) \end{equation} satisfies \begin{equation} R_{n+1} \le A + R_n\frac{\lip^2\Upsilon(0)}{(1-\varepsilon)^\alpha} \qquad\mbox{for all $n \ge0$}, \end{equation} where \begin{equation} A = A(\varepsilon) := c\|u_0\|_{\ell^1(\Z^d)}^2 + \frac{c\|u_0\|_{\ell^1(\Z^d)}^2\lip^2\Upsilon(0)}{\varepsilon ^\alpha( 1-\lip^2\Upsilon(0))}. \end{equation} Since $\lip^2\Upsilon(0)<1$, we can choose $\varepsilon$ sufficiently close to zero to ensure that $\lip^2\Upsilon(0)<(1-\varepsilon)^{1+\alpha}$. For this particular $\varepsilon$, we find that $R_{n+1}\le A + (1-\varepsilon)R_n$ for all $n$. Since $R_0=0$, this proves that $\sup_{n\ge0} R_n\le A/\varepsilon$. Equation \eqref{eq:UB:powerdecay}---whence the proposition---follows from the latter inequality and Fatou's lemma. \section{Proof of Theorem \texorpdfstring{\protect\ref{th:as}}{2.7}} \label{sec:thm1.6} Let us begin with an elementary real-variable inequality. \begin{lemma}\label{lem:arith} For all real numbers $k\ge2$ and $x,y,\delta>0$, \begin{equation} \label{eq:arith} (x+y)^k \le(1+\delta)^{k-1} x^k + \biggl(\frac{1+\delta}{\delta } \biggr)^{k-1} y^k. \end{equation} \end{lemma} This is a consequence of Jensen's inequality when $\delta=1$. We are interested in the regime $\delta\ll1$. \begin{pf} The function $f(z) := (z+1)^k - (1+\delta)^{k-1} z^k$ $(z>0)$ is maximized at $z_* := \delta^{-1}$, and $\max_z f(z) = f(z_*)=\{ (1+\delta)/\delta\}^{k-1}$; that is, $f(x) \le\{(1+\delta)/\delta\}^{k-1}$ for all $x>0$. This is the desired result when $y=1$. We can factor the variable $y$ from both sides of \eqref{eq:arith} in order to reduce the problem to the previously proved case $y=1$. \end{pf} \begin{lemma}\label{lem:transient:ell^2} $\int_0^\infty\llVert p_{s+\tau}-p_s\rrVert _{\ell^2(\Z^d)}^2 \,\d s \le4\Upsilon(0)\tau^2$ for all $\tau\ge0$. \end{lemma} \begin{pf} We apply the Plancherel theorem and \eqref{eq:CHF} in order to deduce that \begin{eqnarray} \llVert p_{s+\tau}-p_s \rrVert _{\ell^2(\Z^d)}^2 &=& (2\pi)^{-d}\int_{(-\pi,\pi)^d} \bigl\llvert {\mathrm{e}}^{-(s+\tau)(1-\varphi (\xi))} -{\mathrm{e}}^{-s(1-\varphi(\xi))}\bigr\rrvert ^2\,\d\xi \nonumber \\ &=&(2\pi)^{-d}\int_{(-\pi,\pi)^d}{\mathrm{e}}^{-2s(1-\Re\varphi(\xi))} \bigl\llvert 1-{\mathrm{e}}^{-\tau(1-\varphi(\xi))}\bigr\rrvert ^2\,\d\xi \\ &\le&\frac{4\tau^2}{(2\pi)^d}\int_{(-\pi,\pi)^d} {\mathrm{e}}^{-2s(1-\Re\varphi (\xi))} \,\d \xi.\nonumber \end{eqnarray} Integrate $[\d s]$ to finish; compare with \eqref{eq:Upsilon}. \end{pf} Recall that $z_k$ denotes the optimal constant in BDG inequality \eqref{eq:Davis}. \begin{lemma}\label{lem:ell^k:subcrit} If $k\in(2 ,\infty)$ satisfies $z_k \lip\sqrt{\Upsilon (0)}<(1+\delta )^{-(k-1)/k}$ for some $\delta>0$, then \begin{equation} \sup_{t\ge0}\E \Bigl( \sup_{x\in\Z^d}\bigl|u_t(x)\bigr|^k \Bigr)\le \sup_{t\ge0}\E \bigl( \llVert u_t\rrVert _{\ell^k(\Z^d)}^k \bigr)<\infty. \end{equation} \end{lemma} \begin{pf} Let $u^{(0)}_t(x):=u_0(x)$, and define $u^{(n)}$ to be the $n$th step Picard approximation to $u$, as in \eqref{eq:Picard}. Define \begin{equation} \bar{M}_t^{(n)} := \E \bigl( \bigl\llVert u^{(n)}_t\bigr\rrVert _{\ell^k(\Z ^d)}^k \bigr)\qquad \mbox{for all $t\ge0$ and $k\ge1$}. \end{equation} Then we can apply Lemma~\ref{lem:arith} and writ \begin{equation} \label{eq:bar{M}} \bar{M}_t^{(n+1)} \le \biggl( \frac{1+\delta}{\delta} \biggr)^{k-1}\sum_{x\in\Z^d} I_x + (1+\delta)^{k-1}\sum_{x\in\Z^d} J_x, \end{equation} where $I_x$ and $J_x$ were defined earlier in \eqref{eq:IJ}. One estimates $\sum_{x\in\Z^d}I_x$ via Jensen's inequality, using $p_t(\bullet-x)$ as the base measure, in order to find that \begin{equation} \label{eq:bar:Ix} \sum_{x\in\Z^d} I_x \le \|u_0\|_{\ell^k(\Z^d)}^k. \end{equation} In order to estimate $\sum_{x\in\Z^d}J_x$, we define---for all $(t ,x)\in\R_+\times\Z^d$---a Borel measure $\rho_{t,x}$ on $\R _+\times\Z^d$ as follows: \begin{equation} \label{eq:rho} \rho_{t,x}(\d s \,\d y) := \bigl[p_{t-s}(y-x) \bigr]^2 \1_{[0,t]}(s) \,\d s \chi(\d y); \end{equation} where $\chi$ denotes the counting measure on $\Z^d$. Because of the transience of $X-X'$, the measure $\rho_{t,x}$ is finite; in fact, \begin{equation} \label{mass:rho} \rho_{t,x}\bigl(\R_+\times\Z^d\bigr) = \int _0^t\|p_s\|_{\ell^2(\Z^d)}^2 \,\d s = \int_0^t\bar{P}(s) \,\d s \le \Upsilon(0). \end{equation} Therefore, we apply \eqref{eq:Jx:prec} and Jensen's inequality, in conjunction, in order to see that \begin{eqnarray} J_x &\le& z_k^k \biggl( \lip^2\int_{[0,t]\times\Z^d} \bigl\{ \E \bigl(\bigl\llvert u^{(n)}_s(y)\bigr\rrvert ^k \bigr) \bigr \}^{2/k} \rho_{t,x}(\d s \,\d y) \biggr)^{k/2} \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&(z_k\lip)^k \bigl[\Upsilon(0)\bigr]^{(k-2)/2} \int_{[0,t]\times\Z^d} \E \bigl( \bigl\llvert u^{(n)}_s(y) \bigr\rrvert ^k \bigr) \rho_{t,x}(\d s \,\d y). \end{eqnarray} Thus \begin{eqnarray} \sum_{x\in\Z^d}J_x & \le&(z_k\lip)^k \bigl[\Upsilon(0)\bigr]^{(k-2)/2}\int _0^t \bar{P}(t-s)\E \bigl(\bigl\llVert u^{(n)}_s\bigr\rrVert _{ \ell^k(\Z^d)}^k \bigr) \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &\le& \bigl(z_k\lip\sqrt{\Upsilon(0)} \bigr)^k\,\cdot\, \sup _{r\ge0}\E \bigl(\bigl\llVert u^{(n)}_r \bigr\rrVert _{ \ell^k(\Z^d)}^k \bigr), \end{eqnarray} thanks to \eqref{eq:Upsilon}. In summary, \eqref{eq:bar{M}} has the following consequence: For all $n\ge0$, \begin{eqnarray}\qquad &&\sup_{t\ge0}\bar{M}^{(n+1)}_t \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le \biggl(\frac{1+\delta}{\delta} \biggr)^{k-1} \|u_0 \|_{\ell^k(\Z^d)}^k + (1+\delta)^{k-1} \bigl(z_k \lip\sqrt{\Upsilon(0)} \bigr)^k \sup_{t\ge0} \bar{M}^{(n)}_t. \end{eqnarray} Since $(1+\delta)^{k-1}(z_k\lip\sqrt{\Upsilon(0)})^k<1$ and $\sup_{t\ge0}\bar{M}^{(0)}_t=\|u_0\|_{\ell^k(\Z^d)}^k$, this\break shows that $C:=\sup_{n\ge0}\sup_{t\ge0}\bar{M}^{(n)}_t<\infty$. Fatou's lemma now implies half of the result, since it shows that $\E(\|u_t\|_{\ell^k(\Z ^d)}^k)\le \liminf_{n\to\infty}\E(\|u_t^{(n)}\|_{\ell^k(\Z^d)}^k)\le C$. The remainder of the proposition follows simply because $\|\bullet\|_{\ell^\infty(\Z ^d)}\le \|\bullet \|_{\ell^k(\Z^d)}$. \end{pf} \begin{proposition}\label{pr:ell^k:modulus} Assume that $\limsup_{t\to\infty}t^{\alpha}\P\{X_t=X'_t\}<1$ for some \mbox{$\alpha>1$}, where $X$ and $X'$ are two independent random walks with generator $\mathscr{L}$. If $k\in(2 ,\infty)$ satisfies $z_k \lip\sqrt{\Upsilon(0)}< (1+\delta)^{-(k-1)/k}$ for some $\delta>0$, then there exists a finite constant $A$---depending only on $\delta$, $\lip$, $\Upsilon(0)$ and $\|u_0\|_{\ell^1(\Z ^d)}$---such that \begin{equation} \label{eq:ell^k:modulus} \E \bigl( \llVert u_{t+\tau}-u_t\rrVert _{\ell^k(\Z^d)}^k \bigr) \le\frac{A\tau^{k/2}}{(1+t)^{\alpha}}\qquad \mbox{for every $t, \tau\ge0$}. \end{equation} Consequently, there exists a H\"older-continuous modification of the process $t\mapsto u_t(\bullet)$ with values in $\ell^\infty(\Z ^d)$. Moreover, for this modification, there is a finite constant $A'$---depending only on $\delta$, $\lip$, $\Upsilon(0)$ and $\|u_0\|_{\ell^1(\Z ^d)}$---such that \begin{equation} \E \biggl( \sup_{ s\neq r\in[t,t+1]} \sup_{x\in\Z^d}\biggl \llvert \frac{u_r(x)-u_s(x)}{|r-s|^\eta}\biggr\rrvert ^k \biggr)\le \frac{A'}{(1+t)^{\alpha}}, \end{equation} as long as $0\le\eta< (k-2)/(2k)$. \end{proposition} \begin{pf} Thanks to Lemma~\ref{lem:ell^k:subcrit}, $\|u_t\|_{\ell^k(\Z^d)}$ has a finite $k$th moment. This observation justifies the use of these moments in the ensuing discussion. Now we begin our proof in earnest. The proof requires us to make a few small adjustments to the derivation of Lemma~\ref{lem:modulus}; specifically we now incorporate the fact that $\lip^2\Upsilon(0)<1$ into that proof. Therefore, we mention only the required changes. We use the notation of the proof of Lemma~\ref{lem:modulus} and write \begin{equation} \bigl\{ \E \bigl(\bigl|u_{t+\tau}(x)-u_t(x)\bigr|^k \bigr) \bigr\}^{1/k} \le |Q_1|+Q_2+Q_3, \end{equation} whence \begin{equation} \E \bigl(\bigl|u_{t+\tau}(x)-u_t(x)\bigr|^k \bigr) \le3^{k-1} \bigl( |Q_1|^k + Q_2^k + Q_3^k \bigr). \end{equation} Note that \begin{eqnarray} \nonumber &&\sum_{x\in\Z^d}|Q_1|^k\\[-2pt] &&\qquad \le\sum_{x\in\Z^d} \biggl( \sum _{y\in\Z^d} u_0(y) \bigl\llvert p_{t+\tau}(y-x)-p_t(y-x) \bigr\rrvert \biggr)^k \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le\|u_0\|_{\ell^1(\Z^d)}^{k-1}\,\cdot\, \sum _{x\in\Z^d}\sum_{y\in\Z^d} u_0(y) \bigl\llvert p_{t+\tau}(y-x)-p_t(y-x) \bigr\rrvert ^k \\ \nonumber &&\qquad=\|u_0\|_{\ell^1(\Z^d)}^k\,\cdot\,\llVert p_{t+\tau} - p_t\rrVert _{\ell^k(\Z^d)}^{k} \le \|u_0\|_{\ell^1(\Z^d)}^k\,\cdot\, \llVert p_{t+\tau}-p_t\rrVert _{\ell^2(\Z^d)}^{k} , \end{eqnarray} thanks to Jensen's inequality. We observe that \begin{eqnarray} \label{eq:p:diff} && \| p_{t+\tau}-p_t \|_{\ell^2(\Z^d)}^2 \nonumber\\ &&\qquad= (2\pi)^{-d} \int _{[-\pi ,\pi]^d} \bigl\llvert {\mathrm{e}}^{-t(1-\varphi(\xi))}\bigr\rrvert ^2 \bigl\llvert {\mathrm{e}}^{-\tau(1-\varphi(\xi))}-1\bigr\rrvert ^2 \,\d\xi \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le\operatorname{const}\,\cdot\,\tau^2 \int _{[-\pi,\pi]^d} \bigl\llvert {\mathrm{e}}^{-t(1-\varphi(\xi))}\bigr\rrvert ^2\,\d\xi =\operatorname{const}\,\cdot\,\tau^2\P\bigl \{X_t=X'_t\bigr\} \\ &&\qquad\le\operatorname{const}\,\cdot\,\frac{\tau^2}{(1+t)^{\alpha} }.\nonumber \end{eqnarray} Consequently, \begin{equation} \label{eq:Q_1:bis} \sum_{x\in\Z^d}|Q_1|^k \le\frac{\operatorname{const}\,\cdot\,\tau ^k}{(1+t)^{\alpha k/2}}. \end{equation} We estimate $Q_2$ slightly differently from the proof of Lemma~\ref {lem:modulus} as well. For every $(t ,x)\in\R_+\times\Z^d$, let us define a similar Borel measure $R_{t,x}$ to $\rho_{t,x}$ [see \eqref{eq:rho}] as follows: \begin{equation} R_{t,x}(\d s \,\d y) := \bigl[p_{t+\tau-s}(y-x)-p_{t-s}(y-x) \bigr]^2 \1_{[0,t]}(s) \,\d s \chi(\d y). \end{equation} Now we reexamine the first line of \eqref{eq:Q2:prec} and note that \begin{eqnarray}\qquad Q_2^2&\le&(z_k\lip)^2 \sum_{y\in\Z^d}\int_0^t \bigl[ p_{t+\tau -s}(y-x)-p_{t-s} (y-x) \bigr]^2 \bigl \{ \E \bigl(\bigl|u_s(y)\bigr|^k \bigr) \bigr\}^{2/k}\,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber & =&(z_k\lip)^2\int_{\R_+\times\Z^d} \bigl\{ \E \bigl(\bigl|u_s(y)\bigr|^k \bigr) \bigr\}^{2/k} R_{t,x}(\d s \,\d y). \end{eqnarray} This follows from \eqref{eq:sigma(0)=0} and \eqref{eq:Q2:prec}, but we use the optimal constant $z_k$ in place of the slightly weaker $2\sqrt{k}$ that came from Lemma~\ref{lem:BDG}. Lemma~\ref{lem:transient:ell^2} implies that $R_{t,x}(\R_+\times\Z^d) = \int_0^t \llVert p_{s+\tau}-p_s\rrVert _{ \ell^2(\Z^d)}^2\,\d s \le4\Upsilon(0)\tau^2$. This bound and Jensen's inequality together show that $\sum_{x\in\Z^d} Q_2^k$ is bounded from above by \begin{eqnarray}\label{eq:Q2:bound} &&(z_k\lip)^k \bigl(4\Upsilon(0)\tau^2 \bigr)^{(k-2)/2} \sum_{x\in\Z^d}\int _{\R_+\times\Z^d}\E \bigl(\bigl\llvert u_s(y) \bigr\rrvert ^k \bigr) R_{t,x}(\d s \,\d y) \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad=(z_k\lip)^k \bigl(4\Upsilon(0) \tau^2\bigr)^{(k-2)/2} \int_0^t \llVert p_{t+\tau -s}-p_{t-s} \rrVert _{\ell^2(\Z^d)}^2 \E \bigl( \|u_s\|_{\ell^k(\Z^d)}^k \bigr) \,\d s. \end{eqnarray} By an argument similar to the one used in Proposition~\ref{pr:L2:WeakDisorder}, one is able to show that $\E( \|u_s\|_{\ell^k(\Z^d)}^k) \le \operatorname{const}\,\cdot\,(1+s)^{-\alpha}$. Here is an outline of the proof: We can follow the proof of Lemma~\ref{lem:ell^k:subcrit}, but derive a better bound on $\sum_{x\in\Z^d} I_x \le\bar{P}(t) \|u_0\|_{\ell^1(\Z^d)}^k $, in order to obtain \begin{eqnarray} \label{eq:ell^k:prop8.2} &&\E \bigl(\bigl\|u_t^{(n+1)} \bigr\|_{\ell^k(\Z^d)}^k \bigr) \nonumber\\ &&\qquad \le \biggl(\frac{1+\delta}{\delta} \biggr)^{k-1} \bar{P}(t) \|u_0\|_{\ell^1(\Z^d)}^k \\ &&\qquad\quad{} +(1+\delta)^{k-1}(z_k\lip)^k \bigl[\Upsilon (0)\bigr]^{(k-2)/2}\int_0^t \bar{P}(t-s)\E \bigl(\bigl\llVert u^{(n)}_s\bigr\rrVert _{ \ell^k(\Z^d)}^k \bigr) \,\d s.\nonumber \end{eqnarray} From here, we proceed along similar lines, as we did from \eqref {eq:ell^2:recur} onward. We follow the proof of Proposition~\ref{pr:L2:int}, using \eqref {eq:ell^k:prop8.2}, in order to derive the following analog of \eqref{eq:F_n:recur}: \begin{equation} \sup_{n\ge0} \int_0^{\infty} F_n(s) \,\d s \le \frac{((1+\delta)/\delta)^{k-1} \|u_0\|_{\ell^1(\Z^d)}^k \Upsilon(0) }{1- (1+\delta)^{k-1}(z_k\lip \Upsilon(0))^k }, \end{equation} where $F_n(t):=\E(\|u_t^{(n+1)}\|_{\ell^k(\Z^d)}^k)$. In this way, we can obtain the bound $\E( \|u_s\|_{\ell^k(\Z^d)}^k) \le\operatorname{const}\cdot\, (1+s)^{-\alpha}$, as was needed. We use this bound, as well as~\eqref{eq:p:diff} in \eqref{eq:Q2:bound}, and split the integral into two parts ($0$ to $t/2$ and $t/2$ to $t$), in order to obtain the following: \begin{equation} \label{eq:Q_2:bis} \sum_{x\in\Z^d} Q_2^k \le\frac{\operatorname{const}\cdot\,\tau ^k}{(1+t)^{\alpha}}. \end{equation} Finally we estimate $\sum_{x\in\Z^d}Q_3^k$ by first modifying \eqref {eq:Q_3:prec} as follows: \begin{eqnarray} Q_3^2 &\le&(z_k\lip)^2 \sum_{y\in\Z^d} \int_t^{t+\tau} \bigl[ p_{t+\tau-s}(y-x) \bigr]^2 \bigl\{ \E \bigl( \bigl\llvert u_s(y)\bigr\rrvert ^k \bigr) \bigr\}^{2/k} \,\d s \nonumber \\[-8pt] \\[-8pt] \nonumber &=&(z_k\lip)^2\int_{\R_+\times\Z^d} \bigl\{ \E \bigl( \bigl\llvert u_s(y)\bigr\rrvert ^k \bigr) \bigr \}^{2/k} \mathcal{R}_{t,\tau,x}(\d s \,\d y), \end{eqnarray} where the Borel measures $\mathcal{R}_{t,\tau,x}$ are defined in a similar manner as in \eqref{eq:rho}; that is, \begin{equation} \mathcal{R}_{t,\tau,x} (\d s \,\d y) := \sum_{y\in\Z^d} \bigl[ p_{t+\tau -s}(y-x) \bigr]^2 \1_{[t,t+\tau]}(s) \,\d s \chi(\d y). \end{equation} Because $\mathcal{R}_{t,\tau,x}(\R_+\times\Z^d) = \int_0^\tau \bar{P}(s)\,\d s\le\tau$, Jensen's inequality assures us that \begin{eqnarray} \label{eq:Q_3:bis} \sum_{x\in\Z^d}Q_3^k &\le&(z_k\lip)^k\tau^{(k-2)/2}\sum _{x\in\Z^d} \int_{\R_+\times\Z^d} \E \bigl(\bigl\llvert u_s(y)\bigr\rrvert ^k \bigr) \mathcal{R}_{t,\tau,x}( \,\d s \,\d y) \nonumber \\ &=&(z_k\lip)^k\tau^{(k-2)/2}\int _t^{t+\tau} \bar{P}(t+\tau-s) \E \bigl( \llVert u_s\rrVert _{\ell^k(\Z^d)}^k \bigr) \,\d s \\ &\le&\frac{\operatorname{const}\cdot\,\tau^{{k}/{2}}}{(1+t)^{\alpha}}, \nonumbe \end{eqnarray} thanks to the bounds $\E( \|u_s\|_{\ell^k(\Z^d)}^k) \le\operatorname{const} \,\cdot\,(1+s)^{-\alpha}$ and $\bar{P}(t+\tau-s)\le1$. Since $\|u_0\|_{\ell^k(\Z^d)}\le\|u_0\|_{\ell^1(\Z^d)}$, displays \eqref{eq:Q_1:bis}, \eqref{eq:Q_2:bis} and \eqref{eq:Q_3:bis} together imply~\eqref{eq:ell^k:modulus}. This yields the first estimate of the proposition. The remaining assertions follow~\eqref{eq:ell^k:modulus}, using a suitable form of the Kolmogorov continuity theorem \cite{RevuzYor}, Theorem~2.1, page 25, and the fact that $\sup_{x\in\Z^d}|u_t(x)-u_s(x)|\le\|u_t-u_s\|_{\ell^k(\Z^d)}$. \end{pf} \begin{pf*}{Proof of Theorem~\ref{th:as}} We apply Proposition~\ref{pr:L2:WeakDisorder} and Chebyshev's inequality in conjunction in order to see that \begin{eqnarray} \sum_{n=1}^\infty\P \Bigl\{ \sup _{x\in\Z^d} \bigl|u_{n}(x)\bigr| > \varepsilon \Bigr\} &\le& \frac{1}{\varepsilon^2} \sum_{n=1}^\infty \E \bigl(\|u_{n}\|_{\ell^2(\Z^d)}^2 \bigr) \nonumber \\[-8pt] \\[-8pt] \nonumber &\le&\frac{\operatorname{const}}{\varepsilon^2}\,\cdot\, \sum_{n=1}^\infty n^{-\alpha}<\infty. \end{eqnarray} Therefore, the Borel--Cantelli lemma implies that \begin{equation} \label{eq:dissipate:subseq} \lim_{n\to\infty} \sup_{x\in\Z^d}\bigl|u_{n}(x)\bigr|=0\qquad \mbox{a.s.} \end{equation} We next note that the Burkholder's constants $z_k$ vary continuously for $k\ge2$, and $z_2=1$ is the minimum; see Davis \cite{Davis}. Davis \cite{Davis} obtains $z_k$ as the largest positive zero of the parabolic cylinder function of parameter $k$ and this varies continuously in $k$; see Abramowitz and Stegun \cite{AS}. If $\lip\sqrt{\Upsilon(0)}<1$, we can find $k>2$ and $\delta>0$ such that \begin{equation} \label{cond:k} z_k\lip\sqrt{\Upsilon(0)}<(1+\delta)^{-(k-1)/k}. \end{equation} We can now use Proposition~\ref{pr:ell^k:modulus} (with $\eta=0$) along with Chebyshev's inequality to control the spacings \begin{eqnarray} &&\P \Bigl\{ \sup_{s\in[n,n+1]}\sup_{x\in\Z^d} \bigl \llvert u_s(x) - u_{n}(x)\bigr\rrvert >\varepsilon \Bigr \} \nonumber \\[-8pt] \\[-8pt] \nonumber &&\qquad\le\frac{1}{\varepsilon^k} \E \Bigl( \sup_{s\in[n,n+1]} \sup _{x\in\Z^d}\bigl\llvert u_t(x)-u_s(x) \bigr\rrvert ^k \Bigr) = O \bigl( n^{-\alpha} \bigr)\qquad \mbox{as $n \to\infty$}. \end{eqnarray} We may use the Borel--Cantelli lemma and \eqref{eq:dissipate:subseq} in order to deduce that $\lim_{t\to\infty}\sup_{x\in\Z ^d}|u_t(x)|= 0$ a.s. Thanks to this fact, Corollary~\ref{co:iff} implies the seemingly stronger assertion that $\lim_{t\to\infty}\|u_t\|_{\ell^2(\Z^d)}^2= 0$ a.s., and completes the proof. \end{pf*} \begin{appendix}\label{app} \section*{Appendix: Some renewal theory} In this appendix we state and prove a few facts from (linear) renewal theory. These facts ought to be well known, but we have not succeeded to find concrete references, and so will describe them in some detail.\vadjust{\goodbreak} Let us suppose that the functions $h,g\dvtx(0 ,\infty)\to\R_+$ are locally integrable (say) and pre-defined, and let us look for a measurable solution $f\dvtx(0 ,\infty)\to\R_+$ to the renewal equation \setcounter{equation}{0} \begin{equation} \label{RE} f(t) = g(t) + \int_0^t h(t-s)f(s) \,\d s\qquad (t\ge0). \end{equation} If $h\in L^1(0 ,\infty)$, then this is a classical subject \cite{FellerVolII}. For a more general treatment, we may proceed with Picard's iteration: Let $f^{(0)}(t)\dvtx(0 ,\infty)\to\R_+$ be a fixed measurable function, and iteratively define \begin{equation} f^{(n+1)}(t) := g(t) + \int_0^t h(t-s)f^{(n)}(s) \,\d s\qquad (t> 0, n\ge0). \end{equation} \begin{lemmaa}\label{lem:RE1} Suppose that there exists a constant $\beta\in\R$ that satisfies the following three conditions: \textup{(i)} $\gamma:= \sup_{t\ge0}[\exp(-\beta t)g(t)]<\infty$; \textup{(ii)} $\rho:=\int_0^\infty\exp(-\beta t)h(t) \,\d t<1$; \textup{(iii)} $\sup_{t\ge0}[\exp(-\beta t)f^{(0)}(t)]<\infty$. Then \eqref{RE} has a unique nonnegative solution $f$ that satisfies the following: \begin{equation} \label{eq:RE:UB(f)} f(t) \le\frac{\gamma{\mathrm{e}}^{\beta t}}{1-\rho}\qquad (t\ge0). \end{equation} Moreover, $\lim_{n\to\infty} \sup_{t\ge0}( {\mathrm{e}}^{-\beta t}| f^{(n)}(t)-f(t)|)=0$. \end{lemmaa} \begin{pf} Choose such a $\beta\in\R$, and define \begin{equation} \gamma:=\sup_{t\ge0} \bigl[ {\mathrm{e}}^{-\beta t}g(t) \bigr],\qquad \rho:=\int_0^\infty{\mathrm{e}}^{-\beta t}h(t) \,\d t<1 \end{equation} and \begin{equation}\qquad C_k := \sup_{t\ge0} \bigl({\mathrm{e}}^{-\beta t} f^{(k)}(t) \bigr), \qquad D_k := \sup_{t\ge0} \bigl({\mathrm{e}}^{-\beta t}\bigl\llvert f^{(k)}(t) - f^{(k-1)}(t) \bigr\rrvert \bigr), \end{equation} for integers $k\ge1$. Thanks to the definition of the $f^{(k)}$'s, \begin{equation} C_{n+1} \le\gamma+ \rho C_n,\qquad D_{n+1} \le\rho D_n\qquad (n\ge0). \end{equation} Consequently, $\sup_{n\ge0}C_n \le\gamma(1-\rho)^{-1}$ and $D_n=O(\rho^n)$. Since $\sum_{n=0}^\infty D_n<\infty$, it follows that there exists a function $f$ such that $\sup_{t\ge0}({\mathrm{e}}^{-\beta t}|f^{(n)}(t)-f(t)|)\to0$ as $n\to \infty$, and $\sup_{t\ge0}({\mathrm{e}}^{-\beta t}f(t))\le\sup_{n\ge0}C_n$. These observations together prove the lemma. \end{pf} The following is the main result of this appendix. \begin{lemmaa}[(Comparison lemma)]\label{lem:RE:comparison} Suppose there exists $\beta\in\R$ such that: \textup{(i)}~$\gamma:= \sup_{t\ge0}[\exp(-\beta t)g(t)]<\infty$ and \textup{(ii)} $\rho:=\int_0^\infty\exp(-\beta t)h(t) \,\d t<1$; and let $f$ denote the unique nonnegative solution to \eqref{RE} that satisfies \eqref{eq:RE:UB(f)}. If $F\dvtx\R_+\to\R_+$ satisfies: \textup{(a)} $\sup_{t\ge0}[\exp(-\beta t)F(t)]<\infty$ and \textup{(b)} \begin{equation} \label{eq:comp1} F(t) \ge g(t) + \int_0^t h(t-s)F(s) \,\d s\qquad (t\ge0), \end{equation} then $f(t)\le F(t)$ for all $t\ge0$. Finally, if we replace condition \eqref{eq:comp1} by \begin{equation} \label{eq:comp2} F(t) \le g(t) + \int_0^t h(t-s)F(s) \,\d s \qquad (t\ge0), \end{equation} then $f(t)\ge F(t)$ for all $t\ge0$. \end{lemmaa} \begin{pf} We will prove \eqref{eq:comp1}; \eqref{eq:comp2} is proved similarly. We apply Picard's iteration with initial function $f^{(0)}:=F$ and note that \begin{equation} f^{(1)}(t) = g(t) + \int_0^t h(t-s) F(s) \,\d s \le F(t) \qquad (t\ge0). \end{equation} This and induction together show that $f^{(n+1)}(t)\le f^{(n)}(t)$ for all $t\ge0$ and $n\ge0$. Let $n\to\infty$ to deduce the lemma from Lemma~\ref{lem:RE1}. \end{pf} \end{appendix} \section*{Acknowledgments} We thank an anonymous referee and an Associate Editor whose remarks and suggestions have helped improve the presentation of this paper.
\section{Introduction} \label{sec:intro} Reversible cellular automata are deterministic, spatially extended, microscopically reversible dynamical systems. They provide a suitable framework --- an alternative to Hamiltonian dynamics --- to examine the dynamical foundations of statistical mechanics with simple caricature models. The intuitive structure of cellular automata makes them attractive to mathematicians, and their combinatorial nature makes them amenable to perfect simulations and computational study. Some reversible cellular automata have long been observed, in simulations, to exhibit ``thermodynamic behavior'': starting from a random configuration, they undergo a transient dynamics until they reach a state of macroscopic (statistical) equilibrium. Which of the equilibrium states the system is going to settle in could often be guessed on the basis of few statistics of the initial configuration. One such example is the \emph{Q2R} cellular automaton~\cite{Vic84}, which is a deterministic dynamics on top of the Ising model. Like the standard Ising model, a configuration of the Q2R model consists of an infinite array of symbols $\symb{+}$ (representing an upward magnetic \emph{spin}) and $\symb{-}$ (a downward spin) arranged on the two-dimensional square lattice. The symbols are updated in iterated succession of two updating stages: at the first stage, the symbols on the even sites (the black cells of the chess board) are updated, and at the second stage, the symbols on the odd sites. The updating of a symbol is performed according to a simple rule: a spin is flipped if and only if among the four neighboring spins, there are equal numbers of upward and downward spins. The dynamics is clearly \emph{reversible} (changing the order of the two stages, we could traverse backward in time). It also conserves the Ising energy (i.e., the number of pairs of adjacent spins that are anti-aligned). Few snapshots from a simulation are shown in Figure~\ref{fig:ising:simulation}. Starting with a random configuration in which the direction of each spin is determined by a biased coin flip, the Q2R cellular automaton evolves towards a state of apparent equilibrium that resembles a sample from the Ising model at the corresponding temperature.\footnote{% The Q2R model has no temperature parameter. A correspondence can however be made with the temperature at which the Ising model has the same expected energy density. } More sophisticated variants of the Q2R model show numerical agreement with the phase diagram of the Ising model, at least away from the critical point~\cite{Cre86}. See~\cite{TofMar87}, Chapter 17, for further simulations and an interesting discussion. \begin{figure} \begin{center} \begin{tabular}{ccc} \begin{minipage}[c]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{figures/p01-200x200-t0.png} \end{minipage} & \begin{minipage}[c]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{figures/p01-200x200-t1000.png} \end{minipage} & \begin{minipage}[c]{0.3\textwidth} \centering \includegraphics[width=\textwidth]{figures/p01-200x200-t10000.png} \end{minipage} \medskip\\ $t=0$ & $t=1000$ & $t=10000$ \end{tabular} \end{center} \caption{% Simulation of the Q2R cellular automaton on a $200\times 200$ toroidal lattice. Black represents an upward spin. The initial configuration is chosen according to a Bernoulli distribution with $1{\,:\,}9$ bias. The ``macroscopic appearance'' of the configuration does not vary significantly after time~$10000$. } \label{fig:ising:simulation} \end{figure} Wolfram was first to study cellular automata from the point of view of statistical mechanics~\cite{Wol83,Wol84} (see also~\cite{Wol02}). He made a detailed heuristic analysis of the so-called elementary cellular automata (those with two states per site and local rule depending on three neighboring sites in one dimension) using computer simulations. One of Wolfram's observations (the randomizing property of the XOR cellular automaton) was mathematically confirmed by Lind~\cite{Lin84}, although the same result had also been obtained independently by Miyamoto~\cite{Miy79}. Motivated by the problem of foundations of statistical mechanics, Takesue made a similar study of elementary \emph{reversible} cellular automata and investigated their ergodic properties and thermodynamic behavior~\cite{Tak87,Tak89,Tak90}. Recognizing the role of conservation laws in presence or absence of thermodynamic behavior, he also started a systematic study of additive conserved quantities in cellular automata~\cite{HatTak91,Tak95}. This article concerns the ``states of macroscopic equilibrium'' and their connection with conservation laws in a class of cellular automata including the reversible ones. As in statistical mechanics, we identify the ``macroscopic states'' of lattice configurations with probability measures on the space of all such configurations. The justification and proper interpretation of this formulation is beyond the scope of this article. We content ourselves with recalling two distinct points of view: the \emph{subjective} interpretation (probability measures are meant to describe the partial states of knowledge of an observer; see~\cite{Jay57}) and the \emph{frequentist} interpretation (a probability measure represents a class of configurations sharing the same statistics). See~\cite{Skl93} for comparison and discussion. If we call tail-measurable observables ``macroscopic'', a probability measure that is trivial on the tail events would give a full description of a macroscopic state (see Paragraph~(7.8) of~\cite{Geo88}). On the other hand, restricting ``macroscopic'' observables to statistical averages (i.e., averages of local observables over the lattice), one could identify the macroscopic states with probability measures that are shift-invariant and ergodic. The configurations in the ergodic set of a shift-ergodic probability measure (i.e., the generic points in its support; see~\cite{Oxt52}) may then be considered as ``typical'' microscopic states for the identified macroscopic state. The interpretation of ``equilibrium'' is another unsettling issue that we leave open. Equilibrium statistical mechanics postulates that the equilibrium states (of a lattice model described by interaction energies) are suitably described by Gibbs measures (associated with the interaction energies)~\cite{Rue04,Isr79,Geo88}. One justification (within the subjective interpretation) is the variational principle that characterizes the shift-invariant Gibbs measures as measures that maximize entropy under a fixed expected energy density constraint. Within a dynamical framework, on the other hand, the system is considered to be in macroscopic equilibrium if its internal fluctuations are not detected by macroscopic observables. One is therefore tempted to identify the equilibrium states of a cellular automaton with (tail-trivial or shift-ergodic) probability measures that are time-invariant. Unfortunately, there are usually an infinity of invariant measures that do not seem to be of physical relevance. For instance, in any cellular automaton, the uniform distribution on the shift and time orbit of a jointly periodic configuration is time-invariant and shift-ergodic, but may hardly be considered a macroscopic equilibrium state. Other conditions such as ``smoothness'' or ``attractiveness'' therefore might be needed. Rather than reversible cellular automata (i.e., those whose trajectories can be traced backward by another cellular automaton), we work with the broader class of surjective cellular automata (i.e., those that act surjectively on the configuration space). Every reversible cellular automaton is surjective, but there are many surjective cellular automata that are not reversible. Surjective cellular automata are nevertheless ``almost injective'' in that the average amount of information per site they erase in each time step is vanishing. They are precisely those cellular automata that preserve the uniform Bernoulli measure (cf.\ Liouville's theorem for Hamiltonian systems). Even if not necessarily physically relevant, they provide a richer source of interesting examples, which could be used in case studies. For instance, most of the known examples of the randomization phenomenon (which, we shall argue, could provide an explanation of approach to equilibrium) are in non-reversible surjective cellular automata. The invariance of Gibbs measures under surjective cellular automata turns out to be associated with their conservation laws. More precisely, if an additive energy-like quantity, formalized by a Hamiltonian, is conserved by a surjective cellular automaton, the cellular automaton maps the simplex of shift-invariant Gibbs measures corresponding to that Hamiltonian onto itself (Theorem~\ref{thm:conservation-invariance:1}). The converse is true in a stronger sense: if a surjective cellular automaton maps a (not necessarily shift-invariant) Gibbs measure for a Hamiltonian to a Gibbs measure for the same Hamiltonian, the Hamiltonian must be conserved by the cellular automaton (Corollary~\ref{cor:conservation-invariance:1:non-shift-invariant}). The proof of this correspondence is an immediate consequence of the variational characterization of shift-invariant Gibbs measures and the fact that surjective cellular automata preserve the average entropy per site of shift-invariant probability measures (Theorem~\ref{thm:pentropy:pre-injective}). An elementary proof of a special case was presented earlier~\cite{KarTaa11}. Note that if a conserved Hamiltonian has a unique Gibbs measure, then that unique Gibbs measure will be invariant under the cellular automaton. This is the case, for example, in one dimension, or when the Hamiltonian does not involve the interaction of more than one site (the Bernoulli case). An important special case is the \emph{trivial} Hamiltonian (all configurations on the same ``energy'' level) which is obviously conserved by every surjective cellular automaton. The uniform Bernoulli measure is the unique Gibbs measure for the trivial Hamiltonian, and we recover the well-known fact that every surjective cellular automaton preserves the uniform Bernoulli measure on its configuration space (i.e., Corollary~\ref{cor:balance:weak}). If, on the other hand, the simplex of shift-invariant Gibbs measures for a conserved Hamiltonian has more than one element, the cellular automaton does not need to preserve individual Gibbs measures in this simplex (Example~\ref{exp:ising:ca:invariant-non-invariant}). We do not know whether, in general, a surjective cellular automaton maps the non-shift-invariant Gibbs measures for a conserved Hamiltonian to Gibbs measures for the same Hamiltonian, but this is known to be the case for a proper subclass of surjective cellular automata including the reversible ones (Theorem~\ref{thm:gibbs:factor-map:complete}), following a result of Ruelle. The essence of the above-mentioned connection between conservation laws and invariant Gibbs measures comes about in a more abstract setting, concerning the pre-injective factor maps between strongly irreducible shifts of finite type. We show that a shift-invariant pre-image of a (shift-invariant) Gibbs measure under such a factor map is again a Gibbs measure (Corollary~\ref{thm:equilibrium:factor-map}). We find a simple sufficient condition under which a pre-injective factor map transforms a shift-invariant Gibbs measure into a measure that is not Gibbs (Proposition~\ref{prop:gibbs-non-gibbs}). An example of a surjective cellular automaton is given that eventually transforms every starting Gibbs measure into a non-Gibbs measure (Example~\ref{exp:xor:non-gibbs:contd}). As an application in the study of phase transitions in equilibrium statistical mechanics, we demonstrate how the result of Aizenman and Higuchi regarding the structure of the simplex of Gibbs measures for the two-dimensional Ising model could be more transparently formulated using a pre-injective factor map (Example~\ref{exp:ising-contour}). The correspondence between invariant Gibbs measures and conservation laws allows us to reduce the problem of invariance of Gibbs measures to the problem of conservation of additive quantities. Conservation laws in cellular automata have been studied by many from various points of view (see e.g.~\cite{Pom84,HatTak91,BocFuk98,Piv02, DurForRok03,ForGra03,MorBocGol04,Ber07,Gar11,ForKarTaa11}). For example, simple algorithms have been proposed to find all additive quantities of up to a given interaction range that are conserved by a cellular automaton. Such an algorithm can be readily applied to find all the full-support Markov measures that are invariant under a surjective cellular automaton (at least in one dimension). We postpone the study of this and similar algorithmic problems to a separate occasion. A highlight of this article is the use of this correspondence to obtain severe restrictions on the existence of invariant Gibbs measures in two interesting classes of cellular automata with strong chaotic behavior. First, we show that a strongly transitive cellular automaton cannot have any invariant Gibbs measure other than the uniform Bernoulli measure (Corollary~\ref{cor:ca:strongly-transitive:rigidity}). The other result concerns the class of one-dimensional reversible cellular automata that are obtained by swapping the role of time and space in positively expansive cellular automata. For such reversible cellular automata, we show that the uniform Bernoulli measure is the unique invariant Markov measure with full support (Corollary~\ref{cor:ca:pexp-transpose:rigidity}). Back to the interpretation of shift-ergodic probability measures as macroscopic states, one might interpret the latter results as an indication of ``absence of phase transitions'' in the cellular automata in question. Much sharper results have been obtained by others for narrower classes of cellular automata having algebraic structures (see the references in Example~\ref{exp:xor:ca}). A mathematical description of approach to equilibrium (as observed in the Q2R example) seems to be very difficult in general. The randomization property of algebraic cellular automata (the result of Miyamoto and Lind and its extensions; see Example~\ref{exp:xor:randomization}) however provides a partial explanation of approach to equilibrium in such cellular automata. Finding ``physically interesting'' cellular automata with similar randomization property is an outstanding open problem. The structure of the paper is as follows. Section~\ref{sec:background} is dedicated to the development of the setting and background material. Given the interdisciplinary nature of the subject, we try to be as self-contained as possible. Basic results regarding the pre-injective factor maps between shifts of finite type as well as two degressing applications appear in Section~\ref{sec:entropy-preserving}. In Section~\ref{sec:ca}, we apply the results of the previous section on cellular automata. Conservation laws in cellular automata are discussed in Section~\ref{sec:ca:conservation-laws}. Proving the absence of non-trivial conservation laws in two classes of chaotic cellular automata in Section~\ref{sec:ca:rigidity}, we obtain results regarding the rigidity of invariant measures for these two classes. Section~\ref{sec:ca:randomization} contains a discussion of the problem of approach to equilibrium. \section{Background} \label{sec:background} \subsection{Observables, Probabilities, and Dynamical Systems} Let $\xSp{X}$ be a compact metric space. By an \emph{observable} we mean a Borel measurable function $f:\xSp{X}\to\xR$. The set of continuous observables on $\xSp{X}$ will be denoted by $C(\xSp{X})$. This is a Banach space with the uniform norm. The default topology on $C(\xSp{X})$ is the topology of the uniform norm. The set of Borel probability measures on $\xSp{X}$ will be denoted by $\xSx{P}(\xSp{X})$. The expectation operator of a Borel probability measure $\pi\in\xSx{P}(\xSp{X})$ is a positive linear (and hence continuous) functional on $C(\xSp{X})$. Conversely, the Riesz representation theorem states that every normalized positive linear functional on $C(\xSp{X})$ is the expectation operator of a unique probability measure on $\xSp{X}$. Therefore, the Borel probability measures can equivalently be identified as normalized positive linear functionals on $C(\xSp{X})$. We assume that $\xSx{P}(\xSp{X})$ is topologized with the weak topology. This is the weakest topology with respect to which, for every observable $f\in C(\xSp{X})$, the mapping $\pi\mapsto \pi(f)$ is continuous. The space $\xSx{P}(\xSp{X})$ under the weak topology is compact and metrizable. If $\delta_x$ denotes the Dirac measure concentrated at $x\in\xSp{X}$, the map $x\mapsto\delta_x$ is an embedding of $\xSp{X}$ into $\xSx{P}(\xSp{X})$. The Dirac measures are precisely the extreme elements of the convex set $\xSx{P}(\xSp{X})$, and by the Krein-Milman theorem, $\xSx{P}(\xSp{X})$ is the closed convex hull of the Dirac measures. Let $\xSp{X}$ and $\xSp{Y}$ be compact metric spaces and $\Phi:\xSp{X}\to\xSp{Y}$ a continuous mapping. We denote the induced mapping $\xSx{P}(\xSp{X})\to\xSx{P}(\xSp{Y})$ by the same symbol $\Phi$; hence $(\Phi\pi)(E)\IsDef\pi(\Phi^{-1}E)$. The dual map $C(\xSp{Y})\to C(\xSp{X})$ is denoted by $\Phi^*$; that is, $(\Phi^*f)(x)\IsDef f(\Phi x)$. The following lemma is well-known. \begin{lemma} \label{lem:onto-one-to-one} Let $\xSp{X}$ and $\xSp{Y}$ be compact metric spaces and $\Phi:\xSp{X}\to\xSp{Y}$ a continuous map. \begin{enumerate}[ \rm a)] \item $\Phi:\xSp{X}\to\xSp{Y}$ is one-to-one if and only if $\Phi^*:C(\xSp{Y})\to C(\xSp{X})$ is onto, which in turn holds if and only if $\Phi:\xSx{P}(\xSp{X})\to\xSx{P}(\xSp{Y})$ is one-to-one. \item $\Phi:\xSp{X}\to\xSp{Y}$ is onto if and only if $\Phi^*:C(\xSp{Y})\to C(\xSp{X})$ is one-to-one, which in turn holds if and only if $\Phi:\xSx{P}(\xSp{X})\to\xSx{P}(\xSp{Y})$ is onto. \end{enumerate} \end{lemma} \begin{proof}\ \\ \begin{enumerate}[ a)] \item Suppose that $\Phi:\xSp{X}\to\xSp{Y}$ is one-to-one. Let $g\in C(\xSp{X})$ be arbitrary. Define $\xSp{Y}_0\IsDef\Phi\xSp{X}$. Then $g\xO\Phi^{-1}:\xSp{Y}_0\to\xR$ is a continuous real-valued function on a closed subset of a compact metric space. Hence, by the Tietze extension theorem, it has an extension $f:\xSp{Y}\to\xR$. We have $f\xO\Phi=g\xO\Phi^{-1}\xO\Phi=g$. Therefore, $\Phi^*:C(\xSp{Y})\to C(\xSp{X})$ is onto. The other implications are trivial. \item Suppose that $\Phi:\xSp{X}\to\xSp{Y}$ is not onto. Let $\xSp{Y}_0\IsDef\Phi\xSp{X}$ and pick an arbitrary $y\in\xSp{Y}\setminus\xSp{Y}_0$. Using the Tietze extension theorem, we can find $f,f'\in C(\xSp{Y})$ such that $\xRest{f}{\xSp{Y}_0}=\xRest{f'}{\xSp{Y}_0}$ but $f(y)\neq f'(y)$. Then $f\xO\Phi=f'\xO\Phi$. Hence, $\Phi^*:C(\xSp{Y})\to C(\xSp{Y})$ is not one-to-one. Next, suppose that $\Phi:\xSp{X}\to\xSp{Y}$ is onto. By the Krein-Milman theorem, the set $\xSx{P}(\xSp{Y})$ is the closed convex hull of Dirac measures on $\xSp{Y}$. Let $\pi=\sum_i \lambda_i\delta_{y_i}\in\xSx{P}(\xSp{Y})$ be a convex combination of Dirac measures. Pick $x_i\in\xSp{X}$ such that $\Phi x_i=y_i$, and define $\nu\IsDef\sum_i\lambda_i\delta_{x_i}\in\xSx{P}(\xSp{X})$. Then $\Phi\nu=\pi$. Therefore, $\Phi\xSx{P}(\xSp{X})$ is dense in $\xSx{P}(\xSp{Y})$. Since $\Phi\xSx{P}(\xSp{X})$ is also closed, we obtain that $\Phi\xSx{P}(\xSp{X})=\xSx{P}(\xSp{Y})$. That is, $\Phi:\xSx{P}(\xSp{X})\to\xSx{P}(\xSp{Y})$ is onto. The remaining implication is trivial. \end{enumerate} \end{proof} By a \emph{dynamical system} we shall mean a compact metric space $\xSp{X}$ together with a continuous action $(i,x)\mapsto \varphi^i x$ of a discrete commutative finitely generated group or semigroup $\xL$ on $\xSp{X}$. In case of a cellular automaton, $\xL$ is the set of non-negative integers $\xN$ (or the set of integers $\xZ$ if the cellular automaton is reversible). For a $d$-dimensional shift, $\xL$ is the $d$-dimensional hyper-cubic lattice~$\xZ^d$. Every dynamical system $(\xSp{X},\varphi)$ has at least one invariant measure, that is, a probability measure $\pi\in\xSx{P}(\xSp{X})$ such that $\varphi^i\pi=\pi$ for every $i\in\xL$. In fact, every non-empty, closed and convex subset of $\xSx{P}(\xSp{X})$ that is closed under the application of $\varphi$ contains an invariant measure. We will denote the set of invariant measures of $(\xSp{X},\varphi)$ by $\xSx{P}(\xSp{X},\varphi)$. We also define $C(\xSp{X},\varphi)$ as the closed linear subspace of $C(\xSp{X})$ generated by the observables of the form $g\xO \varphi^i - g$ for $g\in C(\xSp{X})$ and $i\in\xL$; that is, \begin{align} C(\xSp{X},\varphi) &\IsDef \overline{ \langle g\xO\varphi^i - g: g\in C(\xSp{X}) \text{ and } i\in\xL \rangle } \; \end{align} (see~\cite{Rue04}, Sections~4.7--4.8). Then $\xSx{P}(\xSp{X},\varphi)$ and $C(\xSp{X},\varphi)$ are annihilators of each other: \begin{lemma}[see e.g.~\cite{KatRob01}, Proposition~2.13] \label{lem:annihilators} Let $(\xSp{X},\varphi)$ be a dynamical system. Then, \begin{enumerate}[ \rm a)] \item $\xSx{P}(\xSp{X},\varphi) = \left\{ \pi\in\xSx{P}(\xSp{X}): \text{$\pi(f)=0$ for every $f\in C(\xSp{X},\varphi)$} \right\}$. \item $C(\xSp{X},\varphi) = \left\{ f\in C(\xSp{X}): \text{$\pi(f)=0$ for every $\pi\in\xSx{P}(\xSp{X},\varphi)$} \right\}$. \end{enumerate} \end{lemma} \begin{proof}\ \\ \begin{enumerate}[ a)] \item A probability measure $\pi$ on $\xSp{X}$ is in $\xSx{P}(\xSp{X},\varphi)$ if and only if $\pi(g\xO\varphi^i - g)=0$ for every $g\in C(\xSp{X})$ and $i\in\xL$. Furthermore, for each $\pi$, the set $\{f\in C(\xSp{X}): \pi(f)=0\}$ is a closed linear subspace of $C(\xSp{X})$. Therefore, the equality in~(a) holds. \item Let us denote the righthand side of the claimed equality in~(b) by $D$. The set $D$ is closed and linear, and contains all the elements of the form $g\xO\varphi^i-g$ for all $g\in C(\xSp{X})$ and $i\in\xL$. Therefore, $C(\xSp{X},\varphi)\subseteq D$. Conversely, let $f\in C(\xSp{X})\setminus C(\xSp{X},\varphi)$. Then every element $h\in\langle f, C(\xSp{X},\varphi)\rangle$ has a unique representation $h=a_h f + u_h$ where $a_h\in\xR$ and $u_h\in C(\xSp{X},\varphi)$. Define the linear functional $J:\langle f, C(\xSp{X},\varphi)\rangle\to\xR$ by $J(h)=J(a_h f + u_h)\IsDef a_h$. Then $J$ is bounded ($\norm{h}=\norm{a_h f + u_h}\geq a_h\delta=\abs{J(h)}\delta$, where $\delta>0$ is the distance between $f$ and $C(\xSp{X},\varphi)$), and hence, by the Hahn-Banach theorem, has a bounded linear extension $\widehat{J}$ on $C(\xSp{X})$. According to the Riesz representation theorem, there is a unique signed measure $\pi$ on $\xSp{X}$ such that $\pi(h)=\widehat{J}(h)$ for every $h\in C(\xSp{X})$. Let $\pi=\pi^+-\pi^-$ be the Hahn decomposition of $\pi$. Since $\pi(f)=1$, either $\pi^+(f)>0$ or $\pi^-(f)>0$. If $\pi^+(f)>0$, define $\pi^\ast\IsDef \frac{1}{\pi^+(\xSp{X})}\pi^+$; otherwise $\pi^\ast\IsDef \frac{1}{\pi^-(\xSp{X})}\pi^-$. Then $\pi^*$ is a probability measure with $\pi^*(u)=0$ for every $u\in C(\xSp{X},\varphi)$, which according to part~(a), ensures that $\pi^*\in\xSx{P}(\xSp{X},\varphi)$. On the other hand $\pi^*(f)>0$, and hence $f\notin D$. We conclude that $C(\xSp{X},\varphi)=D$. \end{enumerate} \end{proof} If $K(\xSp{X})$ is a dense subspace of $C(\xSp{X})$, the subspace $C(\xSp{X},\varphi)$ can also be expressed in terms of $K(\xSp{X})$. Namely, if we define \begin{align} \label{eq:def:local-coboundaries} K(\xSp{X},\varphi) &\IsDef \langle h\xO\varphi^i - h: h\in K(\xSp{X}) \text{ and } i\in\xL \rangle \; , \end{align} then $C(\xSp{X},\varphi)=\overline{K(\xSp{X},\varphi)}$. If $D_0$ is a finite generating set for the group/semigroup $\xL$, the subspace $K(\xSp{X},\varphi)$ may also be expressed as \begin{align} \Big\{ \sum_{i\in D_0} (h_i\xO\varphi^i - h_i) : h_i\in K(\xSp{X}) \Big\} \;. \end{align} In particular, if $\xL=\xZ$ or $\xL=\xN$, then every element of $K(\xSp{X},\varphi)$ is of the form $h\xO\varphi-h$ for some $h\in K(\xSp{X})$. A morphism between two dynamical systems $(\xSp{X},\varphi)$ and $(\xSp{Y},\psi)$ is a continuous map $\Theta:\xSp{X}\to\xSp{Y}$ such that $\Theta\varphi=\psi\Theta$. An epimorphism (i.e., an onto morphism) is also called a \emph{factor map}. If $\Theta:\xSp{X}\to\xSp{Y}$ is a factor map, then $(\xSp{Y},\psi)$ is a \emph{factor} of $(\xSp{X},\varphi)$, and $(\xSp{X},\varphi)$ is an \emph{extension} of $(\xSp{Y},\psi)$. A monomorphism (i.e., a one-to-one morphism) is also known as an \emph{embedding}. If $(\xSp{Y},\psi)$ is embedded in $(\xSp{X},\varphi)$ by the inclusion map, $(\xSp{Y},\psi)$ is called a \emph{subsystem} of $(\xSp{X},\varphi)$. A \emph{conjugacy} between dynamical systems is the same as an isomorphism; two systems are said to be conjugate if they are isomorphic. \subsection{Shifts and Cellular Automata} \label{sec:shifts-ca} A cellular automaton is a dynamical system on symbolic configurations on a lattice. The configuration space itself has translational symmetry and can be considered as a dynamical system with the shift action. We allow constraints on the local arrangement of symbols to include models with so-called hard interactions, such as the hard-core model (Example~\ref{exp:hard-core}) or the contour model (Example~\ref{exp:contour}). Such a restricted configuration space is modeled by a (strongly irreducible) shift of finite type. The \emph{sites} of the $d$-dimensional (hypercubic) \emph{lattice} are indexed by the elements of the group~$\xL\IsDef\xZ^d$. A \emph{neighborhood} is a non-empty finite set $N\subseteq\xL$ and specifies a notion of closeness between the lattice sites. The \emph{$N$-neighborhood} of a site $a\in\xL$ is the set $N(a)\IsDef a+N = \{a+i: i\in N\}$. Likewise, the $N$-neighborhood of a set $A\subseteq\xL$ is $N(A)\IsDef A+N \IsDef \{a+i: a\in A \text{ and } i\in N\}$. The (symmetric) \emph{$N$-boundary} of a set $A\subseteq\xL$ is $\partial N(A)\IsDef N(A)\cap N(\xL\setminus A)$. For a set $A\subseteq\xL$, we denote by $N^{-1}(A)\IsDef A-N$ the set of all sites $b\in\xL$ that have an $N$-neighbor in $A$. A \emph{configuration} is an assignment $x:\xL\to S$ of symbols from a finite set $S$ to the lattice sites. The symbol $x(i)$ assigned to a site $i\in\xL$ is also called the \emph{state} of site $i$ in $x$. For two configurations $x,y:\xL\to S$, we denote by $\diff(x,y)\IsDef\{i\in\xL: x(i)\neq y(i)\}$ the set of sites on which $x$ and $y$ disagree. Two configurations $x$ and $y$ are said to be \emph{asymptotic} (or tail-equivalent) if $\diff(x,y)$ is finite. If $D\subseteq\xL$ is finite, an assignment $p:D\to S$ is called a \emph{pattern} on $D$. If $p:D\to S$ and $q:E\to S$ are two patterns (or partial configurations) that agree on $D\cap E$, we denote by $p\lor q$ the pattern (or partial configuration) that agrees with $p$ on $D$ and with $q$ on $E$. Let $S$ be a finite set of symbols with at least two elements. The set $S^\xL$ of all configurations of symbols from $S$ on $\xL$ is given the product topology, which is compact and metrizable. The convergence in this topology is equivalent to site-wise eventual agreement. If $D\subseteq\xL$ is a finite set and $x$ a configuration (or a partial configuration whose domain includes $D$), the set \begin{align} [x]_D &\IsDef \{y \in S^\xL: \xRest{y}{D}=\xRest{x}{D}\} \end{align} is called a \emph{cylinder} with \emph{base} $D$. If $p:D\to S$ is a pattern, we may write more concisely $[p]$ rather than $[p]_D$. In one dimension (i.e., if $\xL=\xZ$), we may also use words to specify cylinder sets: if $u=u_0u_1\cdots u_{n-1}\in S^*$ is a word over the alphabet $S$ and $k\in\xZ$, we write $[u]_k$ for the set of configurations $x\in S^\xZ$ such that $x_{k+i}=u_i$ for each $0\leq i<n$. The cylinders are clopen (i.e., both open and close) and form a basis for the product topology. The Borel $\sigma$-algebra on $S^\xL$ is denoted by $\family{F}$. For $A\subseteq\xL$, the sub-$\sigma$-algebra of events occurring in $A$ (i.e., the $\sigma$-algebra generated by the cylinders whose base is a subset of $A$) will be denoted by $\family{F}_A$. Given a configuration $x:\xL\to S$ and an element $k\in\xL$, we denote by $\sigma^k x$ the configuration obtained by \emph{shifting} (or \emph{translating}) $x$ by vector $k$; that is, $(\sigma^k x)(i)\IsDef x(k+i)$ for every $i\in\xL$. The dynamical system defined by the action of the shift $\sigma$ on $S^\xL$ is called the \emph{full shift}. A closed shift-invariant set $\xSp{X}\subseteq S^\xL$ is called a \emph{shift space} and the subsystem of $(S^\xL,\sigma)$ obtained by restricting $\sigma$ to $\xSp{X}$ is called a \emph{shift} system. We shall use the same symbol $\sigma$ for the shift action of all shift systems. This will not lead to confusion, as the domain will always be clear from the context. A shift space $\xSp{X}\subseteq S^\xL$ is uniquely determined by its \emph{forbidden} patterns, that is, the patterns $p:D\to S$ such that $[p]_D\cap\xSp{X}=\varnothing$. Conversely, every set $F$ of patterns defines a shift space by forbidding the occurrence of the elements of $F$; that is, \begin{align} \xSp{X}_{F} &\IsDef S^\xL \setminus \big(\bigcup_{i\in\xL}\bigcup_{p\in\xSp{F}} \sigma^{-i}[p]\,\big) \;. \end{align} The set of patterns $p:D\to S$ that are allowed in $\xSp{X}$ (i.e., $[p]\cap\xSp{X}\neq\varnothing$) is denoted by $L(\xSp{X})$. If $D\subseteq\xL$ is finite, we denote by $L_D(\xSp{X})\IsDef L(\xSp{X})\cap S^D$ the set of patterns on $D$ that are allowed in $\xSp{X}$. For every pattern $p\in L_D(\xSp{X})$, there is a configuration $x\in\xSp{X}$ such that $p\lor \xRest{x}{\xL\setminus D}\in\xSp{X}$. Given a finite set $D\subseteq\xL$ and a configuration $x\in\xSp{X}$, we write $L_D(\xSp{X}\,|\,x)$ the set of patterns $p\in L_D(\xSp{X})$ such that $p\lor \xRest{x}{\xL\setminus D}\in\xSp{X}$. A shift $(\xSp{X},\sigma)$ (or a shift space $\xSp{X}$) is \emph{of finite type}, if $\xSp{X}$ can be identified by forbidding a finite set of patterns, that is, $\xSp{X}=\xSp{X}_F$ for a finite set $F$. The shifts of finite type have the following gluing property: for every shift of finite type $(\xSp{X},\sigma)$, there is a neighborhood $0\in M\subseteq\xL$ such that for every two sets $A,B\subseteq\xL$ with $M(A)\cap M(B)=\varnothing$ and every two configurations $x,y\in\xSp{X}$ that agree outside $A\cup B$, there is another configuration $z\in\xSp{X}$ that agrees with~$x$ outside~$B$ and with~$y$ outside~$A$. A similar gluing property is the strong irreducibility: a shift $(\xSp{X},\sigma)$ is \emph{strongly irreducible} if there is a neighborhood $0\in M\subseteq\xL$ such that for every two sets $A,B\subseteq\xL$ with $M(A)\cap M(B)=\varnothing$ and every two configurations $x,y\in\xSp{X}$, there is another configuration $z\in\xSp{X}$ that agrees with~$x$ in $A$ and with~$y$ in~$B$. Note that strong irreducibility is a stronger version of topological mixing. A dynamical system $(\xSp{X},\varphi)$ is (topologically) \emph{mixing} if for every two non-empty open sets $U,V\subseteq\xSp{X}$, $U\cap \varphi^{-t}V\neq\varnothing$ for all but finitely many $t$. A one-dimensional shift of finite type is strongly irreducible if and only if it is mixing. Our primary interest in this article will be the shifts of finite type that are strongly irreducible, for these are sufficiently broad to encompass the configuration space of most physically interesting lattice models. The morphisms between shift systems are the same as the sliding block maps. A map $\Theta:\xSp{X}\to\xSp{Y}$ between two shift spaces $\xSp{X}\subseteq S^\xL$ and $\xSp{Y}\subseteq T^\xL$ is a \emph{sliding block map} if there is a neighborhood $0\in N\subseteq\xL$ (a \emph{neighborhood} for $\Theta$) and a function $\theta:L_N(\xSp{X})\to T$ (a \emph{local rule} for $\Theta$) such that \begin{align} (\Theta x)(i) &\IsDef \theta\big(\!\xRest{(\sigma^i x)}{N}\!\big) \end{align} for every configuration $x\in\xSp{X}$ and every site $i\in\xL$. Any sliding block map is continuous and commutes with the shift, and hence, is a morphism. Conversely, every morphism between shift systems is a sliding block map. Finite type property and strong irreducibility are both conjugacy invariants. A morphism $\Theta:\xSp{X}\to\xSp{Y}$ between two shifts $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$ is said to be \emph{pre-injective} if for every two distinct asymptotic configuration $x,y\in\xSp{X}$, the configurations $\Theta x$ and $\Theta y$ are distinct. A \emph{cellular automaton} on a shift space $\xSp{X}$ is a dynamical system identified by an endomorphism $\Phi:\xSp{X}\to\xSp{X}$ of $(\xSp{X},\sigma)$. The evolution of a cellular automaton starting from a configuration $x\in\xSp{X}$ is seen as synchronous updating of the state of different sites in $x$ using the local rule of $\Phi$. A cellular automaton $(\xSp{X},\Phi)$ is said to be \emph{surjective} (resp., \emph{injective}, \emph{pre-injective}, \emph{bijective}) if $\Phi$ is surjective (resp., injective, pre-injective, bijective). If $\Phi$ is bijective, the cellular automaton is further said to be \emph{reversible}, for $(\xSp{X},\Phi^{-1})$ is also a cellular automaton. In this article, we only work with cellular automata that are defined over strongly irreducible shifts of finite type. It is well-known that for cellular automata over strongly irreducible shifts of finite type, surjectivity and pre-injectivity are equivalent (the \emph{Garden-of-Eden} theorem; see below). In particular, every injective cellular automaton is also surjective, and hence reversible. Let $\xSp{X}\subseteq S^\xL$ be a shift space. A linear combination of characteristic functions of cylinder sets is called a \emph{local} observable. An observable $f:\xSp{X}\to\xR$ is local if and only if it is $\family{F}_D$-measurable for a finite set $D\subseteq\xL$. A finite set $D$ with such property is a \emph{base} for $f$; the value of $f$ at a configuration $x$ can be evaluated by looking at $x$ ``through the window $D$''. The set of all local observables on $\xSp{X}$, denoted by $K(\xSp{X})$, is dense in $C(\xSp{X})$. The set of all local observables on $\xSp{X}$ with base $D$ is denoted by $K_D(\xSp{X})$. Let $D\subseteq\xL$ be a non-empty finite set. The \emph{$D$-block presentation} of a configuration $x:\xL\to S$ is a configuration $x^{[D]}:\xL\to S^D$, where $x^{[D]}(i)\IsDef \xRest{x}{i+D}$. If $\xSp{X}$ is a shift space, the set of $D$-block presentations of the elements of $\xSp{X}$ is called the $D$-block presentation of $\xSp{X}$, and is denoted by $\xSp{X}^{[D]}$. The shifts $(\xSp{X},\sigma)$ and $(\xSp{X}^{[D]},\sigma)$ are conjugate via the map $x\mapsto x^{[D]}$. More background on shifts and cellular automata (from the view point of dynamical systems) can be found in the books~\cite{LinMar95,Kit98,Kur03b}. \subsection{Hamiltonians and Gibbs Measures} \label{sec:hamiltonian-gibbs} We will use the term Hamiltonian in more or less the same sense as in the Ising model or other lattice models from statistical mechanics, except that we do not require it to be interpreted as ``energy''. A Hamiltonian formalizes the concept of a local and additive quantity, be it energy, momentum or a quantity with no familiar physical interpretation. Let $\xSp{X}$ be an arbitrary set. A \emph{potential difference} on $\xSp{X}$ is a partial mapping $\Delta:\xSp{X}\times\xSp{X}\to\xR$ such that \begin{enumerate}[ a)] \item $\Delta(x,x)=0$ for every $x\in\xSp{X}$, \item $\Delta(y,x)=-\Delta(x,y)$ whenever $\Delta(x,y)$ exists, and \item $\Delta(x,z)=\Delta(x,y)+\Delta(y,z)$ whenever $\Delta(x,y)$ and $\Delta(y,z)$ both exist. \end{enumerate} Let $\xSp{X}\subseteq S^\xL$ be a shift space. A potential difference $\Delta$ on $\xSp{X}$ is a (relative) \emph{Hamiltonian} if \begin{enumerate}[ a)] \setcounter{enumi}{3} \item $\Delta(x,y)$ exists precisely when $x$ and $y$ are asymptotic, \item $\Delta(\sigma^a x,\sigma^a y)=\Delta(x,y)$ whenever $\Delta(x,y)$ exists and $a\in\xL$, and \item For every finite $D\subseteq\xL$, $\Delta$ is continuous when restricted to pairs $(x,y)$ with $\diff(x,y)\subseteq D$. \label{item:hamiltonian:continuous} \end{enumerate} Note that due to the compactness of $\xSp{X}$, the latter continuity is uniform among all pairs $(x,y)$ with $\diff(x,y)\subseteq D$. If the condition~(\ref{item:hamiltonian:continuous}) is strengthened by the following condition, we say that $\Delta$ is a \emph{finite-range} Hamiltonian. \begin{enumerate}[ a$\!\>'$)] \setcounter{enumi}{5} \item \label{item:hamiltonian:range}% There exists a neighborhood $0\in M\subseteq\xL$ (the \emph{interaction} neighborhood of $\Delta$) such that $\Delta(x,y)$ depends only on the restriction of~$x$ and~$y$ to $M(\diff(x,y))$. \end{enumerate} Hamiltonians in statistical mechanics are usually constructed by assigning interaction energies to different local arrangements of site states. Equivalently, they can be constructed using observables. A local observable $f\in K(\xSp{X})$ defines a finite-range Hamiltonian $\Delta_f$ on $\xSp{X}$ via \begin{equation} \label{eq:hamiltonian:observable} \Delta_f(x,y)\IsDef \sum_{i\in\xL}\left[ f(\sigma^i y) - f(\sigma^i x) \right] \end{equation} for every two asymptotic configurations $x,y\in\xSp{X}$. The value of $f\xO\sigma^i$ is then interpreted as the contribution of site $i$ to the energy-like quantity formalized by $\Delta_f$. The same construction works for non-local observables that are ``sufficiently short-ranged'' (i.e., whose dependence on faraway sites decays rapidly). The \emph{variation} of an observable $f:\xSp{X}\to\xR$ relative to a finite set $A\subseteq\xL$ is defined as \begin{align} \xvar_A(f) &= \sup_{\substack{x,y\in\xSp{X}\\ \text{$x=y$ on $A$}}}\abs{f(y)-f(x)} \;, \end{align} where the supremum is taken over all pairs of configurations $x$ and $y$ in $\xSp{X}$ that agree on $A$. A continuous observable $f$ is said to have \emph{summable variations} if \begin{align} \sum_{n=0}^\infty \abs{\partial I_n}\xvar_{I_n}(f) &< \infty \;, \end{align} where $I_n\IsDef[-n,n]^d$ and $\partial I_n\IsDef I_{n+1}\setminus I_n$. Every observable $f$ that has summable variations defines a Hamiltonian via~(\ref{eq:hamiltonian:observable}), in which the sum is absolutely convergent. We denote the set of observables with summable variations with $SV(\xSp{X})$. Note that $K(\xSp{X})\subseteq SV(\xSp{X})\subseteq C(\xSp{X})$. \begin{question} Is every Hamiltonian on a strongly irreducible shift of finite type generated by an observable with summable variations via~(\ref{eq:hamiltonian:observable})? Is every finite-range Hamiltonian on a strongly irreducible shift of finite type generated by a local observable? \end{question} \begin{proposition} \label{prop:Hamiltonian:observable:local} Every finite-range Hamiltonian on a full shift is generated by a local observable. \end{proposition} \begin{proof} The idea is to write the Hamiltonian as a telescopic sum (see e.g.~\cite{HatTak91}, or~\cite{Kar05}, Section~5). Let $\Delta$ be a finite-range Hamiltonian with interaction range $M$. Let $\xQuiet$ be an arbitrary uniform configuration. Let $\preceq$ be the lexicographic order on $\xL=\xZ^d$, and denote by $\successor(k)$, the successor of site $k\in\xL$ in this ordering. For every configuration $z$ that is asymptotic to $\xQuiet$, we can write \begin{align} \Delta(\xQuiet,z) &= \sum_{k\in\xL} \Delta(z_k,z_{\successor(k)}) \;, \end{align} where $z_k$ is the configuration that agrees with $z$ on every site $i\prec k$ and with $\xQuiet$ on every site $i\succeq k$. Note that all but a finite number of terms in the above sum are~$0$. For every configuration $z$, we define $f(z)\IsDef \Delta(z_0,z_{\successor(0)})$ with the same definition for $z_k$ as above. This is clearly a local observable with base $M$. If $z$ is asymptotic to $\xQuiet$, the above telescopic expansion shows that $\Delta(\xQuiet,z)=\Delta_f(\xQuiet,z)$. If $x$ and $y$ are arbitrary asymptotic configurations, we have $\Delta(x,y)=\Delta(\hat{x},\hat{y})=\Delta(\xQuiet,\hat{y})-\Delta(\xQuiet,\hat{x})$, where $\hat{x}$ and $\hat{y}$ are the configurations that agree, respectively, with $x$ and $y$ on $M^{-1}(M(\diff(x,y)))$ and with $\xQuiet$ everywhere else. Therefore, we can write $\Delta(x,y)=\Delta(\hat{x},\hat{y})=\Delta_f(\hat{x},\hat{y})=\Delta_f(x,y)$. \end{proof} Whether the above proposition extends to finite-range Hamiltonians on strongly irreducible shifts of finite type is not known, but in~\cite{ChaMey13}, examples of shifts of finite type are given on which not every finite-range Hamiltonian is generated by a local observable. On the other hand, the main result of~\cite{ChaHanMarMeyPav13} implies that on a one-dimensional mixing shift of finite type, every finite-range Hamiltonian can be generated by a local observable. The \emph{trivial} Hamiltonian on $\xSp{X}$ (i.e., the Hamiltonian $\Delta$ for which $\Delta(x,y)=0$ for all asymptotic $x,y\in\xSp{X}$) plays a special role as it identifies an important notion of equivalence between observables (see Section~\ref{sec:physical-equivalence}). Another important concept regarding Hamiltonians is that of ground configurations. Let $\xSp{X}\subseteq S^\xL$ be a shift space and $\Delta$ a Hamiltonian on $\xSp{X}$. A \emph{ground} configuration for $\Delta$ is a configuration $z\in\xSp{X}$ such that $\Delta(z,x)\geq 0$ for every configuration $x\in\xSp{X}$ that is asymptotic to $z$. The existence of ground configurations is well known. We shall use it later in the proof of Theorem~\ref{thm:strongly-transitive-CA:no-cons-law}. \begin{proposition} \label{prop:Hamiltonian:ground} Every Hamiltonian on a shift space of finite type has at least one ground configuration. \end{proposition} \begin{proof} Let $\xSp{X}$ be a shift space of finite type and $\Delta$ a Hamiltonian on $\xSp{X}$. Let $I_1\subseteq I_2\subseteq \cdots$ be a chain of finite subsets of $\xL$ that is exhaustive (i.e., $\bigcup_n I_n=\xL$). For example, we could take $I_n=[-n,n]^d$ in $\xL=\xZ^d$. Let $z_0\in\xSp{X}$ be an arbitrary configuration, and construct a sequence of configurations $z_1,z_2,\ldots\in\xSp{X}$ as follows. For each $n$, choose $z_n\in\xSp{X}$ to be a configuration with $\diff(z_0,z_n)\subseteq I_n$ such that $\Delta(z_0,z_n)$ is minimum (i.e., $\Delta(z_0,z_n)\leq\Delta(z_0,x)$ for all $x\in\xSp{X}$ with $\diff(z_0,x)\subseteq I_n$). The minimum exists because $L_{I_n}(\xSp{X}\,|\,z_0)$ is finite. By compactness, there is a subsequence $n_1<n_2<\cdots$ such that $z_{n_i}$ converges. The limit $z\IsDef \lim_{i\to\infty} z_{n_i}$ is a ground configuration. To see this, let $x\in\xSp{X}$ be asymptotic to $z$, and choose $k$ such that $I_k\supseteq\diff(z,x)$. Since $\xSp{X}$ is of finite type, there is a $l\geq k$ such that for every two configuration $u,v\in\xSp{X}$ that agree on $I_l\setminus I_k$, there is a configuration $w\in\xSp{X}$ that agrees with $u$ on $I_l$ and with $v$ outside $I_k$. In particular, for every sufficiently large $i$, $x$ (and $z$) agree with $z_{n_i}$ on $I_l\setminus I_k$, and hence there is a configuration $x_{n_i}$ that agrees with $x$ on $I_l$ and with $z_{n_i}$ outside $I_k$. Then, $\diff(z_{n_i},x_{n_i})=\diff(z,x)$. Since $z_{n_i}\to z$, we also get $x_{n_i}\to x$. The continuity property of $\Delta$ now implies that $\Delta(z_{n_i},x_{n_i})\to\Delta(z,x)$. On the other hand, $\Delta(z_{n_i},x_{n_i})=\Delta(z_0,x_{n_i})-\Delta(z_0,z_{n_i})\geq 0$. Therefore, $\Delta(z,x)\geq 0$. \end{proof} \begin{example}[Ising model] \label{exp:ising} The \emph{Ising model} is a simple model on the lattice designed to give a statistical explanation of the phenomenon of spontaneous magnetization in ferromagnetic material (see e.g.~\cite{Vel09,Geo88}). The configuration space of the $d$-dimensional Ising model is the full shift $\xSp{X}\IsDef\{\symb{-},\symb{+}\}^{\xL}$, where $\xL=\xZ^d$, and where having $\symb{+}$ and~$\symb{-}$ at a site $i$ is interpreted as an upward or downward magnetization of the tiny segment of the material approximated by site $i$. The state of site $i$ is called the \emph{spin} at site $i$. The interaction between spins is modeled by associating an interaction energy $-1$ to every two adjacent spins that are aligned (i.e., both are upward or both downward) and energy $+1$ to every two adjacent spins that are not aligned. Alternatively, we can specify the energy using the energy observable $f\in K(\xSp{X})$ defined by \begin{align} f(x) &\IsDef \begin{cases} \frac{1}{2}(n^{\symb{-}}(x) - n^{\symb{+}}(x)) & \text{if $x(0)=\symb{+}$,} \\ \frac{1}{2}(n^{\symb{+}}(x) - n^{\symb{-}}(x)) & \text{if $x(0)=\symb{-}$,} \end{cases} \end{align} where $n^{\symb{+}}(x)$ and $n^{\symb{-}}(x)$ are, respectively, the number of upward and downward spins adjacent to site $0$. This defines a Hamiltonian $\Delta_f$. The two uniform configurations (all sites $\symb{+}$ and all sites $\symb{-}$) are ground configurations for $\Delta_f$, although $\Delta_f$ has many other ground configurations. \hfill$\ocircle$ \end{example} \begin{example}[Contour model] \label{exp:contour} The \emph{contour model} was originally used to study phase transition in the Ising model. Each site of two-dimensional lattice $\xL=\xZ^2$ may take a state from the set \begin{align} T &\IsDef \{ \,\ContourEmpty\,, \,\ContourH\,, \,\ContourV\,, \,\ContourLD\,, \,\ContourRD\,, \,\ContourRU\,, \,\ContourLU\,, \,\ContourX\, \} \;. \end{align} Not all configurations are allowed. The allowed configurations are those in which the state of adjacent sites match in the obvious fashion. For example, $\ContourLD$ can be placed on the right side of $\ContourH$ but not on top of it, and $\ContourEmpty$ can be placed on the left side of $\ContourRU$ but not on its right side. The allowed configurations depict decorations of the lattice formed by closed or bi-infinite paths (see Figure~\ref{fig:ising-contour}b). These paths are referred to as \emph{contours}. The space of allowed configurations $\xSp{Y}\subseteq T^\xZ$ is a shift space of finite type. It is also easy to verify that $(\xSp{Y},\sigma)$ is strongly irreducible. Define the local observable $g\in K(\xSp{Y})$, where \begin{align} g(x) &\IsDef \begin{cases} 0 &\text{if $x(0)=\ContourEmpty$,}\\ 1 &\text{if $x(0)\in \{ \,\ContourH\,, \,\ContourV\,, \,\ContourLD\,, \,\ContourRD\,, \,\ContourRU\,, \,\ContourLU\, \}$,} \\ 2 &\text{if $x(0)=\ContourX$.} \end{cases} \end{align} The Hamiltonian $\Delta_g$ simply compares the length of the contours in two asymptotic configurations. The uniform configuration in which every site is in state $\ContourEmpty$ is a ground configuration for $\Delta_g$. Any configuration with a single bi-infinite horizontal (or vertical) contour is also a ground configuration for $\Delta_g$. \hfill$\ocircle$ \end{example} \begin{example}[Hard-core gas] \label{exp:hard-core} Let $0\subseteq W\subseteq\xL$ be a neighborhood, and define a shift space $\xSp{X}\subseteq\{\symb{0},\symb{1}\}^\xL$ consisting of all configurations $x$ for which $W(i)\cap W(j)=\varnothing$ for every distinct $i,j\in\xL$ with $x(i)=x(j)=\symb{1}$. This is the configuration space of the \emph{hard-core gas} model. A site having state $\symb{1}$ is interpreted as containing a particle, whereas a site in state $\symb{0}$ is thought of to be empty. It is assumed that each particle occupies a volume $W$ and that the volume of different particles cannot overlap. The one-dimensional version of the hard-core shift with volume $W=\{0,1\}$ is also known as the \emph{golden mean} shift. The hard-core shift is clearly of finite type. It is also strongly irreducible. In fact, $\xSp{X}$ has a stronger irreducibility property: for every two asymptotic configurations $x,y\in\xSp{X}$, there is a sequence $x=x_0,x_1,\ldots,x_n=y$ of configurations in $\xSp{X}$ such that $\diff(x_i,x_{i+1})$ is singleton. In particular, Proposition~\ref{prop:Hamiltonian:observable:local} can be adapted to cover the Hamiltonians on~$\xSp{X}$. Let $h(x)\IsDef 1$ if $x(0)=\symb{1}$ and $h(x)\IsDef 0$ otherwise. The Hamiltonian $\Delta_h$ compares the number of particles on two asymptotic configurations. The empty configuration is the unique ground configuration for $\Delta_h$.% \hfill$\ocircle$ \end{example} Gibbs measures are a class of probability measures identified by Hamiltonians. Let $\xSp{X}$ be a shift space. A \emph{Gibbs measure} for a finite-range Hamiltonian $\Delta$ is a probability measure $\pi\in\xSx{P}(\xSp{X})$ satisfying \begin{align} \label{eq:gibbs:def:local} \pi([y]_E) &= \xe^{-\Delta(x,y)}\,\pi([x]_E) \end{align} for every two asymptotic configurations $x,y\in\xSp{X}$ and all sufficiently large $E$. (If $M$ is the interaction neighborhood of $\Delta$, the above equality will hold for every $E\supseteq M(\diff(x,y))$.) More generally, if $\Delta$ is an arbitrary Hamiltonian on $\xSp{X}$, a probability measure $\pi\in\xSx{P}(\xSp{X})$ is said to be a Gibbs measure for $\Delta$ if \begin{align} \label{eq:gibbs:def} \lim_{E\nearrow\xL} \frac{\pi([y]_E)}{\pi([x]_E)} &= \xe^{-\Delta(x,y)} \end{align} for every configuration $x\in\xSp{X}$ that is in the support of $\pi$ and every configurations $y\in\xSp{X}$ that is asymptotic to $x$. The limit is taken along the directed family of finite subsets of $\xL$ with inclusion.\footnote{% The original definition of a Gibbs measure given by Dobrushin, Lanford and Ruelle is via conditional probabilities (see e.g.~\cite{Geo88}). The definition given here can be shown to be equivalent to the original definition using the martingale convergence theorem. See Appendix~\ref{apx:gibbs:def:equivalence}. } The above limit is in fact uniform among all pairs of configurations $x,y$ in the support of $\pi$ whose disagreements $\diff(x,y)$ are included in a finite set $D\subseteq\xL$ (see Appendix~\ref{apx:gibbs:def:equivalence}). Note also that if $\xSp{X}$ is strongly irreducible, every Gibbs measure on $\xSp{X}$ has full support, and therefore, the relation~(\ref{eq:gibbs:def}) must hold for every two asymptotic $x,y\in\xSp{X}$. The set of Gibbs measures for a Hamiltonian $\Delta$, denoted by $\xSx{G}_\Delta(\xSp{X})$, is non-empty, closed and convex. According to the Krein-Milman theorem, the set $\xSx{G}_\Delta(\xSp{X})$ coincides with the closed convex hull of its extremal elements. The extremal elements of $\xSx{G}_\Delta(\xSp{X})$ are mutually singular. The subset $\xSx{G}_\Delta(\xSp{X},\sigma)$ of shift-invariant elements of $\xSx{G}_\Delta(\xSp{X})$ is also non-empty (using convexity and compactness), closed and convex, and hence equal to the closed convex hull of its extremal elements. The extremal elements of $\xSx{G}_\Delta(\xSp{X},\sigma)$ are precisely its ergodic elements, and hence again mutually singular. The Gibbs measures associated to finite-range Hamiltonians have the Markov property. A measure $\pi$ on a shift space $\xSp{X}\subseteq S^\xL$ is called a \emph{Markov measure} if there is a neighborhood $0\in M\subseteq\xL$ such that for every two finite sets $D,E\subseteq\xL$ with $M(D)\subseteq E$ and every pattern $p:E\to S$ with $\pi([p]_{E\setminus D})>0$ it holds \begin{equation} \pi\left([p]_D\,\middle|\, [p]_{E\setminus D}\right) = \pi\left([p]_D\,\middle|\, [p]_{M(D)\setminus D}\right) \;. \end{equation} The data contained in the conditional probabilities $\pi\left([p]_D\,\middle|\, [p]_{M(D)\setminus D}\right)$ for all choices of $D$, $E$ and $p$ is called the \emph{specification} of the Markov measure $\pi$. The specification of a Gibbs measure associated to a finite-range Hamiltonian is \emph{positive} (i.e., all the conditional distributions are positive) and \emph{shift-invariant}. Conversely, every positive shift-invariant Markovian specification is the specification of a Gibbs measure. In fact, Equation~(\ref{eq:gibbs:def:local}) identifies a one-to-one correspondence between finite-range Hamiltonians and the positive shift-invariant Markovian specifications. The uniform Bernoulli measure on a full shift $\xSp{X}\subseteq S^\xL$ is the unique Gibbs measure for the trivial Hamiltonian on $\xSp{X}$. More generally, the shift-invariant Gibbs measures on a strongly irreducible shift of finite type associated to the trivial Hamiltonian are precisely the measures that maximize the entropy (see below). We shall call a Gibbs measure \emph{regular} if its corresponding Hamiltonian is generated by an observable with summable variations. \subsection{Entropy, Pressure, and The Variational Principle} \label{sec:entropy-pressure} Statistical mechanics attempts to explain the macroscopic behaviour of a physical system by statistical analysis of its microscopic details. In the subjective interpretation (see~\cite{Jay57}), the probabilities reflect the partial knowledge of an observer. A suitable choice for a probability distribution over the possible microscopic states of a system is therefore one which, in light of the available partial observations, is least presumptive. The standard approach to pick the least presumptive probability distribution is by maximizing entropy. The characterization of the uniform probability distribution over a finite set as the probability distribution that maximizes entropy is widely known. Maximizing entropy subject to partial observations leads to Boltzmann distribution. The infinite systems based on lattice configurations have a similar (though more technical) picture. Below, we give a minimum review necessary for our discussion. Details and more information can be found in the original monographs and textbooks~\cite{Rue04,Isr79,Geo88,Sim93,Kel98}. The equilibrium statistical mechanics, which can be built upon the maximum entropy postulate, has been enormously successful in predicting the absence or presence of phase transitions, and in describing the qualitative features of the phases; see~\cite{Geo88}. Let $\xSp{X}\subseteq S^\xL$ be a shift space and $\pi\in\xSx{P}(\xSp{X})$ a probability measure on $\xSp{X}$. The \emph{entropy} of a finite set $A\subseteq\xL$ of sites under $\pi$ is \begin{equation} H_\pi(A) \IsDef -\sum_{p\in L_A(\xSp{X})} \pi([p]_A)\log\pi([p]_A) \;. \end{equation} (By convention, $0\log 0\IsDef 0$.) This is the same as the Shannon entropy of the random variable $\xv{x}_A$ in the probability space $(\xSp{X},\pi)$, where $\xv{x}_A$ is the projection $x\mapsto\xRest{x}{A}$. Let us recall few basic properties of the entropy. The entropy $H(\xv{x})$ of a random variable $\xv{x}$ is non-negative. If $\xv{x}$ takes its values in a finite set of cardinality $n$, then $H(\xv{x})\leq\log n$. The entropy is sub-additive, meaning that $H((\xv{x},\xv{y}))\leq H(\xv{x})+H(\xv{y})$ for every two random variables $\xv{x}$ and $\xv{y}$. If $\xv{y}=f(\xv{x})$ depends deterministically on $\xv{x}$, we have $H(f(\xv{x}))\leq H(\xv{x})$. Let $I_n\IsDef [-n,n]^d\subseteq\xL$ be the centered $(2n+1)\times (2n+1)\times \cdots\times (2n+1)$ box in the lattice. If $\pi$ is shift-invariant, the sub-additivity of $A\mapsto H_\pi(A)$ ensures that the limit \begin{equation} h_\pi(\xSp{X},\sigma)\IsDef \lim_{n\to\infty} \frac{H_\pi(I_n)}{\abs{I_n}} = \inf_{n\geq 0} \frac{H_\pi(I_n)}{\abs{I_n}} \end{equation} exists (Fekete's lemma). The limit value $h_\pi(\xSp{X},\sigma)$ is the average entropy per site of $\pi$ over $\xSp{X}$. It is also referred to as the \emph{(Kolmogorov-Sinai) entropy} of the dynamical system $(\xSp{X},\sigma)$ under $\pi$ (see~\cite{Wal82}, Theorem~4.17). The entropy functional $\pi \mapsto h_\pi(\xSp{X},\sigma)$ is non-negative and affine. Although it is not continuous, it is upper semi-continuous. \begin{proposition}[Upper Semi-continuity] If $\lim_{i\to\infty}\pi_i=\pi$, then $\limsup_{i\to\infty}h_{\pi_i}(\xSp{X},\sigma) \leq h_\pi(\xSp{X},\sigma)$. \end{proposition} \begin{proof} The pointwise infimum of a family of continuous functions is upper semi-continuous. \end{proof} The entropy functional is also bounded. Due to the compactness of $\xSx{P}(\xSp{X},\sigma)$ and the upper semi-continuity of $\pi\mapsto h_\pi(\xSp{X},\sigma)$, the entropy $h_\pi(\xSp{X},\sigma)$ takes its maximum value at some measures $\pi\in\xSx{P}(\xSp{X},\sigma)$. This maximum value coincides with the \emph{topological entropy} of the shift $(\xSp{X},\sigma)$, defined by \begin{align} h(\xSp{X},\sigma)\IsDef \lim_{n\to\infty} \frac{\log\abs{L_{I_n}(\xSp{X})}}{\abs{I_n}} = \inf_{n\geq 0} \frac{\log\abs{L_{I_n}(\xSp{X})}}{\abs{I_n}} \;, \end{align} which is the average combinatorial entropy per site of $\xSp{X}$. The following propositions are easy to prove, and are indeed valid for arbitrary dynamical systems. \begin{proposition}[Factoring] \label{prop:entropy:factoring} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a factor map between two shifts $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$ and $\pi\in\xSx{P}(\xSp{X},\sigma)$ a probability measure on $\xSp{X}$. Then, $h_{\Phi\pi}(\xSp{Y},\sigma)\leq h_\pi(\xSp{X},\sigma)$. \end{proposition} \begin{proposition}[Embedding] Let $\Phi:\xSp{Y}\to\xSp{X}$ be an embedding of a shift $(\xSp{Y},\sigma)$ in a shift $(\xSp{X},\sigma)$ and $\pi\in\xSx{P}(\xSp{Y},\sigma)$ a probability measure on $\xSp{Y}$. Then, $h_\pi(\xSp{Y},\sigma)= h_{\Phi\pi}(\xSp{X},\sigma)$. \end{proposition} Given a continuous observable $f\in C(\xSp{X})$, the mapping $\pi\in\xSx{P}(\xSp{X},\sigma)\mapsto \pi(f)$ is continuous and affine. Its range is closed, bounded, and convex, that is, a finite closed interval $[e_{\min},e_{\max}]\subseteq\xR$. For each $e\in [e_{\min},e_{\max}]$, let us define \begin{equation} s_f(e) \IsDef \sup\left\{ h_\pi(\xSp{X},\sigma): \pi\in\xSx{P}(\xSp{X},\sigma) \text{ and } \pi(f)=e \right\} \;. \end{equation} Let $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ denote the set of measures $\pi\in\xSx{P}(\xSp{X},\sigma)$ with $\pi(f)=e$ and $h_\pi(\xSp{X},\sigma)=s_f(e)$, that is, the measures $\pi$ that maximize entropy under the constraint $\pi(f)=e$. By the compactness of $\xSx{P}(\xSp{X},\sigma)$ and the upper semi-continuity of $\pi\mapsto h_\pi(\xSp{X},\sigma)$, the set $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ is non-empty (as long as $e\in [e_{\min},e_{\max}]$). The mapping $s_f(\cdot)$ is concave and continuous. The measures in $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ (and more generally, the solutions of similar entropy maximization problems with multiple contraints $\pi(f_1)=e_1$, $\pi(f_2)=e_2$, \ldots, $\pi(f_n)=e_n$) could be implicitly identified after a Legendre transform. The \emph{pressure} associated to $f\in C(\xSp{X})$ could be defined as \begin{align} \label{eq:pressure:def} P_f(\xSp{X},\sigma) &= \sup_{\nu\in\xSx{P}(\xSp{X},\sigma)}[ h_\nu(\xSp{X},\sigma) - \nu(f) ] \;. \end{align} The functional $f\mapsto P_f(\xSp{X},\sigma)$ is convex and Lipschitz continuous. It is the convex conjugate of the entropy functional $\nu\mapsto h_\nu(\xSp{X},\sigma)$ (up to a negative sign), and we also have \begin{align} h_\pi(\xSp{X},\sigma) &= \inf_{g\in C(\xSp{X})}[ P_g(\xSp{X},\sigma) + \pi(g) ] \end{align} (see~\cite{Rue04}, Theorem~3.12). Note that the pressure $P_0(\xSp{X},\sigma)$ associated to $0$ is the same as the topological entropy of $(\xSp{X},\sigma)$. Again, the compactness of $\xSx{P}(\xSp{X},\sigma)$ and the upper semi-continuity of $\nu\mapsto h_\nu(\xSp{X},\sigma)$ ensure that the supremum in~(\ref{eq:pressure:def}) can be achieved. The set of shift-invariant probability measures $\pi\in\xSx{P}(\xSp{X},\sigma)$ for which the equality in \begin{align} h_\pi(\xSp{X},\sigma) - P_f(\xSp{X},\sigma) \leq \pi(f) \end{align} is satisfied will be denoted by $\xSx{E}_f(\xSp{X},\sigma)$. Following the common terminology of statistical mechanics and ergodic theory, we call the elements of $\xSx{E}_f(\xSp{X},\sigma)$ the \emph{equilibrium measures} for $f$. Let us emphasize that this terminology lacks a dynamical justification that we are striving for. The Bayesian justification is further clarified below. A celebrated theorem of Dobrushin, Lanford and Ruelle characterizes the equilibrium measures (for ``short-ranged'' observables over strongly irreducible shift spaces of finite type) as the associated shift-invariant Gibbs measures. \begin{theorem}[% Characterization of Equilibrium Measures; see~\cite{Rue04}, Theorem~4.2, and~\cite{Kel98}, Sections~5.2 and~5.3, and~\cite{Mey13}% ] \label{thm:equilibrium:Gibbs} Let $\xSp{X}\subseteq S^\xL$ be a strongly irreducible shift space of finite type. Let $f\in SV(\xSp{X})$ be an observable with summable variations and $\Delta_f$ the Hamiltonian it generates. The set of equilibrium measures for $f$ coincides with the set of shift-invariant Gibbs measures for $\Delta_f$. \end{theorem} Consider now an observable $f\in C(\xSp{X})$, and as before, let $[e_{\min},e_{\max}]$ be the set of possible values $\nu(f)$ for $\nu\in\xSx{P}(\xSp{X},\sigma)$. For every $\beta\in\xR$, we have \begin{align} P_{\beta f}(\xSp{X},\sigma) &= \sup\{s_f(e) - \beta e: e\in [e_{\min},e_{\max}]\} \;. \end{align} That is, $\beta\in\xR\mapsto P_{\beta f}$ is the Legendre transform of $e\mapsto s_f(e)$. If $f$ has summable variations and the Hamiltonian $\Delta_f$ is not trivial, it can be shown that $\beta\mapsto P_{\beta f}$ is strictly convex (see~\cite{Rue04}, Section~4.6, or~\cite{Isr79}, Section~III.4). It follows that $e\mapsto s_f(e)$ is continuously differentiable everywhere except at $e_{\min}$ and $e_{\max}$, and \begin{align} s_f(e) &= \inf\{P_{\beta f}(\xSp{X},\sigma)+\beta e: \beta\in\xR\} \end{align} for every $e\in (e_{\min},e_{\max})$. For $e\in (e_{\min},e_{\max})$, the above theorem identifies the elements of $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ as the shift-invariant Gibbs measures for $\beta_e f$, where $\beta_e\in\xR$ is the unique value at which $\beta\mapsto -P_{\beta f}(\xSp{X},\sigma)$ has a tangent with slope~$e$. The mapping $e\mapsto\beta_e$ is continuous and non-increasing. The set of slopes of tangents to $\beta\mapsto -P_{\beta f}(\xSp{X},\sigma)$ at a point $\beta\in\xR$ is a closed interval $[e^-_\beta,e^+_\beta]\subseteq(e_{\min},e_{\max})$. We have \begin{equation} \xSx{E}_{\beta f}(\xSp{X},\sigma)= \bigcup_{e\in[e^-_\beta,e^+_\beta]} \xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma) \;. \end{equation} When $f$ is interpreted as the energy contribution of a single site, $1/\beta$ is interpreted as the temperature and $e$ as the mean energy per site. By a Bayesian reasoning, if $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ is singleton, its unique element is an appropriate choice of the probability distribution of the system in thermal equilibrium when the mean energy per site is $e$. If $\xSx{E}_{\beta f}(\xSp{X},\sigma)$ is singleton, the unique element is interpreted as a description of the system in thermal equilibrium at temperature $1/\beta$. The existence of more than one element in $\xSx{E}_{\beta f}(\xSp{X},\sigma)$ (or in $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$) is interpreted as the existence of more than one phase (e.g., liquid or gas) at temperature $1/\beta$ (resp., with energy density $e$). The presence of distinct tangents to $\beta\mapsto-P_{\beta f}(\xSp{X},\sigma)$ at a given inverse temperature $\beta$ implies the existence of distinct phases at temperature $1/\beta$ having different mean energy per site. Note that since the elements of $\xSx{E}_{\beta f}(\xSp{X},\sigma)=\xSx{G}_{\beta\Delta_f}(\xSp{X},\sigma)$ are shift-invariant, they only offer a description of the equilibrium states that respect the translation symmetry of the model. By extrapolating the interpretation, one could consider the Gibbs measures $\pi\in\xSx{G}_{\beta\Delta_f}(\xSp{X})$ that are not shift-invariant as states of equilibrium in which the translation symmetry is broken.% \subsection{Physical Equivalence of Observables} \label{sec:physical-equivalence} Let $\xSp{X}\subseteq S^{\xL}$ be a strongly irreducible shift space of finite type. Every local observable generates a finite-range Hamiltonian via Equation~(\ref{eq:hamiltonian:observable}). However, different local observables may generate the same Hamiltonians. Two local observables $f,g\in K(\xSp{X})$ are \emph{physically equivalent} (see~\cite{Rue04}, Sections~4.6-4.7, \cite{Isr79}, Sections~I.4 and~III.4, or~\cite{Geo88}, Section~2.4), $f\sim g$ in symbols, if they identify the same Hamiltonian, that is, if $\Delta_f=\Delta_g$. The following proposition gives an alternate characterization of physical equivalence, which will allow us to extend the notion of physical equivalence to~$C(\xSp{X})$. \begin{proposition} \label{prop:physical-equivalence:local} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type. Two observables $f,g\in K(\xSp{X})$ are physically equivalent, if and only if there is a constant $c\in\xR$ such that $\pi(f)=\pi(g)+c$ for every probability measure $\pi\in\xSx{P}(\xSp{X},\sigma)$. \end{proposition} \begin{proof}\ \\ \begin{enumerate}[ $\Rightarrow$)] \item[ $\Rightarrow$)] Let $h\IsDef f-g$. Let us pick an arbitrary configuration $\xQuiet\in\xSp{X}$ with the property that the spatial average \begin{equation} c\IsDef \lim_{n\to\infty} \frac{\sum_{i\in I_n} h(\sigma^i \xQuiet)}{\abs{I_n}} \end{equation} (where $I_n\IsDef[-n,n]^d\subseteq\xL$) exists. That such a configuration exists follows, for example, from the ergodic theorem.\footnote{% We could simply choose $\xQuiet$ to be a periodic configuration if we knew such a configuration existed. Unfortunately, it is not known whether every strongly irreducible shift of finite type (in more than two dimensions) has a periodic configuration. } We claim that indeed \begin{equation} \lim_{n\to\infty} \frac{\sum_{i\in I_n} h(\sigma^i x)}{\abs{I_n}} = c \end{equation} for every configuration $x\in\xSp{X}$. This follows from the fact that $\Delta_h=\Delta_f-\Delta_g=0$. More specifically, let $0\in M\subseteq\xL$ be a neighborhood that witnesses the strong irreducibility of $\xSp{X}$, and let $D\subseteq\xL$ be a finite base for $h$ (i.e., $h$ is $\family{F}_D$-measurable). For each configuration $x\in\xSp{X}$ and each $n\geq 0$, let $x_n$ be a configuration that agrees with $x$ on $I_n+D$ and with $\xQuiet$ off $I_n+D+M-M$. Then \begin{align} \sum_{i\in I_n} h(\sigma^i x) &= \sum_{i\in I_n} h(\sigma^i x_n) \\ &= \sum_{i\in I_n} h(\sigma^i\xQuiet) + \Delta_h(\xQuiet,x_n) + \smallo(\abs{I_n}) \\ &= \sum_{i\in I_n} h(\sigma^i\xQuiet) + \smallo(\abs{I_n}) \;, \end{align} and the claim follows. Now, the dominated convergence theorem concludes that $\pi(f)-\pi(g)=\pi(h)=c$, for every $\pi\in\xSx{P}(\xSp{X},\sigma)$. \item[ $\Leftarrow$)] Following the definition, $P_f(\xSp{X},\sigma)=P_g(\xSp{X},\sigma)-c$, and $f$ and $g$ have the same equilibrium measures. Theorem~\ref{thm:equilibrium:Gibbs} then implies that the shift-invariant Gibbs measures of $\Delta_f$ and $\Delta_g$ coincide, which in turn concludes that $\Delta_f=\Delta_g$. \end{enumerate} \end{proof} As a corollary, the physical equivalence relation is closed in $K(\xSp{X})\times K(\xSp{X})$: \begin{corollary} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type. Let $h_1,h_2,\ldots$ be local observables on $\xSp{X}$ such that $\Delta_{h_i}=0$ for each $i$. If $h_i$ converge to a local observable $h$, then $\Delta_h=0$. \end{corollary} The continuous extension of this relation (i.e., the closure of $\sim$ in $C(\xSp{X})\times C(\xSp{X})$) gives a notion of physical equivalence of arbitrary continuous observables. \begin{proposition} \label{prop:physical-equivalence:continuous} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type. Two observables $f,g\in C(\xSp{X})$ are physically equivalent if and only if there is a constant $c\in\xR$ such that $\pi(f)=\pi(g)+c$ for every probability measure $\pi\in\xSx{P}(\xSp{X},\sigma)$. \end{proposition} \begin{proof} First, suppose that $f$ and $g$ are physically equivalent. Then, there exist sequences $f_1,f_2,\ldots$ and $g_1,g_2,\ldots$ of local observables such that $f_i\to f$, $g_i\to g$ and $f_i\sim g_i$. By Proposition~\ref{prop:physical-equivalence:local}, there are real numbers $c_i$ such that for every $\pi\in\xSx{P}(\xSp{X},\sigma)$, $\pi(f_i)-\pi(g_i)=c_i$. Taking the limits as $i\to\infty$, we obtain $\pi(f)-\pi(g)=c$, where $c\IsDef\lim_i c_i$ is independent of $\pi$. Conversely, suppose there is a constant $c\in\xR$ such that $\pi(f)=\pi(g)+c$ for every probability measure $\pi\in\xSx{P}(\xSp{X},\sigma)$. Let $h\IsDef f-g-c$. Then, according to Lemma~\ref{lem:annihilators}, $h\in C(\xSp{X},\sigma)$. Therefore, by the denseness of $K(\xSp{X})$ in $C(\xSp{X})$, there exists a sequence of \emph{local} observables $h_i\in C(\xSp{X},\sigma)$ such that $h_i\to h$. Choose another sequence of local observables $g_i$ that converges to $g$, and set $f_i\IsDef h_i+g_i+c$. By Lemma~\ref{lem:annihilators}, $\pi(f_i)=\pi(g_i)+c$ for every $\pi\in\xSx{P}(\xSp{X},\sigma)$, which along with Proposition~\ref{prop:physical-equivalence:local}, implies that $\Delta_{f_i}=\Delta_{g_i}$. Taking the limit, we obtain that $f$ and $g$ are physically equivalent. \end{proof} Using Lemma~\ref{lem:annihilators}, we also get the following characterization. \begin{corollary} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type. Two observables $f,g\in C(\xSp{X})$ are physically equivalent if and only if $f-g-c\in C(\xSp{X},\sigma)$ for some $c\in\xR$. \end{corollary} Physically equivalent observables define the same set of equilibrium measures. Moreover, the equilibrium measures of two observables with summable variations that are not physically equivalent are disjoint. (However, continuous observables that are not physically equivalent might in general share equilibrium measures; see~\cite{Rue04}, Corollary~3.17.) \begin{proposition} \label{prop:equilibrium:equivalence} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type. If two observables $f,g\in C(\xSp{X})$ are physically equivalent, they have the same set of equilibrium measures. Conversely, if two observables $f,g\in SV(\xSp{X})$ with summable variations share an equilibrium measure, they are physically equivalent. \end{proposition} \begin{proof} The first claim is an easy consequence of the characterization of physical equivalence given in Proposition~\ref{prop:physical-equivalence:continuous}. The converse follows from the characterization of equilibrium measures as Gibbs measures (Theorem~\ref{thm:equilibrium:Gibbs}). \end{proof} \section{Entropy-preserving Maps} \label{sec:entropy-preserving} \subsection{Entropy and Pre-injective Maps} The Garden-of-Eden theorem states that a cellular automaton over a strongly irreducible shift of finite type is surjective if and only if it is pre-injective~\cite{Moo62,Myh63,Hed69,CecMacSca99,Fio03}. This is one of the earliest results in the theory of cellular automata, and gives a characterization of when a cellular automaton has a so-called \emph{Garden-of-Eden}, that is, a configuration with no pre-image. The Garden-of-Eden theorem can be proved by a counting argument. Alternatively, the argument can be phrased in terms of entropy (see~\cite{LinMar95}, Theorem~8.1.16 and~\cite{CecCor10}, Chapter~5). \begin{theorem}[see~\cite{LinMar95}, Theorem~8.1.16 and~\cite{MeeSte01}, Theorem~3.6] \label{thm:entropy:pre-injective} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a factor map from a strongly irreducible shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Then, $h(\xSp{Y},\sigma)\leq h(\xSp{X},\sigma)$ with equality if and only if $\Phi$ is pre-injective. \end{theorem} \begin{theorem}[see~\cite{CovPau74}, Theorem~3.3, and~\cite{MeeSte01}, Lemma~4.1, and~\cite{Fio03}, Lemma~4.4] \label{thm:entropy:embedding} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type and $\xSp{Y}\subseteq\xSp{X}$ a proper subsystem. Then, $h(\xSp{Y},\sigma)<h(\xSp{X},\sigma)$. \end{theorem} \begin{corollary}[Garden-of-Eden Theorem~\cite{Moo62,Myh63,CecMacSca99,Fio03}] Let $\Phi:\xSp{X}\to\xSp{X}$ be a cellular automaton on a strongly irreducible shift of finite type $(\xSp{X},\sigma)$. Then, $\Phi$ is surjective if and only if it is pre-injective. \end{corollary} \begin{proof} \begin{align} \begin{array}{c@{\ }c@{\ }c@{\ }c@{\ }c} \xSp{X} & \xrightarrow{\ \quad\Phi\ \quad} & \Phi\xSp{X} & \xrightarrow{\ \quad\subseteq\ \quad} & \xSp{X} \;, \\[1ex] \mathclap{h(\xSp{X},\sigma)} & \geq & \mathclap{h(\Phi\xSp{X},\sigma)} & \leq & \mathclap{h(\xSp{X},\sigma)} \;. \end{array} \end{align} \end{proof} Another corollary of Theorem~\ref{thm:entropy:pre-injective} (along with Lemma~\ref{lem:onto-one-to-one} and Proposition~\ref{prop:entropy:factoring}) is the so-called \emph{balance} property of pre-injective cellular automata. \begin{corollary}[see~\cite{CovPau74}, Theorem~2.1, and~\cite{MeeSte01}, Theorems~3.3 and~3.6] \label{cor:balance:weak} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map from a strongly irreducible shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Every maximum entropy measure $\nu\in\xSx{P}(\xSp{Y},\sigma)$ has a maximum entropy pre-image $\pi\in\xSx{P}(\xSp{X},\sigma)$. \end{corollary} \noindent In particular, a cellular automaton on a full shift is surjective if and only if it preserves the uniform Bernoulli measure~\cite{Hed69,MarKim76}. In Section~\ref{sec:ca:invariance}, we shall find a generalization of this property. The probabilistic version of Theorem~\ref{thm:entropy:pre-injective} states that the pre-injective factor maps preserve the entropy of shift-invariant probability measures, and seems to be part of the folklore (see e.g.~\cite{HelLinNor07}). \begin{theorem} \label{thm:pentropy:pre-injective} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a factor map from a shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Let $\pi\in\xSx{P}(\xSp{X},\sigma)$ be a probability measure. Then, $h_{\Phi\pi}(\xSp{Y},\sigma)\leq h_\pi(\xSp{X},\sigma)$ with equality if $\Phi$ is pre-injective. \end{theorem} \begin{proof} For any factor map $\Phi$, the inequality $h_{\Phi\pi}(\xSp{Y},\sigma)\leq h_\pi(\xSp{X},\sigma)$ holds by Proposition~\ref{prop:entropy:factoring}. Suppose that $\Phi$ is pre-injective. It is enough to show that $h_{\Phi\pi}(\xSp{Y},\sigma)\geq h_\pi(\xSp{X},\sigma)$. Let $0\subseteq M\subseteq\xL$ be a neighborhood for $\Phi$ and a witness for the finite-type gluing property of $\xSp{X}$ (see Section~\ref{sec:shifts-ca}). Let $A\subseteq\xL$ be a finite set. By the pre-injectivity of $\Phi$, for every $x\in\xSp{X}$, the pattern $\xRest{x}{A\setminus M(\xCmpl{A})}$ is uniquely determined by $\xRest{x}{\partial M(A)}$ and $\xRest{(\Phi x)}{A}$. Indeed, suppose that $x'\in\xSp{X}$ is another configuration with $\xRest{x'}{\partial M(A)}=\xRest{x}{\partial M(A)}$ and $\xRest{(\Phi x')}{A}=\xRest{(\Phi x)}{A}$. Then, the configuration $x''$ that agrees with $x$ on $M(\xCmpl{A})$ and with $x'$ on $M(A)$ is in $\xSp{X}$ and asymptotic to~$x$. Since $\xRest{(\Phi x'')}{A}=\xRest{(\Phi x')}{A}=\xRest{(\Phi x)}{A}$, it follows that $\Phi x''=\Phi x$. Therefore, $x''=x$, and in particular, $\xRest{x'}{A\setminus M(\xCmpl{A})}=\xRest{x''}{A\setminus M(\xCmpl{A})}=\xRest{x}{A\setminus M(\xCmpl{A})}$. From the basic properties of the Shannon entropy, it follows that \begin{align} H_\pi(\partial M(A)) + H_{\Phi\pi}(A) &\geq H_\pi(A\setminus M(\xCmpl{A})) \;. \end{align} Now, choose the neighborhood $M$ to be $I_r\IsDef [-r,r]^d\subseteq\xL$ for a sufficiently large $r$. For $A\IsDef I_n=[-n,n]^d$, we obtain \begin{align} H_{\Phi\pi}(I_n) &\geq H_\pi(I_{n-r}) - H_\pi(I_{n+r}\setminus I_{n-r}) \;. \end{align} Dividing by $\abs{I_n}$ we get \begin{align} \frac{H_{\Phi\pi}(I_n)}{\abs{I_n}} &\geq \frac{\abs{I_{n-r}}}{\abs{I_n}}\cdot\frac{H_\pi(I_{n-r})}{\abs{I_{n-r}}} - \frac{\smallo(\abs{I_n})}{\abs{I_n}} \;, \end{align} which proves the theorem by letting $n\to\infty$. \end{proof} From Theorem~\ref{thm:pentropy:pre-injective} and Lemma~\ref{lem:onto-one-to-one}, it immediately follows that the functionals $f\mapsto P_f$ and $f\mapsto s_f(\cdot)$ are preserved under the dual of a pre-injective factor map. \begin{corollary} \label{cor:pressure:pre-injective} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a factor map from a shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Let $f\in C(\xSp{Y})$ be an observable. Then, $P_{f\xO\Phi}(\xSp{X},\sigma)\geq P_f(\xSp{Y},\sigma)$ with equality if $\Phi$ is pre-injective. \end{corollary} \begin{corollary} \label{cor:bayesian:entropy:factor-map} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a factor map from a shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Let $f\in C(\xSp{Y})$ be an observable. Then, $s_{f\xO\Phi}(\cdot)\geq s_f(\cdot)$ with equality if $\Phi$ is pre-injective. \end{corollary} Central to this article is the the following correspondence between the equilibrium (Gibbs) measures of a model and its pre-injective factors. \begin{corollary} \label{thm:equilibrium:factor-map} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map from a shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Let $f\in C(\xSp{Y})$ be an observable and $\pi\in \xSx{P}(\xSp{X},\sigma)$ a probability measure. Then $\pi\in\xSx{E}_{f\xO\Phi}(\xSp{X},\sigma)$ if and only if $\Phi\pi\in\xSx{E}_f(\xSp{Y},\sigma)$. \end{corollary} \begin{corollary} \label{thm:bayesian:solutions:factor-map} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map from a shift of finite type $(\xSp{X},\sigma)$ onto a shift $(\xSp{Y},\sigma)$. Let $f\in C(\xSp{Y})$ be an observable, $\pi\in \xSx{P}(\xSp{X},\sigma)$ a probability measure, and $e\in\xR$. Then, $\pi\in\xSx{E}_{\langle f\xO\Phi\rangle=e}(\xSp{X},\sigma)$ if and only if $\Phi\pi\in\xSx{E}_{\langle f\rangle=e}(\xSp{Y},\sigma)$. \end{corollary} \begin{example} Let $\xSp{X}\subseteq\{\symb{0},\symb{1},\symb{2}\}^\xZ$ be the shift obtained by forbidding $\symb{1}\symb{1}$ and $\symb{2}\symb{2}$, and $\xSp{Y}\subseteq\{\symb{0},\symb{1},\symb{2}\}^\xZ$ the shift obtained by forbidding $\symb{2}\symb{2}$ and $\symb{2}\symb{1}$. Then, both $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$ are mixing shifts of finite type. For every configuration $x\in\xSp{X}$, let $\Phi x\in\{\symb{0},\symb{1},\symb{2}\}^\xZ$ be the configuration in which \begin{align} (\Phi x)(i) &= \begin{cases} \symb{1} \qquad & \text{if $x(i)=\symb{2}$ and $x(i+1)=\symb{1}$,}\\ x(i) & \text{otherwise.} \end{cases} \end{align} Then, $\Phi$ is a pre-injective factor map from $\xSp{X}$ onto $\xSp{Y}$. Consider the observables $g_0,g_1,g_2:\xSp{Y}\to\xR$ defined by \begin{align} g_0(y) &\IsDef 1_{[\symb{0}]_0}(y)\;, & g_1(y) &\IsDef 1_{[\symb{1}]_0}(y)\;, & g_2(y) &\IsDef 1_{[\symb{2}]_0}(y) \end{align} for $y\in\xSp{Y}$. The Hamiltonians $\Delta_{g_0}$, $\Delta_{g_1}$ and $\Delta_{g_2}$ count the number of $\symb{0}$s, $\symb{1}$s and $\symb{2}$s, respectively. The unique Gibbs measures for $\Delta_{g_0}$, $\Delta_{g_1}$ and $\Delta_{g_2}$ are, respectively, the distribution of the bi-infinite Markov chains with transition matrices\\ \begin{align} A_0 &\approx {\small% \begin{matrix} \tsymb{0} \\ \tsymb{1} \\ \tsymb{2} \end{matrix} \begin{bmatrix} \OverLabel{0.230}{\symb{0}} & \OverLabel{0.626}{\symb{1}} & \OverLabel{0.144}{\symb{2}} \\ 0.230 & 0.626 & 0.144 \\ 1 & 0 & 0 \end{bmatrix} } \;, & A_1 &\approx {\small% \begin{matrix} \tsymb{0} \\ \tsymb{1} \\ \tsymb{2} \end{matrix} \begin{bmatrix} \OverLabel{0.528}{\symb{0}} & \OverLabel{0.194}{\symb{1}} & \OverLabel{0.278}{\symb{2}} \\ 0.528 & 0.194 & 0.278 \\ 1 & 0 & 0 \end{bmatrix} } \;, & A_2 &\approx {\small% \begin{matrix} \tsymb{0} \\ \tsymb{1} \\ \tsymb{2} \end{matrix} \begin{bmatrix} \OverLabel{0.461}{\symb{0}} & \OverLabel{0.461}{\symb{1}} & \OverLabel{0.078}{\symb{2}} \\ 0.461 & 0.461 & 0.078 \\ 1 & 1 & 0 \end{bmatrix} } \;. \end{align} In general, every finite-range Gibbs measure on a one-dimensional mixing shift of finite type is the distribution of a bi-infinite Markov chain and vice versa (see~\cite{Geo88}, Theorem~3.5 and~\cite{ChaHanMarMeyPav13}). The observables induced by $g_0$, $g_1$ and $g_2$ on $\xSp{X}$ via $\Phi$ satisfy \begin{align} (g_0\xO\Phi)(x) &= 1_{[\symb{0}]_0}(x)\;, & (g_1\xO\Phi)(x) &\IsDef (1_{[\symb{1}]_0}+1_{[\symb{2}\symb{1}]_0})(x)\;, & (g_2\xO\Phi)(x) &\IsDef 1_{[\symb{2}\symb{0}]_0}(x)\;. \end{align} for every $x\in\xSp{X}$. The unique Gibbs measures for $\Delta_{g_0\xO\Phi}$, $\Delta_{g_1\xO\Phi}$ and $\Delta_{g_2\xO\Phi}$ are, respectively, the distribution of the bi-infinite Markov chains with transition matrices\\ \begin{align} A_0 &\approx {\small% \begin{matrix} \tsymb{0} \\ \tsymb{1} \\ \tsymb{2} \end{matrix} \begin{bmatrix} \OverLabel{0.230}{\symb{0}} & \OverLabel{0.385}{\symb{1}} & \OverLabel{0.385}{\symb{2}} \\ 0.374 & 0 & 0.626 \\ 0.374 & 0.626 & 0 \end{bmatrix} } \;, & A_1 &\approx {\small% \begin{matrix} \tsymb{0} \\ \tsymb{1} \\ \tsymb{2} \end{matrix} \begin{bmatrix} \OverLabel{0.528}{\symb{0}} & \OverLabel{0.163}{\symb{1}} & \OverLabel{0.310}{\symb{2}} \\ 0.630 & 0 & 0.370 \\ 0.898 & 0.102 & 0 \end{bmatrix} } \;, & A_2 &\approx {\small% \begin{matrix} \tsymb{0} \\ \tsymb{1} \\ \tsymb{2} \end{matrix} \begin{bmatrix} \OverLabel{0.461}{\symb{0}} & \OverLabel{0.316}{\symb{1}} & \OverLabel{0.224}{\symb{2}} \\ 0.673 & 0 & 0.327 \\ 0.350 & 0.650 & 0 \end{bmatrix} } \;. \end{align} By Corollary~\ref{thm:equilibrium:factor-map}, we have $\Phi\pi_0=\nu_0$, $\Phi\pi_1=\nu_1$, and $\Phi\pi_2=\nu_2$. \hfill$\ocircle$ \end{example} \subsection{Complete Pre-injective Maps} In this section, we discuss the extension of Corollary~\ref{thm:equilibrium:factor-map} to the case of non-shift-invariant Gibbs measures (Conjecture~\ref{conj:gibbs:factor-map}, Theorem~\ref{thm:gibbs:factor-map:complete}, and Corollary~\ref{cor:gibbs:conjugacy}). We start with an example that deviates from the main line of this article (i.e., understanding macroscopic equilibrium in surjective cellular automata) but rather demonstrates an application of factor maps as a tool in the study of phase transitions in equilibrium statistical mechanics models. The (trivial) argument used in this example however serves as a model for the proof of Theorem~\ref{thm:gibbs:factor-map:complete}. \begin{example}[Ising and contour models] \label{exp:ising-contour} There is a natural correspondence between the two-dimensional Ising model (Example~\ref{exp:ising}) and the contour model (Example~\ref{exp:contour}). As before, let $\xSp{X}=\{\symb{+},\symb{-}\}^{\xZ^2}$ and $\xSp{Y}\IsDef T^{\xZ^2}$ denote the configuration spaces of the Ising model and the contour model. Define a sliding block map $\Theta:\xSp{X}\to T^{\xZ^2}$ with neighborhood $N\IsDef\{(0,0),(0,1),(1,0),(1,1)\}$ and local rule $\theta:\{\symb{+},\symb{-}\}^N\to T$, specified by {\small% \begin{align} \begin{array}{% c@{\>\mapsto\>}c|c@{\>\mapsto\>}c|c@{\>\mapsto\>}c|c@{\>\mapsto\>}c |c@{\>\mapsto\>}c|c@{\>\mapsto\>}c|c@{\>\mapsto\>}c|c@{\>\mapsto\>}c } \SymbCellBlock{+}{+}{+}{+}&\ContourEmpty & \SymbCellBlock{+}{+}{+}{-}&\ContourRD & \SymbCellBlock{+}{+}{-}{+}&\ContourLD & \SymbCellBlock{+}{+}{-}{-}&\ContourH & \SymbCellBlock{+}{-}{+}{+}&\ContourRU & \SymbCellBlock{+}{-}{+}{-}&\ContourV & \SymbCellBlock{+}{-}{-}{+}&\ContourX & \SymbCellBlock{+}{-}{-}{-}&\ContourLU \\ \hline \SymbCellBlock{-}{-}{-}{-}&\ContourEmpty & \SymbCellBlock{-}{-}{-}{+}&\ContourRD & \SymbCellBlock{-}{-}{+}{-}&\ContourLD & \SymbCellBlock{-}{-}{+}{+}&\ContourH & \SymbCellBlock{-}{+}{-}{-}&\ContourRU & \SymbCellBlock{-}{+}{-}{+}&\ContourV & \SymbCellBlock{-}{+}{+}{-}&\ContourX & \SymbCellBlock{-}{+}{+}{+}&\ContourLU \end{array} \end{align} }% (see Figure~\ref{fig:ising-contour}). Then, $\Theta$ is a factor map onto $\xSp{Y}$ and is pre-injective. In fact, $\Theta$ is $2$-to-$1$: every configuration $y\in\xSp{Y}$ has exactly two pre-images $x,x'\in\xSp{X}$, where $x'=-x$ (i.e., $x'$ is obtained from $x$ by flipping the direction of the spin at every site). Moreover, if $f$ denotes the energy observable for the Ising model and $g$ the contour length observable for the contour model, we have $\Delta_f=2\Delta_{g\xO\Theta}$. \begin{figure} \begin{center} \begin{tabular}{ccc} \begin{minipage}[c]{0.4\textwidth} \centering \includegraphics[width=0.95\textwidth]{figures/ising-40x30-b04.pdf} \end{minipage} & & \begin{minipage}[c]{0.4\textwidth} \centering \includegraphics[width=0.912\textwidth]{figures/contour-40x30-b04.pdf} \end{minipage} \medskip\\ (a) && (b) \end{tabular} \end{center} \caption{% A configuration of the Ising model (a) and its corresponding contour configuration (b). Black represents an upward spin. } \label{fig:ising-contour} \end{figure} This relationship, which was first discovered by Peierls~\cite{Pei36}, is used to reduce the study of the Ising model to the study of the contour model (see e.g.~\cite{Gri64,Vel09}). The Gibbs measures for $\beta\Delta_f$ represent the states of thermal equilibrium for the Ising model at temperature $1/\beta$. According to Corollary~\ref{thm:equilibrium:factor-map} (and Theorem~\ref{thm:equilibrium:Gibbs}), the shift-invariant Gibbs measures $\pi\in\xSx{G}_{\beta\Delta_f}(\xSp{X},\sigma)$ are precisely the $\Theta$-pre-images of the shift-invariant Gibbs measures $\nu\in\xSx{G}_{2\beta\Delta_g}(\xSp{Y},\sigma)$ for the contour model. In fact, in this case it is also easy to show that the $\Theta$-image of every Gibbs measure for $\beta\Delta_f$ (not necessarily shift-invariant) is a Gibbs measure for $2\beta\Delta_g$. Indeed, suppose that $\pi\in\xSx{G}_{\beta\Delta_f}(\xSp{X})$ is a Gibbs measure for $\beta\Delta_f$ and $\nu\IsDef\Theta\pi$ its image. Let $y,y'\in\xSp{Y}$ be asymptotic configurations, and $E\supseteq\diff(y,y')$ a sufficiently large finite set of sites. Let $x_1,x_2\in\xSp{X}$ be the pre-images of $y$, and $x'_1,x'_2\in\xSp{X}$ the pre-images of $y'$. Without loss of generality, we can assume that $x_1$ is asymptotic to $x'_1$, and $x_2$ is asymptotic $x'_2$. It is easy to see that $\Theta^{-1}[y]_E=[x_1]_{N(E)}\cup[x_2]_{N(E)}$ and $\Theta^{-1}[y']_E=[x'_1]_{N(E)}\cup[x'_2]_{N(E)}$. Note that the cylinders $[x_1]_{N(E)}$ and $[x_2]_{N(E)}$ are disjoint, and so are the cylinders $[x'_1]_{N(E)}$ and $[x'_2]_{N(E)}$. Therefore, \begin{align} \frac{\nu([y']_E)}{\nu([y]_E)} &= \frac{\pi(\Theta^{-1}[y']_E)}{\pi(\Theta^{-1}[y]_E)} = \frac{\pi([x'_1]_{N(E)}) + \pi([x'_2]_{N(E)})}{\pi([x_1]_{N(E)}) + \pi([x_2]_{N(E)})} \;. \end{align} Since $N(E)$ is large and $\pi$ is a Gibbs measure for $\beta\Delta_f$, we have $\pi([x'_1]_{N(E)})=\xe^{-\beta\Delta_f(x_1,x'_1)}\pi([x_1]_{N(E)})$ and $\pi([x'_2]_{N(E)})=\xe^{-\beta\Delta_f(x_2,x'_2)}\pi([x_2]_{N(E)})$. Since, $\Delta_f(x_1,x'_1)=\Delta_f(x_2,x'_2)=2\Delta_g(y,y')$, it follows that $\nu([y']_E)=\xe^{-2\beta\Delta_g(y,y')}\nu([y]_E)$. It has been proved that for any $0<\beta<\infty$, the contour model with Hamiltonian $2\beta\Delta_g$ has a unique Gibbs measure~\cite{Aiz80,Hig81}; the main difficulty is to show that the infinite contours are ``unstable'', in the sense that, under every Gibbs measure, the probability of appearance of an infinite contour is zero.\footnote{% In fact, the theorem of Aizenman and Higuchi states that the simplex of Gibbs measures for the two-dimensional Ising model at any temperature has at most two extremal elements. However, the uniqueness of the Gibbs measure for the contour model is implicit in their result, and constitutes the main ingredient of the proof. } Let us denote the unique Gibbs measure for $2\beta\Delta_g$ by $\nu_\beta$. It follows that the simplex of Gibbs measures for the Ising model at temperature $1/\beta$ is precisely $\Theta^{-1}\nu_\beta$. For, the set $\Theta^{-1}\nu_\beta$ includes $\xSx{G}_{\beta\Delta_f}(\xSp{X})$ (by the above observation) and is included in $\xSx{G}_{\beta\Delta_f}(\xSp{X},\sigma)$ (because $\nu_\beta$ must be shift-invariant). Therefore, the Gibbs measures for the Ising model at any temperature $1/\beta$ are shift-invariant and $\xSx{G}_{\beta\Delta_f}(\xSp{X})=\xSx{G}_{\beta\Delta_f}(\xSp{X},\sigma)=\Theta^{-1}\nu_\beta$. It is not difficult to show that if $\Phi:\xSp{X}\to\xSp{Y}$ is a continuous $k$-to-$1$ map between two compact metric spaces, then every probability measure on $\xSp{Y}$ has at most $k$ mutually singular pre-images under $\Phi$. In particular, the simplex $\xSx{G}_{\beta\Delta_f}(\xSp{X})=\xSx{G}_{\beta\Delta_f}(\xSp{X},\sigma)=\Theta^{-1}\nu_\beta$ of Gibbs measures for the Ising model at temperature $1/\beta$ has at most $2$~ergodic elements. Whether the Ising model at temperature $1/\beta$ has two ergodic Gibbs measures or one depends on a specific geometric feature of the typical contour configurations under the measure~$\nu_\beta$. Roughly speaking, the contours of a contour configuration divide the two-dimensional plane into disjoint clusters. A configuration with no infinite contour generates either one or no infinite cluster, depending on whether each site is surrounded by a finite or infinite number of contours. Note that since $\nu_\beta$ is ergodic, the number of infinite clusters in a random configuration chosen according to $\nu_\beta$ is almost surely constant. If $\nu_\beta$-almost every configuration has an infinite cluster, then it follows by symmetry that $\Theta^{-1}\nu_\beta$ contains two distinct ergodic measures, one in which the infinite cluster is colored with~$\symb{+}$ and one with~$\symb{-}$. The converse is also known to be true~\cite{Rus79}: if $\nu_\beta$-almost every configuration has no infinite cluster, then $\Theta^{-1}\nu_\beta$ has only one element. Contour representations are used to study a wide range of statistical mechanics models, and are particularly fruitful to prove the ``stability'' of ground configurations at low temperature (see e.g.~\cite{Sin82,Fer98}). \hfill$\ocircle$ \end{example} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map between two strongly irreducible shifts of finite type $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$. Let $f\in SV(\xSp{Y})$ be an observable having summable variations and $\Delta_f$ the Hamiltonian defined by $f$. Then, according to Theorem~\ref{thm:equilibrium:Gibbs}, the equilibrium measures of $f$ and $f\xO\Phi$ are precisely the shift-invariant Gibbs measures for the Hamiltonians $\Delta_f$ and $\Delta_{f\xO\Phi}$. A natural question is whether Corollary~\ref{thm:equilibrium:factor-map} remains valid for arbitrary Gibbs measures (not necessarily shift-invariant). If $\Phi:\xSp{X}\to\xSp{Y}$ is a morphism between two shifts and $\Delta$ is a Hamiltonian on $\xSp{Y}$, let us denote by $\Phi^*\Delta$, the Hamiltonian on $\xSp{X}$ defined by $(\Phi^*\Delta)(x,y)\IsDef \Delta(\Phi x,\Phi y)$. \begin{conjecture} \label{conj:gibbs:factor-map} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map between two strongly irreducible shifts of finite type $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$. Let $\Delta$ be a Hamiltonian on $\xSp{Y}$, and $\pi$ a probability measure on $\xSp{X}$. Then, $\pi$ is a Gibbs measure for~$\Phi^*\Delta$ if and only if $\Phi\pi$ is a Gibbs measure for~$\Delta$. \end{conjecture} One direction of the latter conjecture is known to be true for a subclass of pre-injective factor maps. Let us say that a pre-injective factor map $\Phi:\xSp{X}\to\xSp{Y}$ between two shifts is \emph{complete} if for every configuration $x\in\xSp{X}$ and every configuration $y'\in\xSp{Y}$ that is asymptotic to $y\IsDef\Phi x$, there is a (unique) configuration $x'\in\xSp{X}$ asymptotic to $x$ such that $\Phi x'= y'$. \begin{lemma} \label{lem:pre-injective:complete} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a complete pre-injective factor map between two shifts of finite type $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$. For every finite set $D\subseteq\xL$, there is a finite set $E\subseteq\xL$ such that every two asymptotic configurations $x,x'\in\xSp{X}$ with $\diff(\Phi x,\Phi x')\subseteq D$ satisfy $\diff(x,x')\subseteq E$. \end{lemma} \begin{proof} See Figure~\ref{fig:lemma:complete-pre-injective} for an illustration. For a configuration $x\in\xSp{X}$, let $\event{A}_x$ be the set of all configurations $x'$ asymptotic to $x$ such that $\diff(\Phi x,\Phi x')\subseteq D$. The set $\event{A}_x$ is finite. Therefore, there is a finite set $E_x$ such that all the elements of $\event{A}_x$ agree outside $E_x$. We claim that if $C_x\supseteq E_x$ is a large enough finite set of sites, then for every configuration $x_1\in [x]_{C_x}$, all the elements of $\event{A}_{x_1}$ agree outside $E_x$. To see this, suppose that $C_x$ is large, and consider a configuration $x_1\in[x]_{C_x}$. Let $x'_1$ be a configuration asymptotic to $x_1$ such that $\diff(\Phi x_1,\Phi x'_1)\subseteq D$. By the gluing property of $\xSp{Y}$, there is a configuration $y'\in\xSp{Y}$ that agrees with $\Phi x'_1$ in a large neighborhood of $D$ and with $\Phi x$ outside~$D$. Since $\Phi$ is a complete pre-injective factor map, there is a unique configuration $x'$ asymptotic to $x$ such that $\Phi x'=y'$. Now, by the gluing property of $\xSp{X}$, there is a configuration $x''_1$ that agrees with $x'$ in $C_x$ and with $x_1$ outside $E_x$. Since $C_x$ was chosen large, it follows that $\Phi x''_1=\Phi x'_1$. Since $x'_1$ and $x''_1$ are asymptotic, the pre-injectivity of $\Phi$ ensures that $x''_1=x'_1$. Therefore, $x_1$ and $x'_1$ agree outside $E_x$. The cylinders $[x]_{C_x}$ form an open cover of $\xSp{X}$. Therefore, by the compactness of $\xSp{X}$, there is a finite set $\event{I}\subseteq\xSp{X}$ such that $\bigcup_{x\in\event{I}} [x]_{C_x}\supseteq\xSp{X}$. The set $E\IsDef \bigcup_{x\in\event{I}} E_x$ has the desired property. \end{proof} \begin{figure} \begin{center} \begin{tikzpicture}[scale=1,line cap=round, >=stealth', hachA/.style={ decorate,decoration={snake,amplitude=1pt,segment length=5pt},thick }, hachB/.style={ decorate,decoration={crosses,segment length=3pt,shape height=3pt,shape width=3pt},thick }, hachAA/.style={ decorate,decoration={zigzag,amplitude=1pt,segment length=5pt},thick }, hachBB/.style={ decorate,decoration={shape backgrounds,shape=diamond,shape size=3pt,shape sep=4pt},thick } ] \def\vdist{0.4} \def\eps{0.1} \def\vgap{3} \def6{6} \def\dwin{0.5} \def\ewin{1.7} \def\cwin{5.3} \def\crwin{4.8} \def\xwin{5.6} \def0.35{0.35} \coordinate[label=left:$\scriptstyle x$] (x) at (-6,0); \coordinate[label=left:$\scriptstyle x_1$] (x1) at (-6,-\vdist); \coordinate[label=left:$\scriptstyle x'_1$] (xp1) at (-6,-2*\vdist); \coordinate[label=left:$\scriptstyle x'$] (xp) at (-6,-3*\vdist); \coordinate[label=left:$\scriptstyle x''_1$] (xpp1) at (-6,-4*\vdist); \coordinate [label=left:$\scriptstyle \Phi x$] (fx) at (-6,-\vgap); \coordinate [label=left:$\scriptstyle \Phi x_1$] (fx1) at (-6,-\vgap-\vdist); \coordinate [label=left:$\scriptstyle \Phi x'_1$] (fxp1) at (-6,-\vgap-2*\vdist); \coordinate [label=left:$\scriptstyle y'$] (yp) at (-6,-\vgap-3*\vdist); \draw[very thick] (-6,0) -- (6,0); \draw[hachB] (-6,-\vdist) -- (-\cwin,-\vdist); \draw[hachB] (6,-\vdist) -- (\cwin,-\vdist); \draw[very thick] (-\cwin,-\vdist) -- (\cwin,-\vdist); \draw[hachB] (-6,-2*\vdist) -- (-\xwin,-2*\vdist); \draw[hachB] (6,-2*\vdist) -- (\xwin,-2*\vdist); \draw[dashed] (-\xwin,-2*\vdist) -- (\xwin,-2*\vdist); \draw[very thick] (-6,-3*\vdist) -- (-\ewin,-3*\vdist); \draw[very thick] (6,-3*\vdist) -- (\ewin,-3*\vdist); \draw[hachA] (-\ewin,-3*\vdist) -- (\ewin,-3*\vdist); \draw[hachB] (-6,-4*\vdist) -- (-\cwin,-4*\vdist); \draw[hachB] (6,-4*\vdist) -- (\cwin,-4*\vdist); \draw[very thick] (-\cwin,-4*\vdist) -- (-\ewin,-4*\vdist); \draw[very thick] (\cwin,-4*\vdist) -- (\ewin,-4*\vdist); \draw[hachA] (-\ewin,-4*\vdist) -- (\ewin,-4*\vdist); \draw[very thick] (-6,-\vgap) -- (6,-\vgap); \draw[hachBB] (-6,-\vgap-\vdist) -- (-\crwin,-\vgap-\vdist); \draw[hachBB] (6,-\vgap-\vdist) -- (\crwin,-\vgap-\vdist); \draw[very thick] (-\crwin,-\vgap-\vdist) -- (\crwin,-\vgap-\vdist); \draw[hachBB] (-6,-\vgap-2*\vdist) -- (-\crwin,-\vgap-2*\vdist); \draw[hachBB] (6,-\vgap-2*\vdist) -- (\crwin,-\vgap-2*\vdist); \draw[very thick] (-\crwin,-\vgap-2*\vdist) -- (-\dwin,-\vgap-2*\vdist); \draw[very thick] (\crwin,-\vgap-2*\vdist) -- (\dwin,-\vgap-2*\vdist); \draw[hachAA] (-\dwin,-\vgap-2*\vdist) -- (\dwin,-\vgap-2*\vdist); \draw[very thick] (-6,-\vgap-3*\vdist) -- (-\dwin,-\vgap-3*\vdist); \draw[very thick] (6,-\vgap-3*\vdist) -- (\dwin,-\vgap-3*\vdist); \draw[hachAA] (-\dwin,-\vgap-3*\vdist) -- (\dwin,-\vgap-3*\vdist); \draw[help lines] (-\ewin,0.35+\eps) -- (-\ewin,-4*\vdist-\eps); \draw[help lines] (\ewin,0.35+\eps) -- (\ewin,-4*\vdist-\eps); \draw[help lines, <->, shorten <=1pt, shorten >=1pt] (-\ewin,0.35) -- (\ewin,0.35) node[fill=white,pos=0.5] {$\scriptstyle E_x$}; \draw[help lines] (-\cwin,2*0.35+\eps) -- (-\cwin,-4*\vdist-\eps); \draw[help lines] (\cwin,2*0.35+\eps) -- (\cwin,-4*\vdist-\eps); \draw[help lines, <->, shorten <=1pt, shorten >=1pt] (-\cwin,2*0.35) -- (\cwin,2*0.35) node[fill=white,pos=0.5] {$\scriptstyle C_x$}; \draw[help lines] (-\dwin,-\vgap+0.35+\eps) -- (-\dwin,-\vgap-3*\vdist-\eps); \draw[help lines] (\dwin,-\vgap+0.35+\eps) -- (\dwin,-\vgap-3*\vdist-\eps); \draw[help lines, <->, shorten <=1pt, shorten >=1pt] (-\dwin,-\vgap+0.35) -- (\dwin,-\vgap+0.35) node[fill=white,pos=0.5] {$\scriptstyle D$}; \draw[help lines] (-\crwin,-\vgap+\eps) -- (-\crwin,-\vgap-3*\vdist-\eps); \draw[help lines] (\crwin,-\vgap+\eps) -- (\crwin,-\vgap-3*\vdist-\eps); \end{tikzpicture} \end{center} \caption{% Illustration of the proof of Lemma~\ref{lem:pre-injective:complete}. } \label{fig:lemma:complete-pre-injective} \end{figure} \begin{theorem}[see~\cite{Rue04}, Proposition~2.5] \label{thm:gibbs:factor-map:complete} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a complete pre-injective factor map between two strongly irreducible shifts of finite type $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$. Let $\Delta$ be a Hamiltonian on $\xSp{Y}$, and $\pi$ a probability measure on $\xSp{X}$. If $\pi$ is a Gibbs measure for~$\Phi^*\Delta$, then $\Phi\pi$ is a Gibbs measure for~$\Delta$. \end{theorem} \begin{proof} Let $0\in N\subseteq\xL$ be a neighborhood for $\Phi$. Let $0\in M\subseteq\xL$ be a neighborhood that witnesses the finite type gluing property of both $\xSp{X}$ and $\xSp{Y}$. We write $\tilde{N}\IsDef N^{-1}(N)$ and $\tilde{M}\IsDef M^{-1}(M)$. Let $y$ and $y'$ be two asymptotic configurations in $\xSp{Y}$, and set $D\IsDef\diff(y,y')$. For every configuration $x\in\Phi^{-1}[y]_{\tilde{M}(D)}$, there is a unique configuration $x'\in\Phi^{-1}[y']_{\tilde{M}(D)}$ that is asymptotic to $x$ and such that $\diff(\Phi x,\Phi x')\subseteq D$. (Namely, by the gluing property of $\xSp{Y}$, the configuration $y'_x$ that agrees with $y'$ in $\tilde{M}(D)$ and with $\Phi x$ outside $D$ is in $\xSp{Y}$. Since $\Phi$ is a complete pre-injective factor map, there is a unique configuration $x'$ that is asymptotic to $x$ and $\Phi x'=y'_x$.) The relation $x\mapsto x'$ is a one-to-one correspondence. By Lemma~\ref{lem:pre-injective:complete}, there is a large enough finite set $E\subseteq\xL$ such that for every $x\in\Phi^{-1}[y]_{\tilde{M}(D)}$, it holds $\diff(x,x')\subseteq E$. Consider a large finite set $\hat{D}\subseteq\xL$ and another finite set $\hat{E}\subseteq\xL$ that is much larger than $\hat{D}$. (More precisely, we need $\hat{D}\supseteq \tilde{M}(D)$ and $\hat{E}\supseteq N(\hat{D})\cup\tilde{N}(E)$.) Let $P_y$ denote the set of patterns $p\in L_{\hat{E}}(\xSp{X})$ such that $\Phi[p]_{\hat{E}}\subseteq [y]_{\hat{D}}$. Then, $\Phi^{-1}[y]_{\hat{D}}=\bigcup_{p\in P_y} [p]_{\hat{E}}$ (provided $\hat{E}\supseteq N(\hat{D})$). Let $\event{I}\subseteq\xSp{X}$ be a finite set consisting of one representative from each cylinder $[p]_{\hat{E}}$, for $p\in A$. Then, $\Phi^{-1}[y]_{\hat{D}}=\bigcup_{x\in\event{I}}[x]_{\hat{E}}$. Moreover, $\Phi^{-1}[y']_{\hat{D}}=\bigcup_{x\in\event{I}}[x']_{\hat{E}}$ (provided $\hat{E}\supseteq\tilde{N}(E)$). Since the terms in each of the two latter unions are disjoint, we have \begin{align} (\Phi\pi)[y]_{\hat{D}} = \sum_{x\in\event{I}} \pi([x]_{\hat{E}}) \qquad\quad\text{and}\qquad\quad (\Phi\pi)[y']_{\hat{D}} = \sum_{x\in\event{I}} \pi([x']_{\hat{E}}) \;. \end{align} For each $x\in \Phi^{-1}[y]_{\hat{D}}$, we have \begin{align} \frac{\pi([x']_{\hat{E}})}{\pi([x]_{\hat{E}})} - \xe^{-\Delta(y,y')} &= \left(\frac{\pi([x']_{\hat{E}})}{\pi([x]_{\hat{E}})} - \xe^{-(\Phi^*\Delta)(x,x')}\right) + \left(\xe^{-\Delta(\Phi x,\Phi x')} - \xe^{-\Delta(y,y')}\right) \;. \end{align} Let us denote the first term on the righthand side by $\delta_{\hat{E}}(x)$ and the second term by $\gamma_{\hat{D}}(x)$. Note that, since $\pi$ is a Gibbs measure for $\Phi^*\Delta$ and $\diff(x,x')\subseteq E$, $\delta_{\hat{E}}(x)\to 0$ uniformly over $\Phi^{-1}[y]_{\hat{D}}$ as $\hat{E}\nearrow\xL$. Note also that, by the continuity property of $\Delta$, $\gamma_{\hat{D}}(x)\to 0$ uniformly over $\Phi^{-1}[y]_{\hat{D}}$ as $\hat{D}\nearrow\xL$. We can now write \begin{align} \frac{(\Phi\pi)[y']_{\hat{D}}}{(\Phi\pi)[y]_{\hat{D}}} = \frac{\displaystyle{% \sum_{x\in\event{I}} \pi([x']_{\hat{E}}) }}{\displaystyle{% \sum_{x\in\event{I}} \pi([x]_{\hat{E}}) }} &= \frac{\displaystyle{% \sum_{x\in\event{I}} \left( \xe^{-\Delta(y,y')} + \delta_{\hat{E}}(x) + \gamma_{\hat{D}}(x) \right)\; \pi([x]_{\hat{E}}) }}{\displaystyle{% \sum_{x\in\event{I}} \pi([x]_{\hat{E}}) }} \\ &= \xe^{-\Delta(y,y')} + \frac{\displaystyle{% \sum_{x\in\event{I}} \left(\delta_{\hat{E}}(x) + \gamma_{\hat{D}}(x)\right)\; \pi([x]_{\hat{E}}) }}{\displaystyle{% \sum_{x\in\event{I}} \pi([x]_{\hat{E}}) }} \;. \end{align} Consider a small number $\varepsilon>0$. If $\hat{D}$ is sufficiently large, we have $\abs{\gamma_{\hat{D}}}<\varepsilon/2$. Moreover, after choosing $\hat{D}$, we can choose $\hat{E}$ large enough so that $\abs{\delta_{\hat{E}}}<\varepsilon/2$. Therefore, for $\hat{D}$ sufficiently large we get \begin{align} \abs{\frac{(\Phi\pi)[y']_{\hat{D}}}{(\Phi\pi)[y]_{\hat{D}}} - \xe^{-\Delta(y,y)}} &\leq \varepsilon. \end{align} It follows that \begin{align} \abs{\frac{(\Phi\pi)[y']_{\hat{D}}}{(\Phi\pi)[y]_{\hat{D}}} - \xe^{-\Delta(y,y)}} &\to 0 \end{align} as $\hat{D}\nearrow \xL$. Since this is valid for every two asymptotic configurations $y,y'\in\xSp{Y}$, we conclude that $\Phi\pi$ is a Gibbs measure for $\Delta$. \end{proof} \begin{corollary} \label{cor:gibbs:conjugacy} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a conjugacy between two strongly irreducible shifts of finite type $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$. Let $\Delta$ be a Hamiltonian on $\xSp{Y}$, and $\pi$ a probability measure on $\xSp{X}$. Then, $\pi$ is a Gibbs measure for~$\Phi^*\Delta$ if and only if $\Phi\pi$ is a Gibbs measure for~$\Delta$. \end{corollary} \subsection{The Image of a Gibbs Measure} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map between two strongly irreducible shifts of finite type. According to Corollary~\ref{thm:equilibrium:factor-map}, a pre-image of a shift-invariant Gibbs measure under the induced map $\xSx{P}(\xSp{X},\sigma)\to\xSx{P}(\xSp{Y},\sigma)$ is again a Gibbs measure. The image of a Gibbs measure, however, does not need to be a Gibbs measure as the following example demonstrates. \begin{example}[XOR map] \label{exp:xor:non-gibbs} Let $\xSp{X}=\xSp{Y}\IsDef\{\symb{0},\symb{1}\}^{\xZ}$ be the binary full shift and $\Phi$ the so-called \emph{XOR map}, defined by $(\Phi x)(i)\IsDef x(i)+x(i+1) \pmod{2}$. Let $\pi$ be the shift-invariant Bernoulli measure on $\xSp{X}$ with marginals $\symb{1}\mapsto p$ and $\symb{0}\mapsto 1-p$, where $0< p< 1$. This is a Gibbs measure for the Hamiltonian $\Delta_f$, where $f:\xSp{X}\to\xR$ is the single-site observable defined by $f(x)\IsDef-\log p$ if $x(0)=\symb{1}$ and $f(x)\IsDef-\log(1-p)$ if $x(0)=\symb{0}$. We claim that unless $p=\frac{1}{2}$, $\Phi\pi$ is not a regular Gibbs measure (i.e., a Gibbs measure for a Hamiltonian generated by an observable with summable variations). Suppose, on the contrary, that $p\neq\frac{1}{2}$ and $\Phi\pi$ is a Gibbs measure for $\Delta_g$ for some $g\in SV(\xSp{X})$. Then $\pi$ is also an equilibrium measure for $g\xO\Phi$ (Corollary~\ref{thm:equilibrium:factor-map}), implying that $f$ and $g\xO\Phi$ are physically equivalent (Proposition~\ref{prop:equilibrium:equivalence}). Consider the two uniform configurations $\underline{\symb{0}}$ and $\underline{\symb{1}}$, where $\underline{\symb{0}}(i)\IsDef \symb{0}$ and $\underline{\symb{1}}(i)\IsDef \symb{1}$ for every $i\in\xZ$. We have $f(\underline{\symb{0}})=-\log p\neq -\log(1-p)=f(\underline{\symb{1}})$, whereas $g\xO\Phi(\underline{\symb{0}})=g\xO\Phi(\underline{\symb{1}})$. If $\delta_{\underline{\symb{0}}}$ and $\delta_{\underline{\symb{1}}}$ are, respectively, the probability measures concentrated on $\underline{\symb{0}}$ and $\underline{\symb{1}}$, we get that $\delta_{\underline{\symb{0}}}(f)-\delta_{\underline{\symb{0}}}(g\xO\Phi) \neq \delta_{\underline{\symb{1}}}(f)-\delta_{\underline{\symb{1}}}(g\xO\Phi)$. This is a contradiction with the physical equivalence of $f$ and $g\xO\Phi$, because $\delta_{\underline{\symb{0}}},\delta_{\underline{\symb{1}}}\in\xSx{P}(\xSp{X},\sigma)$ (Proposition~\ref{prop:physical-equivalence:continuous}). In fact, the same argument shows that none of the $n$-fold iterations $\Phi^n\pi$ are regular Gibbs measures, because $\Phi^n(\underline{\symb{0}})=\Phi^n(\underline{\symb{1}})$ for every $n\geq 1$. On the other hand, it has been shown~\cite{Miy79,Lin84}, that $\Phi^n\pi$ converges in density to the uniform Bernoulli measure, which is a Gibbs measure and is invariant under $\Phi$. The question of approach to equilibrium will be discussed in Section~\ref{sec:ca:randomization}.% \hfill$\ocircle$ \end{example} The latter example was first suggested by van~den~Berg (see~\cite{LorMaeVan98}, Section~3.2) as an example of a measure that is \emph{strongly} non-Gibbsian, in the sense that attempting to define a Hamiltonian for it via~(\ref{eq:gibbs:def}) would lead to a function $\Delta$ for which the continuity property fails \emph{everywhere}. The question of when a measure is Gibbsian and the study of the symptoms of being non-Gibbsian is an active area of research as non-Gibbsianness sets boundaries on the applicability of the so-called renormalization group technique in statistical mechanics (see e.g.~\cite{EntFerSok93,Fer06}). The observation in Example~\ref{exp:xor:non-gibbs} can be generalized as follows. \begin{proposition} \label{prop:gibbs-non-gibbs} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type and $\pi\in\xSx{P}(\xSp{X},\sigma)$ a Gibbs measure for a Hamiltonian $\Delta_f$, where $f\in SV(\xSp{X})$. Suppose that $\Phi:\xSp{X}\to\xSp{Y}$ is a pre-injective factor map from $(\xSp{X},\sigma)$ onto another shift of finite type $(\xSp{Y},\sigma)$. A necessary condition for $\Phi\pi$ to be a regular Gibbs measure is that for every two measures $\mu_1,\mu_2\in\xSx{P}(\xSp{X},\sigma)$ with $\mu_1(f)\neq\mu_2(f)$ it holds $\Phi\mu_1\neq\Phi\mu_2$. \end{proposition} \begin{example}[XOR map; Example~\ref{exp:xor:non-gibbs} continued] \label{exp:xor:non-gibbs:contd} The argument of Example~\ref{exp:xor:non-gibbs} can be stretched to show that the iterations of the XOR map turn every Gibbs measure other than the uniform Bernoulli measure eventually to a non-Gibbs measure. More specifically, for every observable $f\in SV(\xSp{X})$ that is not physically equivalent to~$0$ and every shift-invariant Gibbs measure $\pi$ for $\Delta_f$, there is an integer $n_0\geq 1$ such that for any $n\geq n_0$, the measure $\Phi^n\pi$ is not a regular Gibbs measure. This is a consequence of the self-similar behaviour of the XOR map. Namely, the map $\Phi$ satisfies $(\Phi^{2^k}x)(i) = x(i) + x(i+2^k) \pmod{2}$ for every $i\in\xZ$ and every $k\geq 1$. If $f$ is not physically equivalent to $0$, two periodic configurations $x,y\in\xSp{X}$ with common period $2^k$ can be found such that $2^{-k}\sum_{i=0}^{2^k-1}f(\sigma^i x)\neq 2^{-k}\sum_{i=0}^{2^k-1}f(\sigma^i y)$. If $\mu_x$ and $\mu_y$ denote, respectively, the shift-invariant measures concentrated at the shift orbits of $x$ and $y$, we obtain that $\mu_x(f)\neq \mu_y(f)$. Nevertheless, $\Phi^n x=\Phi^n y=\underline{\symb{0}}$ for all $n\geq 2^k$, implying that $\Phi^n \mu_x = \Phi^n \mu_y = \delta_{\underline{\symb{0}}}$. Therefore, according to Proposition~\ref{prop:gibbs-non-gibbs}, the measure $\Phi^n\pi$ cannot be a regular Gibbs measure. \hfill$\ocircle$ \end{example} With the interpretation of the shift-ergodic measures as the macroscopic states (see the \hyperref[sec:intro]{Introduction}), the above proposition reads as follows: a sufficient condition for the non-Gibbsianness of $\Phi\pi$ is that there are two macroscopic states that are distinguishable by the density of $f$ and are mapped to the same state by $\Phi$. If the induced map $\Phi:\xSx{P}(\xSp{X},\sigma)\to\xSx{P}(\xSp{Y},\sigma)$ is not one-to-one, then there are Gibbs measures (even Markov measures) whose images are not Gibbs. For, suppose $\mu_1,\mu_2\in\xSx{P}(\xSp{X},\sigma)$ are distinct measures with $\Phi\mu_1=\Phi\mu_2$. Then, there is a local observable $f\in K(\xSp{X})$ such that $\mu_1(f)\neq\mu_2(f)$. Every shift-invariant Gibbs measure for $\Delta_f$ is mapped by $\Phi$ to a measure that is not regular Gibbs. \begin{question} Let $\Phi:\xSp{X}\to\xSp{Y}$ be a pre-injective factor map between two strongly irreducible shifts of finite type $(\xSp{X},\sigma)$ and $(\xSp{Y},\sigma)$, and suppose that the induced map $\Phi:\xSx{P}(\xSp{X},\sigma)\to\xSx{P}(\xSp{Y},\sigma)$ is injective. Does $\Phi$ map every (regular, shift-invariant) Gibbs measure to a Gibbs measure? \end{question} \section{Cellular Automata} \label{sec:ca} \subsection{Conservation Laws} \label{sec:ca:conservation-laws} Let $\Phi:\xSp{X}\to\xSp{X}$ be a cellular automaton on a strongly irreducible shift of finite type $(\xSp{X},\sigma)$. We say that $\Phi$ \emph{conserves} (the energy-like quantity formalized by) a Hamiltonian $\Delta$ if $\Delta(\Phi x,\Phi y)=\Delta(x,y)$ for every two asymptotic configurations $x,y\in\xSp{X}$. If $\Delta=\Delta_f$ is the Hamiltonian generated by a local observable $f\in K(\xSp{X})$, then we may also say that $\Phi$ conserves $f$ (in the aggregate). More generally, we say that a continuous observable $f\in C(\xSp{X})$ is conserved by $\Phi$ if $f$ and $f\xO\Phi$ are physically equivalent. According to Proposition~\ref{prop:physical-equivalence:continuous}, this is equivalent to the existence of a constant $c\in\xR$ such that $(\Phi\pi)(f)=\pi(f)+c$ for every shift-invariant probability measure $\pi$. However, in this case $c$ is always $0$. \begin{proposition} \label{prop:conservation:char} Let $\Phi:\xSp{X}\to\xSp{X}$ be a cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$. A continuous observable $f\in C(\xSp{X})$ is conserved by $\Phi$ if and only if $(\Phi\pi)(f)=\pi(f)$ for every probability measure $\pi\in\xSx{P}(\xSp{X},\sigma)$. \end{proposition} \begin{proof} Let $c\in\xR$ be such that $(\Phi\pi)(f)=\pi(f)+c$ for every $\pi\in\xSx{P}(\xSp{X},\sigma)$. Then, for every $n>0$, $(\Phi^n\pi)(f)=\pi(f)+nc$. However, every continuous function on a compact space is bounded. Therefore, $c=0$. \end{proof} If an observable $f$ is conserved by a cellular automaton $\Phi$, we say that $f$ is bound by a \emph{conservation law} under $\Phi$. There is also a concept of \emph{local} conservation law. Let $D_0$ be a finite generating set for the group $\xL=\xZ^d$. Suppose that $f\in C(\xSp{X})$ is an observable that is conserved by a cellular automaton $\Phi:\xSp{X}\to\xSp{X}$. By Proposition~\ref{prop:conservation:char} and Lemma~\ref{lem:annihilators}, this means that $f\xO\Phi-f\in C(\xSp{X},\sigma)$, that is \begin{align} f\xO\Phi-f &= \lim_{n\to\infty} \sum_{i\in D_0} (h_i^{(n)}\xO\sigma^i - h_i^{(n)}) \end{align} for some $h_i^{(n)}\in K(\xSp{X})$ (for $i\in D_0$ and $n=0,1,2,\ldots$). In other words, for every configuration $x\in\xSp{X}$ it holds \begin{align} f(\Phi x) &= f(x) + \lim_{n\to\infty}\left[ \sum_{i\in D_0} h_i^{(n)}(\sigma^i x) - \sum_{i\in D_0} h_i^{(n)}(x) \right] \;. \end{align} If furthermore $f\xO\Phi-f\in K(\xSp{X},\sigma)\subseteq C(\xSp{X},\sigma)$ (where $K(\xSp{X},\sigma)$ is defined as in~(\ref{eq:def:local-coboundaries})), then we have the more intuitive equation \begin{align} \label{eq:cons-law:local} f(\Phi x) &= f(x) + \sum_{i\in D_0} h_i(\sigma^i x) - \sum_{i\in D_0} h_i(x) \end{align} for some $h_i\in K(\xSp{X})$. In this case, we say that $f$ is \emph{locally conserved} by $\Phi$ (or satisfies a \emph{local conservation law} under $\Phi$). The value $h_i(\sigma^k x)$ is then interpreted as the \emph{flow} (of the energy-like quantity captured by $f$) from site $k$ to site $k-i$. The latter equation is a \emph{continuity equation}, stating that at each site $k$, the changes in the observed quantity after one step should balance with the incoming and the outgoing flows. If $\xSp{X}$ is a full shift, it is known that every conserved local observable is locally conserved. The proof is similar to that of Proposition~\ref{prop:Hamiltonian:observable:local}. Local conservation laws enjoy a somewhat symmetric relationship with time and space. Namely, an observable $f\in K(\xSp{X})$ is locally conserved by $\Phi$ if and only if the observable $\alpha\IsDef f\xO\Phi-f$ is in $K(\xSp{X},\Phi)\cap K(\xSp{X},\sigma)$. Moreover, to every observable $\alpha\in K(\xSp{X},\Phi)\cap K(\xSp{X},\sigma)$, there corresponds at least one observable $f\in K(\xSp{X})$ such that $\alpha=f\xO\Phi-f$ and $f$ is locally conserved by $\Phi$. In general, there might be several observables $f$ with the latter property. If $\alpha=f\xO\Phi-f=f'\xO\Phi-f'$ for two observables $f,f'\in K(\xSp{X})$, then $(f-f')=(f-f')\xO\Phi$; that is, $f-f'$ is \emph{invariant} under $\Phi$. Every constant observable is invariant under any cellular automaton. The following is an example of a cellular automaton with non-constant invariant local observables. \begin{example}[Invariant observables] \label{exp:almost-equicontinuous} Let $\Phi:\{\symb{0},\symb{1},\symb{2}\}^\xZ\to\{\symb{0},\symb{1},\symb{2}\}^\xZ$ be the cellular automaton with \begin{align} (\Phi x)(i) &\IsDef \begin{cases} \symb{2} & \text{if $x(i)=\symb{2}$,}\\ x(i) + x(i+1) \pmod{2}\qquad{} & \text{otherwise} \end{cases} \end{align} (see Figure~\ref{fig:almost-equicontinuous}). The observable $f:\{\symb{0},\symb{1},\symb{2}\}^\xZ\to\xR$ defined by $f(x)\IsDef 1$ if $x(i)=\symb{2}$ and $f(x)\IsDef 0$ otherwise is obviously invariant. The Hamiltonian $\Delta_f$ counts the number of occurrences of symbol~$\symb{2}$ and is conserved by $\Phi$. In fact, there are infinitely many linearly independent, physically non-equivalent observables that are invariant under $\Phi$. Namely, the relative position of the occurrences of~$\symb{2}$ remain unchanged, and hence, for any finite set $D\subseteq\xZ$, the logical conjunction of $f\xO\sigma^i$ for $i\in D$ is invariant. It follows that $\Phi$ has infinitely many distinct (and linearly independent) conservation laws. Such abundance of conservation laws is common among all cellular automata having non-constant invariant local observables (see Lemma~2 of~\cite{ForKarTaa11}), and has been suggested as the reason behind the ``non-physical'' behavior in these cellular automata (see e.g.~\cite{Tak87}). Every surjective equicontinuous cellular automaton is periodic~\cite{BlaTis00,Gam06} and hence has non-constant invariant local observables. It follows that every surjective cellular automaton that has a non-trivial equicontinuous cellular automaton as factor has non-constant invariant local observables and an infinity of linearly independent conservation laws. \hfill$\ocircle$ \end{example} \begin{question} Does every surjective cellular automaton with equicontinuous points have non-constant local observables? \end{question} \begin{figure} \begin{center} \includegraphics[width=0.6\textwidth]{figures/aequi-40x16.pdf} \end{center} \caption{% A cellular automaton with infinitely many distinct conservation laws (see Example~\ref{exp:almost-equicontinuous}). Black square represents state $\symb{2}$ and black dot represents $\symb{1}$. Time goes downward. } \label{fig:almost-equicontinuous} \end{figure} Every cellular automaton $\Phi:\xSp{X}\to\xSp{X}$ conserves the trivial Hamiltonian $\Delta\equiv 0$ on $\xSp{X}$. Furthermore, every observable $f\in C(\xSp{X})$ that is physically equivalent to $0$ (i.e., $f-c\in C(\xSp{X},\sigma)$ for some $c\in\xR$) is \emph{trivially} conserved by $\Phi$. Likewise, a local observable $f\in K(\xSp{X})$ is \emph{trivially} locally conserved by $\Phi$ if $f-c\in K(\xSp{X},\sigma)$ for some $c\in\xR$. We shall say that two local observables $f,g\in K(\xSp{X})$ are \emph{locally physically equivalent} if $f-g-c\in K(\xSp{X},\sigma)$ for some $c\in\xR$. The following proposition is the analogue of Proposition~\ref{prop:conservation:char}. \begin{proposition} Let $\Phi:\xSp{X}\to\xSp{X}$ be a cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$. A local observable $f\in K(\xSp{X})$ is locally conserved by $\Phi$ if and only if $f$ and $f\xO\Phi$ are locally physically equivalent. \end{proposition} \subsection{Invariance of Gibbs Measures} \label{sec:ca:invariance} As a corollary of the results of Section~\ref{sec:entropy-preserving}, we obtain a correspondence between the conservation laws of a surjective cellular automata and its invariant Gibbs measures. It is well-known that every surjective cellular automaton over a full shift preserves the uniform Bernoulli measure (see~\cite{Hed69}, Theorem~5.4 and~\cite{MarKim76}). The invariance of the uniform Bernoulli measure is sometimes called the \emph{balance property} of (the local update rule of) the surjective cellular automata. In case of surjective cellular automata over strongly irreducible shifts of finite type, a similar property is known to hold: every measure of maximum entropy is mapped to a measure of maximum entropy (see~\cite{CovPau74}, Corollary~2.3 and~\cite{MeeSte01}, Theorems~3.3 and~3.6). The following two theorems can be seen as further generalizations of the balance property. Indeed, choosing $f\equiv 0$ in either of the two theorems implies that a surjective cellular automaton maps each measure of maximum entropy to a measure of maximum entropy. An elementary proof of Theorem~\ref{thm:conservation-invariance:1} in the special case of surjective cellular automata on one-dimensional full shifts and single-site observables was earlier presented in~\cite{KarTaa11}. \begin{theorem} \label{thm:conservation-invariance:1} Let $\Phi:\xSp{X}\to\xSp{X}$ be a surjective cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$, and let $f\in SV(\xSp{X})$ be an observable with summable variations. The following conditions are equivalent: \begin{enumerate}[ \rm a)] \item $\Phi$ conserves $f$. \item $\Phi$ maps the set $\xSx{E}_f(\xSp{X},\sigma)$ of equilibrium measures for $f$ onto itself. \item There exist a measure in $\xSx{E}_f(\xSp{X},\sigma)$ whose $\Phi$-image is also in $\xSx{E}_f(\xSp{X},\sigma)$. \end{enumerate} If $f\in C(\xSp{X})$ does not have summable variations, condition~(a) still implies the other two conditions. \end{theorem} \begin{proof}\ \\ \begin{enumerate}[ a $\Rightarrow$ b)] \item[a $\Rightarrow$ b)] Suppose that $\Phi$ conserves $f$. By Proposition~\ref{prop:equilibrium:equivalence} and Corollary~\ref{thm:equilibrium:factor-map} we have $\pi\in\xSx{E}_f(\xSp{X},\sigma)$ if and only if $\Phi\pi\in\xSx{E}_f(\xSp{X},\sigma)$. Using Lemma~\ref{lem:onto-one-to-one}, we obtain $\Phi\xSx{E}_f(\xSp{X},\sigma)=\xSx{E}_f(\xSp{X},\sigma)$. \item[b $\Rightarrow$ c)] Trivial. \item[c $\Rightarrow$ a)] Let $f$ have summable variations. Then, so does $f\xO\Phi$. Suppose that there exists a measure $\pi\in\xSx{E}_f(\xSp{X},\sigma)$ such that $\Phi\pi\in\xSx{E}_f(\xSp{X},\sigma)$. By Corollary~\ref{thm:equilibrium:factor-map}, we also have $\pi\in\xSx{E}_{f\xO\Phi}(\xSp{X},\sigma)$. Therefore, $\xSx{E}_f(\xSp{X},\sigma)\cap \xSx{E}_{f\xO\Phi}(\xSp{X},\sigma)\neq\varnothing$ and by Proposition~\ref{prop:equilibrium:equivalence}, $f$ and $f\xO\Phi$ are physically equivalent. That is, $\Phi$ conserves $f$. \end{enumerate} \end{proof} \begin{theorem} \label{thm:conservation-invariance:2} Let $\Phi:\xSp{X}\to\xSp{X}$ be a surjective cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$. Let $f\in C(\xSp{X})$ be an observable and $e\in\xR$. If $\Phi$ conserves $f$, then $\Phi$ maps $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ onto itself. \end{theorem} From Theorems~\ref{thm:conservation-invariance:1} and~\ref{thm:conservation-invariance:2} it follows that each of the (convex and compact) sets $\xSx{E}_f(\xSp{X},\sigma)$ and $\xSx{E}_{\langle f\rangle=e}(\xSp{X},\sigma)$ contains an invariant measure for $\Phi$, provided that $\Phi$ conserves $f$. However, following the common reasoning of statistical mechanics (see the \hyperref[sec:intro]{Introduction}), such an invariant measure should not be considered as a macroscopic equilibrium state unless it is shift-ergodic (see Example~\ref{exp:ising:ca:invariant-non-invariant} below). In the implication (c $\Rightarrow$ a) of Theorem~\ref{thm:conservation-invariance:1}, the set $\xSx{E}_f(\xSp{X},\sigma)$ of equilibrium measures for $f$ can be replaced by the potentially larger set $\xSx{G}_{\Delta_f}(\xSp{X})$ of Gibbs measures for $\Delta_f$. \begin{corollary} \label{cor:conservation-invariance:1:non-shift-invariant} Let $\Phi:\xSp{X}\to\xSp{X}$ be a surjective cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$, and let $f\in SV(\xSp{X})$ be an observable with summable variations. Suppose that there is a Gibbs measure for $\Delta_f$ whose $\Phi$-image is also a Gibbs measure for $\Delta_f$. Then, $\Phi$ conserves~$\Delta_f$. \end{corollary} \begin{proof} Let $\pi$ be a probability measure on $\xSp{X}$ such that $\pi,\Phi\pi\in\xSx{G}_{\Delta_f}(\xSp{X})$. Let $\xSx{H}$ denote the closed convex hull of the measures $\sigma^k\pi$ for $k\in\xL$. Then, $\xSx{H}$ is a closed, convex, shift-invariant set, and therefore, contains a shift-invariant element $\nu$. Moreover, both $\xSx{H}$ and $\Phi\xSx{H}$ are subsets of $\xSx{G}_{\Delta_f}(\xSp{X})$. In particular, $\nu,\Phi\nu\in\xSx{G}_{\Delta_f}(\xSp{X})$. Hence, $\nu,\Phi\nu\in\xSx{E}_f(\xSp{X},\sigma)$, and the claim follows from Theorem~\ref{thm:conservation-invariance:1}. \end{proof} For reversible cellular automata, Corollary~\ref{cor:gibbs:conjugacy} leads to a variant of Theorem~\ref{thm:conservation-invariance:1} concerning all (not necessarily shift-invariant) Gibbs measures. \begin{theorem} \label{thm:conservation-invariance:reversible} Let $\Phi:\xSp{X}\to\xSp{X}$ be a reversible cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$, and let $\Delta$ be a Hamiltonian on $\xSp{X}$. The following conditions are equivalent: \begin{enumerate}[ \rm a)] \item $\Phi$ conserves $\Delta$. \item A probability measure is in $\xSx{G}_\Delta(\xSp{X})$ if and only if its $\Phi$-image is in $\xSx{G}_\Delta(\xSp{X})$. \item There exists a measure in $\xSx{G}_\Delta(\xSp{X})$ whose $\Phi$-image is also in $\xSx{G}_\Delta(\xSp{X})$. \end{enumerate} \end{theorem} \begin{proof} If $\Phi$ conserves $\Delta$, we have, by definition, $\Phi^*\Delta=\Delta$, and Corollary~\ref{cor:gibbs:conjugacy} (and Lemma~\ref{lem:onto-one-to-one}) imply that $\Phi^{-1}\xSx{G}_\Delta(\xSp{X})=\xSx{G}_\Delta(\xSp{X})$. Conversely, suppose that $\pi$ is a probability measure such that $\pi,\Phi\pi\in\xSx{G}_\Delta(\xSp{X})$. Then, by Corollary~\ref{cor:gibbs:conjugacy}, $\pi\in\xSx{G}_\Delta(\xSp{X})\cap\xSx{G}_{\Phi^*\Delta}(\xSp{X})$, and it follows from the definition of a Gibbs measure that $\Phi^*\Delta=\Delta$. That is, $\Phi$ conserves $\Delta$. \end{proof} \begin{example}[Q2R cellular automaton] \label{exp:ising:ca:invariant-non-invariant} The Q2R model discussed in the \hyperref[sec:intro]{Introduction} is not, strictly speaking, a cellular automaton (with the standard definition), as it involves alternate application of two maps that do not commute with the shift. Simple tricks can however be used to turn it into a standard cellular automaton (see e.g.~\cite{TofMar90}, Section~5.2). Let $\xSp{X}\IsDef\{\symb{+},\symb{-}\}^{\xZ^2}$ be the space of spin configurations, and denote by $\Phi_{\mathsf{e}}$ the mapping $\xSp{X}\to\xSp{X}$ that updates the even sites. That is, \begin{align} (\Phi_{\mathsf{e}} x)(i) &\IsDef \begin{cases} \overline{x_i}\qquad{} & \text{if $i$ an even site and $n_i^{\symb{+}}(x)=n_i^{\symb{-}}(x)$,} \\ x_i & \text{otherwise,} \end{cases} \end{align} where the spin-flipping operation is denoted by overline, and $n_i^{\symb{+}}(x)$ (resp., $n_i^{\symb{-}}(x)$) represents the number of sites $j$ among the four immediate neighbors of $i$ such that $x(j)=\symb{+}$ (resp., $x(j)=\symb{-}$). Similarly, let $\Phi_{\mathsf{o}}$ denotes the mapping that updates the odd sites. The composition $\Phi\IsDef \Phi_{\mathsf{o}}\Phi_{\mathsf{e}}$ commutes with the shifts $\sigma^k$, for $k$ in the sub-lattice $(2\xZ)^2$, and (after a recoding) could be considered as a cellular automaton. Let $f$ denote the energy observable defined in Example~\ref{exp:ising}. For every $\beta>0$, the Hamiltonian $\Delta_{\beta f}$ is conserved by $\Phi$. Therefore, according to Theorem~\ref{thm:conservation-invariance:reversible}, the set $\xSx{G}_{\Delta_{\beta f}}(\xSp{X})$ of Gibbs measures for $\Delta_{\beta f}$ is invariant under $\Phi$. In fact, in this example, it is easy to show that $\Phi$ preserves every individual Gibbs measure in $\xSx{G}_{\Delta_{\beta f}}(\xSp{X})$. It is natural to ask whether the preservation of individual elements of $\xSx{G}_{\Delta_{\beta f}}(\xSp{X})$ holds in general. This is however not the case. When $\beta$ large enough, it is known that $\xSx{G}_{\Delta_{\beta f}}(\xSp{X})$ contains two distinct shift-ergodic measures, obtained from each other by a spin flip transformation (see Example~\ref{exp:ising-contour}). The cellular automaton $\Phi' x\IsDef \overline{\Phi x}$, which flips every spin after applying $\Phi$, conserves $\Delta_{\beta f}$ but does not preserve either of the two distinct shift-ergodic Gibbs measures for $\Delta_{\beta f}$. \hfill$\ocircle$ \end{example} \subsection{Absence of Conservation Laws} \label{sec:ca:rigidity} In light of the above connection, every statement about conservation laws in surjective cellular automata has an interpretation in terms of invariance of Gibbs measures, and vice versa. In this section, we see an example of such reinterpretation that leads to otherwise non-trivial results. Namely, proving the abscence of conservation laws in two relatively rich families of surjective and reversible cellular automata, we obtain strong constraints on the invariant measures of the cellular automata within each family. Roughly speaking, strong chaotic behavior is incompatible with the presence of conservation laws. In contrast, any surjective cellular automaton with a non-trivial equicontinuous factor has an infinity of linearly independent conservation laws (see Example~\ref{exp:almost-equicontinuous}). We say that a dynamical system $(\xSp{X},\Phi)$ is \emph{strongly transitive} if for every point $z\in\xSp{X}$, the set $\bigcup_{i=0}^\infty \Phi^{-i}z$ is dense in $\xSp{X}$. Strong transitivity is stronger than transitivity (!)\ and weaker than minimality. A dynamical system $\Phi:\xSp{X}\to\xSp{X}$ is \emph{minimal} if it has no non-trivial closed subsystems, and is \emph{transitive} if for every pair of non-empty open sets $A,B\subseteq\xSp{X}$, there is an integer $n\geq 0$ such that $A\cap \Phi^{-n}B\neq\varnothing$. In our setting (i.e., $\xSp{X}$ being compact), minimality is equivalent to the property that the only closed sets $E\subseteq\xSp{X}$ with $E\subseteq\Phi E$ are $\varnothing$ and $\xSp{X}$, which is easily seen to imply strong transitivity. However, note that cellular automata over non-trivial strongly irreducible shifts of finite type cannot be minimal. This is because every strongly irreducible shift of finite type has configurations that are periodic in at least one direction. (More specifically, for each $k\in\xL\setminus\{0\}$, there is a configuration $x$ such that $\sigma^{pk}x=x$ for some $p>0$.) Transitivity is often considered as one of the main indicators of chaos (see e.g.~\cite{Kol04,Bla09}). Every transitive cellular automaton is known to be sensitive to initial conditions (i.e., uniformly unstable)~\cite{CodMar96,Kur97}.\footnote{% In fact, transitive cellular automata are weakly mixing~\cite{Moo05}, which, in addition to sensitivity, also implies chaos in the sense of Li and Yorke (see~\cite{Kol04,Bla09}). } \begin{example}[XOR cellular automata] \label{exp:xor:ca} The $d$-dimensional \emph{XOR cellular automaton} with neighborhood $N\subseteq\xZ^d$ is defined by the map $\Phi:\{\symb{0},\symb{1}\}^{\xZ^d}\to\{\symb{0},\symb{1}\}^{\xZ^d}$, where $(\Phi x)(i)\IsDef \sum_{i\in N} x(i) \pmod{2}$. To avoid the trivial case, we assume that the neighborhood has at least two elements. Examples~\ref{exp:xor:non-gibbs} and~\ref{exp:xor:non-gibbs:contd} were about the one-dimensional XOR cellular automaton with neighborhood $\{0,1\}$. Figure~\ref{fig:xor:transpose}a depicts a sample run of the one-dimensional model with neighborhood $\{-1,1\}$. The XOR cellular automaton is strongly transitive. An argument similar to that in Example~\ref{exp:xor:non-gibbs:contd} shows that the uniform Bernoulli measure is the only regular Gibbs measure that is invariant under an XOR cellular automaton. Note, however, that there are many other (non-Gibbs) invariant measures. For example, the Dirac measure concentrated at the uniform configuration with $\symb{0}$ everywhere is invariant. So is the (atomic) measure uniformly distributed over any jointly periodic orbit (i.e., a finite orbit of $(\sigma,\Phi)$). In fact, much more is known about the invariant measures of the XOR cellular automata, with a strong indication that the uniform Bernoulli measure is the only ``state of macroscopic equilibrium''. For instance, the uniform Bernoulli measure on $\{0,1\}^\xZ$ is known to be the only shift-ergodic probability measure that is invariant and of positive entropy for the XOR cellular automaton with neighborhood $\{0,1\}$~\cite{HosMaaMar03}. Another such result states that the only measures that are strongly mixing for the shift and invariant under the XOR cellular automaton with neighborhood $\{-1,1\}$ are the uniform Bernoulli measure and the Dirac measure concentrated at the uniform configuration with $\symb{0}s$ everywhere~\cite{Miy79}. (Note that the one-dimensional Gibbs measures are all strongly mixing.) Similar results have been obtained for broad classes of cellular automata with algebraic structure (e.g.~\cite{Piv05,Sab07,Sob08}). See~\cite{Piv09} for a survey. \hfill$\ocircle$ \end{example} The following theorem is a slight generalization of Theorem~5 in~\cite{ForKarTaa11}. \begin{theorem} \label{thm:strongly-transitive-CA:no-cons-law} Let $\Phi:\xSp{X}\to\xSp{X}$ be a strongly transitive cellular automaton over a shift of finite type $(\xSp{X},\sigma)$. Then, $\Phi$ does not conserve any non-trivial Hamiltonian. \end{theorem} \begin{proof} Let $0\in M\subseteq\xL$ be a finite window that witnesses the finite type gluing property of~$\xSp{X}$. Let $\Delta$ be a non-trivial Hamiltonian on $\xSp{X}$, and suppose there exist two asymptotic configurations $u$ and $v$ such that $\varepsilon\IsDef\Delta(u,v)>0$. By the continuity property of $\Delta$, there is a finite set $D\supseteq M(M^{-1}(\diff(u,v)))$ such that for every two asymptotic configurations $u'\in [u]_D$ and $v'\in [v]_D$ with $\diff(u',v')=\diff(u,v)$, $\Delta(u',v')\geq\varepsilon/2>0$. Let $z$ be a ground configuration for $\Delta$ (see Proposition~\ref{prop:Hamiltonian:ground}). Since $\Phi$ is strongly transitive, there is a configuration $x\in[v]_D$ and a time $t\geq 0$ such that $\Phi^t x=z$. Construct a configuration $y\in\xSp{X}$ that agrees with $u$ on $D$ and with $x$ outside $\diff(u,v)$. In particular, $y\in[u]_D$. Then, $\Delta(y,x)\geq\varepsilon/2$, whereas $\Delta(\Phi^t y,\Phi^t x)\leq 0$. Therefore, $\Delta$ is not conserved by $\Phi$. \end{proof} \begin{corollary} \label{cor:ca:strongly-transitive:rigidity} Let $\Phi:\xSp{X}\to\xSp{X}$ be a strongly transitive cellular automaton over a strongly irreducible shift of finite type $(\xSp{X},\sigma)$. Then, $\Phi$ does not preserve any regular Gibbs measure other than the Gibbs measures for the trivial Hamiltonian. \end{corollary} A special case of the above corollary (for the permutive cellular automata and Bernoulli measures) is also proved in~\cite{BanChaChe11} (Corollary 3.6). Let us recall that the \emph{shift-invariant} Gibbs measures for the trivial Hamiltonian on $\xSp{X}$ coincide with the measures of maximum entropy for $(\xSp{X},\sigma)$ (Theorem~\ref{thm:equilibrium:Gibbs}). Therefore, according to Corollary~\ref{cor:ca:strongly-transitive:rigidity}, if $\Phi$ is strongly transitive, the measures of maximum entropy for $(\xSp{X},\sigma)$ are the only candidates for Gibbs measures that are invariant under both~$\sigma$ and~$\Phi$. Since the set of measures with maximum entropy for $(\xSp{X},\sigma)$ is closed and convex, and is preserved under $\Phi$, it follows that at least one measure with maximum entropy is invariant under~$\Phi$. However, this measure does not need to be ergodic for the shift. Next, we are going to introduce a class of one-dimensional \emph{reversible} cellular automata with no local conservation law. The proof will be via reduction to Theorem~\ref{thm:strongly-transitive-CA:no-cons-law}. Note that reversible cellular automata over non-trivial strongly irreducible shifts of finite type cannot be strongly transitive: the inverse of a strongly transitive system is minimal, and as mentioned above, cellular automata over non-trivial strongly irreducible shifts of finite type cannot be minimal. \begin{example}[Transpose of XOR] \label{exp:xor:transpose} Figure~\ref{fig:xor:transpose}b depicts a sample space-time diagram of the reversible cellular automaton $\Phi$ on $(\{\symb{0},\symb{1}\}\times\{\symb{0},\symb{1}\})^\xZ$ with neighbourhood $\{0,1\}$ and local rule $((a,b),(c,d))\mapsto (b,a+d)$, where the addition is modulo~$2$. Observe that rotating a space-time diagram of $\Phi$ by $90$ degrees, we obtain what is essentially a space-time diagram of the XOR cellular automaton with neighbourhood $\{-1,1\}$ (see Figure~\ref{fig:xor:transpose}a and Example~\ref{exp:xor:ca}). As in Example~3 of~\cite{ForKarTaa11}, it is possible to show that $\Phi$ has no non-trivial finite-range conservation law. Below, we shall present an alternative proof (using its connection with the XOR cellular automaton) that covers a large class of similar reversible cellular automata. \hfill$\ocircle$ \end{example} \begin{figure} \begin{center} \begin{tabular}{ccc} \begin{minipage}[c]{0.35\textwidth} \centering \includegraphics[width=0.95\textwidth]{figures/xor-40x40.pdf} \end{minipage} & & \begin{minipage}[c]{0.35\textwidth} \centering \includegraphics[width=0.95\textwidth]{figures/xort-40x40.pdf} \end{minipage} \medskip\\ (a) && (b) \end{tabular} \end{center} \caption{% Sample runs of (a) the one-dimensional XOR cellular automaton with neighborhood $\{-1,1\}$ (see Example~\ref{exp:xor:ca}) and (b) its transpose (Example~\ref{exp:xor:transpose}). Time goes downward. } \label{fig:xor:transpose} \end{figure} We shall say that two surjective one-dimensional cellular automata are \emph{transpose} of each other if the bi-infinite space-time diagrams of each is obtained (up to a conjugacy) from the bi-infinite space-time diagrams of the other by swapping the role of space and time. To be more specific, let $\Phi:\xSp{X}\to\xSp{X}$ be a surjective cellular automaton on a one-dimensional mixing shift space of finite type $\xSp{X}\subseteq S^\xZ$. Define the continuous map $\Theta:S^{\xZ\times\xZ}\to S^\xZ$ where $(\Theta z)(i)\IsDef z(i,0)$, and let $\tilde{\xSp{X}}$ be the two-dimensional shift space formed by all configurations $z\in S^{\xZ\times\xZ}$ such that \begin{align} \ldots \,\xrightarrow[]{\Phi}\, \Theta\sigma^{(0,-1)}z \,\xrightarrow[]{\Phi}\, \Theta z \,\xrightarrow[]{\Phi}\, \Theta\sigma^{(0,1)} z \,\xrightarrow[]{\Phi}\, \ldots \end{align} is a bi-infinite orbit of $\Phi$, that is $\Theta\sigma^{(0,k+1)} z = \Phi\Theta\sigma^{(0,k)} z$ for each $k\in\xZ$. Set $\xVer\IsDef(0,1)$ and $\xHor\IsDef(1,0)$. The dynamical system $(\tilde{\xSp{X}},\sigma^\xVer,\sigma^\xHor)$ (together with the map $\Theta$) is the \emph{natural extension} of $(\xSp{X},\Phi,\sigma)$. Now, let $\Psi:\xSp{Y}\to\xSp{Y}$ be another surjective cellular automaton on a one-dimensional mixing shift space of finite type $\xSp{Y}\subseteq T^\xZ$. We say $\Psi$ is a \emph{transpose} of $\Phi$ if its natural extension is conjugate to $(\tilde{\xSp{X}},\sigma^\xHor,\sigma^\xVer)$. The transpose of $\Phi$ (if it exists) is unique only up to conjugacy. When there is no danger of confusion, we denote any representative of the transpose conjugacy class by~$\Phi^\intercal$. \begin{proposition} \label{prop:transpose:mixing} A surjective cellular automaton on a one-dimensional mixing shift of finite type is mixing provided it has a transpose (acting on a mixing shift of finite type). \end{proposition} \begin{proof} A dynamical system is mixing if and only if its natural extension is mixing. \end{proof} Obviously, not every cellular automaton has a transpose. A class of cellular automata that do have transposes is the class of those that are positively expansive. A dynamical system $(\xSp{X},\Phi)$ is \emph{positively expansive} if there exists a real number $\varepsilon>0$ such that for every two distinct points $x,y\in \xSp{X}$, there is a time $t\geq 0$ such that $\Phi^t x$ and $\Phi^t y$ have distance at least $\varepsilon$. If $(\xSp{X},\sigma)$ is a mixing shift of finite type and $\Phi:\xSp{X}\to\xSp{X}$ is a positively expansive cellular automaton, then $\Phi$ is surjective, and it is known that a transpose of $\Phi$ exists and is a reversible cellular automaton on a mixing shift of finite type (see~\cite{Kur03b}, Section~5.5\footnote{% The proof in~\cite{Kur03b} is presented for the case that $(\xSp{X},\sigma)$ is a full shift, but the same proof, with slight adaptation, works for any arbitrary mixing shift of finite type. To prove the openness of $\Phi$, see~\cite{Nas83}, Theorems~6.3 and~6.4, and note that $\Phi$ is both left- and right-closing. }). If, furthermore, $(\xSp{X},\sigma)$ is a full shift, then the transpose of $\Phi$ also acts on a full shift (see~\cite{Nas95}, Theorem~3.12). \begin{proposition} Every positively expansive cellular automaton on a one-dimensional mixing shift of finite type is strongly transitive. \end{proposition} \begin{proof} Any continuous map $\Phi:\xSp{X}\to\xSp{X}$ on a compact metric space that is transitive, open, and positively expansive is strongly transitive~\cite{Kam02}. Every positively expansive cellular automaton on a mixing shift of finite type is itself mixing (see the above paragraph) and open (see~\cite{Kur03b}, Theorem~5.45). Alternatively, every positively expansive cellular automaton on a mixing shift of finite type is conjugate to a mixing \emph{one-sided} shift of finite type (see~\cite{Kur03b}, Theorem~5.49), and hence strongly transitive. \end{proof} The local conservation laws of a cellular automaton and its transpose are in one-to-one correspondence. \begin{theorem} \label{thm:transpose:cons-law} Let $\Phi:\xSp{X}\to\xSp{X}$ and $\Phi^\intercal:\xSp{X}^\intercal\to\xSp{X}^\intercal$ be surjective cellular automata over one-dimensional mixing shifts of finite type $\xSp{X}$ and $\xSp{X}^\intercal$, and suppose that $\Phi$ and $\Phi^\intercal$ are transpose of each other. There is a one-to-one correspondence (up to local physical equivalence) between the observables $f\in K(\xSp{X})$ that are locally conserved by $\Phi$ and the observables $f^\intercal\in K(\xSp{X}^\intercal)$ that are locally conserved by $\Phi^\intercal$. Moreover, $f$ is locally physically equivalent to $0$ if and only if $f^\intercal$ is so. \end{theorem} \begin{proof} Recall that an observable $f\in K(\xSp{X})$ is locally conserved by $\Phi$ if and only if it satisfies the continuity equation \begin{align} f\xO\Phi &= f + g\xO\sigma - g \end{align} for some observable $g\in K(\xSp{X})$, where the terms $g\xO\sigma$ and $g$ are interpreted, respectively, as the flow pouring into a site from its right neighbour and the flow leaving that site towards its left neighbour. This equation may alternatively be written as \begin{align} g\xO\sigma &= g + f\xO\Phi - f \;, \end{align} which can be interpreted as the local conservation of the observable $g$ when the role of $\Phi$ and $\sigma$ are exchanged. To specify the correspondence between local conservation laws of $\Phi$ and $\Phi^\intercal$ more precisely, let $\tilde{\xSp{X}}$ be the shift space of the space-time diagrams of $\Phi$, so that $(\tilde{\xSp{X}},\sigma^\xVer,\sigma^\xHor)$ is the natural extension of $(\xSp{X},\Phi,\sigma)$, and $(\tilde{\xSp{X}},\sigma^\xHor,\sigma^\xVer)$ is the natural extension of $(\xSp{X}^\intercal,\Phi^\intercal,\sigma)$, and let $\Theta:\tilde{\xSp{X}}\to\xSp{X}$ and $\Theta^\intercal:\tilde{\xSp{X}}\to\xSp{X}^\intercal$ be, respectively, the corresponding factor maps, extracting (up to a conjugacy) the $0$th row and the $0$th column of $\tilde{\xSp{X}}$. Let us use the following notation. Suppose that local observables $f,g\in K(\xSp{X})$ and $f^\intercal,g^\intercal\in K(\xSp{X}^\intercal)$ are such that $f\xO\Theta = g^\intercal\xO\Theta^\intercal$ and $f^\intercal\xO\Theta^\intercal=g\xO\Theta$, and setting $\tilde{f}_\xVer\IsDef f\xO\Theta = g^\intercal\xO\Theta^\intercal$ and $\tilde{f}_\xHor\IsDef f^\intercal\xO\Theta^\intercal=g\xO\Theta$, it holds \begin{align} \tilde{f}_\xVer\xO\sigma^\xVer - \tilde{f}_\xVer &= \tilde{f}_\xHor\xO\sigma^\xHor - \tilde{f}_\xHor \;. \end{align} Then, we write $f_1\perp f^\intercal_1$ for any two local observables $f_1\in K(\xSp{X})$ and $f^\intercal_1\in K(\xSp{X})$ that are locally physically equivalent to $f$ and $f^\intercal$, respectively. We verify that \begin{enumerate}[i)] \item a local observable $f\in K(\xSp{X})$ is locally conserved by $\Phi$ if and only if $f\perp f^\intercal$ for some local observable $f^\intercal\in K(\xSp{X}^\intercal)$, \item the relation $\perp$ is linear, and \item $f\perp 0$ if and only if $f$ is locally physically equivalent to $0$. \end{enumerate} Note that these three statements (along with the similar statements obtained by swapping~$f$ and~$f^\intercal$) would imply that $\perp$ is a one-to-one correspondence with the desired properties. To prove the first statement, suppose that $f\perp f^\intercal$. Then, \begin{align} \tilde{f}_\xVer\xO\sigma^\xVer - \tilde{f}_\xVer &= \tilde{f}_\xHor\xO\sigma^\xHor - \tilde{f}_\xHor \;, \end{align} where $\tilde{f}_\xVer = f\xO\Theta$ and $\tilde{f}_\xHor=g\xO\Theta$ for some $g\in K(\xSp{X})$. Rewriting this equation as \begin{align} (f\xO\Phi - f)\xO\Theta &= (g\xO\sigma - g)\xO\Theta \;, \end{align} we obtain, using Lemma~\ref{lem:onto-one-to-one} and the surjectivity of $\Theta$, that $f$ is locally conserved by $\Phi$. Conversely, suppose that $\Phi$ locally conserves $f$, and let $g\in K(\xSp{X})$ be such that $f\xO\Phi - f=g\xO\sigma - g$. Therefore, \begin{align} f\xO\Theta\xO\sigma^\xVer - f\xO\Theta &= g\xO\Theta\xO\sigma^\xHor - g\xO\Theta \;. \end{align} Since $f\xO\Theta$ and $g\xO\Theta$ are local observables, there exists a finite region $D\subseteq\xZ\times\xZ$ such that $f\xO\Theta,g\xO\Theta\in K_D(\tilde{\xSp{X}})$. By the definition of natural extension, there is an integer $k>0$ such that $\xRest{z}{D}$ is uniquely and continuously determined by $\Theta^\intercal\sigma^{-k\xHor}z$ (i.e., column $-k$ of the space-time shift of $\Phi$). Hence, there exist observables $f^\intercal,g^\intercal\in K(\xSp{X}^\intercal)$ such that $f\xO\Theta\xO\sigma^{k\xHor}=g^\intercal\xO\Theta^\intercal$ and $g\xO\Theta\xO\sigma^{k\xVer}=f^\intercal\xO\Theta^\intercal$. Now, setting $\tilde{f}_\xVer\IsDef f\xO\sigma^k\xO\Theta=g^\intercal\xO\Theta^\intercal$ and $\tilde{f}_\xHor\IsDef g\xO\sigma^k\xO\Theta=f^\intercal\xO\Theta^\intercal$, we can write \begin{align} \tilde{f}_\xVer\xO\sigma^\xVer - \tilde{f}_\xVer &= \tilde{f}_\xHor\xO\sigma^\xHor - \tilde{f}_\xHor \;, \end{align} which means $f\xO\sigma^k \perp f^\intercal$. Finally, note that $f$ and $f\xO\sigma^k$ are locally physically equivalent. The linearity of $\perp$ and the fact that $0\perp 0$ are clear. It remains to show that if $f_1\in K(\xSp{X})$ is a local observable such that $f_1\perp 0$, then $f_1$ is locally physically equivalent to $0$. Suppose that $f_1\perp 0$. Then, there is an observable $f\in K(\xSp{X})$ locally physically equivalen to $f_1$, and an observable $f^\intercal\in K(\xSp{X}^\intercal)$ locally physically equivalent to $0$ such that \begin{align} f\xO\Theta\xO\sigma^\xVer - f\xO\Theta &= f^\intercal\xO\Theta^\intercal\xO\sigma^\xHor - f^\intercal\xO\Theta^\intercal \;. \end{align} Since $f^\intercal$ is locally physically equivalent to $0$, it has the form $f^\intercal=h^\intercal\xO\sigma - h^\intercal + c$ for some observable $h^\intercal\in K(\xSp{X}^\intercal)$ and some constant $c\in\xR$. Therefore, \begin{align} \label{eq:thm:transpose:1} f\xO\Theta\xO\sigma^\xVer - f\xO\Theta &= h^\intercal\xO\Theta^\intercal\xO\sigma^\xVer\xO\sigma^\xHor - h^\intercal\xO\Theta^\intercal\xO\sigma^\xVer - h^\intercal\xO\Theta^\intercal\xO\sigma^\xHor + h^\intercal\xO\Theta^\intercal \;. \end{align} Since $h^\intercal$ is a local observable, we can find, as before, an integer $l>0$ and a local observable $h\in K(\xSp{X})$ such that $h^\intercal\xO\sigma^l\xO\Theta^\intercal = h\xO\Theta$. Therefore, composing both sides of~(\ref{eq:thm:transpose:1}) with $\sigma^{l\xVer}$ leads to \begin{align} f\xO\Phi^l\xO\Theta\xO\sigma^\xVer - f\xO\Phi^l\xO\Theta &= h\xO\Theta\xO\sigma^\xVer\xO\sigma^\xHor - h\xO\Theta\xO\sigma^\xVer - h\xO\Theta\xO\sigma^\xHor + h\xO\Theta \;, \end{align} which, together with Lemma~\ref{lem:onto-one-to-one}, gives \begin{align} f\xO\Phi^l\xO\Phi - f\xO\Phi^l &= h\xO\Phi\xO\sigma - h\xO\Phi - h\xO\sigma + h \;. \end{align} The latter equation can be rewritten as \begin{align} (f\xO\Phi^l - h\xO\sigma + h)\xO\Phi &= (f\xO\Phi^l - h\xO\sigma + h) \;, \end{align} which says that $f\xO\Phi^l - h\xO\sigma + h$ is invariant under $\Phi$. On the other hand, since $(\xSp{X}^\intercal,\sigma)$ is a mixing shift, it follows from Proposition~\ref{prop:transpose:mixing} that $(\xSp{X},\Phi)$ is also mixing. As a consequence, every continuous observable that is invariant under $\Phi$ is constant. In particular, $f\xO\Phi^l - h\xO\sigma + h=c'$ for some constant $c'\in \xR$, which means $f\xO\Phi^l$ is locally physically equivalent to $0$. Since $f$ is locally conserved by $\Phi$, the observable $f\xO\Phi^l$ is also locally physically equivalent to $f$, and this completes the proof. \end{proof} \begin{corollary} \label{cor:ca:pexp-transpose:no-cons-law} Let $\Phi:\xSp{X}\to\xSp{X}$ be a reversible cellular automaton on a one-dimensional mixing shift of finite type $(\xSp{X},\sigma)$, and suppose that $\Phi$ has a positively expansive transpose. Then, $\Phi$ has no non-trivial local conservation law. \end{corollary} As mentioned in Section~\ref{sec:ca:conservation-laws}, for cellular automata on full shifts, every conserved local observable is locally conserved. \begin{corollary} \label{cor:ca:pexp-transpose:rigidity} Let $\Phi:\xSp{X}\to\xSp{X}$ be a reversible cellular automaton on a one-dimensional full shift $(\xSp{X},\sigma)$, and suppose that $\Phi$ has a positively expansive transpose. The uniform Bernoulli measure is the only finite-range Gibbs measure ($\equiv$ full-support Markov measure) that is invariant under~$\Phi$. \end{corollary} \begin{example}[Non-additive positively expansive] \label{exp:permutive:transpose} Let $\Phi:\{\symb{0},\symb{1},\symb{2}\}^\xZ\to\{\symb{0},\symb{1},\symb{2}\}^\xZ$ be the cellular automaton defined by neighborhood $N\IsDef\{-1,0,1\}$ and local update rule $\varphi:\{\symb{0},\symb{1},\symb{2}\}^N\to \{\symb{0},\symb{1},\symb{2}\}$ defined by \begin{align} \varphi(a,b,c) &\IsDef \begin{cases} a+c+\symb{1} \pmod{3}\qquad\ & \text{if $b=\symb{2}$,}\\ a+c \pmod{3} & \text{otherwise.} \end{cases} \end{align} See Figure~\ref{fig:permutive:transpose}a for a sample run. Note that the local rule is both left- and right-\emph{permutive} (i.e., $a\mapsto\varphi(a,b,c)$ and $c\mapsto\varphi(a,b,c)$ are permutations). It follows that $\Phi$ is positively expansive, and hence also strongly transitive. Therefore, according to Theorem~\ref{thm:strongly-transitive-CA:no-cons-law} and Corollary~\ref{cor:ca:strongly-transitive:rigidity}, $\Phi$ has no non-trivial conservation law and the uniform Bernoulli measure is the only regular Gibbs measure on $\{\symb{0},\symb{1},\symb{2}\}^\xZ$ that is invariant under $\Phi$. The cellular automaton $\Phi$ has a transpose $\Phi^\intercal:(\{\symb{0},\symb{1},\symb{2}\}\times\{\symb{0},\symb{1},\symb{2}\})^\xZ \to (\{\symb{0},\symb{1},\symb{2}\}\times\{\symb{0},\symb{1},\symb{2}\})^\xZ$ defined with neighborhood $\{0,1\}$ and local rule \begin{align} ((a,b), (a',b')) &\mapsto \begin{cases} (b, b'-a-\symb{1})\qquad\ &\text{if $b=\symb{2}$,}\\ (b, b'-a) &\text{otherwise,}\\ \end{cases} \end{align} where the subtractions are modulo~$3$ (see Figure~\ref{fig:permutive:transpose}b). This is a reversible cellular automaton. It follows from Corollaries~\ref{cor:ca:pexp-transpose:no-cons-law} and~\ref{cor:ca:pexp-transpose:rigidity} that $\Phi^\intercal$ has no non-trivial local conservation law and no invariant full-support Markov measure other than the uniform Bernoulli measure. \hfill$\ocircle$ \end{example} \begin{figure} \begin{center} \begin{tabular}{ccc} \begin{minipage}[c]{0.35\textwidth} \centering \includegraphics[width=0.95\textwidth]{figures/biperm-40x40.pdf} \end{minipage} & & \begin{minipage}[c]{0.35\textwidth} \centering \includegraphics[width=0.95\textwidth]{figures/bipermt-40x40.pdf} \end{minipage} \medskip\\ (a) && (b) \end{tabular} \end{center} \caption{% Sample runs of (a) a one-dimensional bi-permutive cellular automaton and (b) its transpose. See Example~\ref{exp:permutive:transpose}. Time goes downward. Black circle represents~$\symb{2}$ and white circle~$\symb{1}$. } \label{fig:permutive:transpose} \end{figure} According to Corollary~\ref{cor:ca:strongly-transitive:rigidity}, if $\Phi$ is a strongly transitive cellular automaton on a full shift $\xSp{X}$, the uniform Bernoulli measure on $\xSp{X}$ is the only regular Gibbs measure that is preserved by $\Phi$. Likewise, Corollary~\ref{cor:ca:pexp-transpose:rigidity} states that for a class of one-dimensional reversible cellular automata, the uniform Bernoulli measure is the only invariant full-support Markov measure. Note that even with these constraints, a cellular automaton in either of these two classes still has a large collection of other invariant measures. For example, for every $d$ linearly independent vectors $k_1,k_2,\ldots,k_d\in\xZ^d$, the set of $d$-dimensional spatially periodic configurations having $k_i$ as periods (i.e., $\{x: \sigma^{k_i} x=x \text{ for $i=1,2,\ldots,d$}\}$) is finite and invariant under any cellular automaton, and therefore any cellular automaton has an (atomic) invariant measure supported at such a set. Nevertheless, if we restrict our attention to sufficiently ``smooth'' measures, the uniform Bernoulli measure becomes the ``unique'' invariant measure for a cellular automaton in either of the above classes.\footnote{ The term ``smoothness'' here refers to the continuity of the conditional probabilities $\pi([p]_D\,|\,\family{F}_D)(z)$ for Gibbs measures (which is a defining proeprty). Unfortunately, Corollary~\ref{cor:ca:strongly-transitive:rigidity} restricts only the invariance of \emph{regular} Gibbs measures (see Section~\ref{sec:hamiltonian-gibbs}). We do not know if every Hamiltonian is generated by an observable with summable variations. However, see~\cite{Sul73} in this direction. } In this sense, Corollaries~\ref{cor:ca:strongly-transitive:rigidity} and~\ref{cor:ca:pexp-transpose:rigidity} may be interpreted as weak indications of ``absence of phase transition'' for cellular automata in the two classes in question. \begin{question} Let $(\xSp{X},\sigma)$ be a strongly irreducible shift of finite type. Which shift-ergodic measures can be invariant under a strongly transitive cellular automaton? Can a shift-ergodic measure with positive but sub-maximum entropy on $(\xSp{X},\sigma)$ be invariant under a strongly transitive cellular automaton? \end{question} \subsection{Randomization and Approach to Equilibrium} \label{sec:ca:randomization} This section contains a few remarks and open questions regarding the problem of approach to equilibrium in surjective cellular automata. \begin{example}[Randomization in XOR cellular automata] \label{exp:xor:randomization} The XOR cellular automata (Examples~\ref{exp:xor:ca},~\ref{exp:xor:non-gibbs} and~\ref{exp:xor:non-gibbs:contd}) exhibit the same kind of ``approach to equilibrium'' as observed in the Q2R model (see the \hyperref[sec:intro]{Introduction}). Starting from a biased Bernoulli random configuration, the system quickly reaches a uniformly random state, where it remains (see Figure~\ref{fig:xor:randomization}). A mathematical explanation of this behavior was first found independently by Miyamoto~\cite{Miy79} and Lind~\cite{Lin84} (following Wolfram~\cite{Wol83}) and has since been extended and strengthened by others. Let $\xSp{X}\IsDef\{\symb{0},\symb{1}\}^\xZ$, and consider the XOR cellular automaton $\Phi:\xSp{X}\to\xSp{X}$ with neighborhood $\{0,1\}$. If $\pi$ is a shift-invariant probability measure on $\xSp{X}$, the convergence of $\Phi^t\pi$ as $t\to\infty$ fails as long as $\pi$ is strongly mixing and different from the uniform Bernoulli measure and the Dirac measures concentrated at one of the two uniform configurations~\cite{Miy79,Miy94}. However, if $\pi$ is a non-degenerate Bernoulli measure, the convergence holds if a negligible set of time steps are ignored. More precisely, there is a set $J\subseteq\xN$ of density~$1$ such that for every non-degenerate Bernoulli measure $\pi$, the sequence $\{\Phi^t\pi\}_{t\in J}$ converges, as $t\to\infty$, to the uniform Bernoulli measure~$\mu$. In particular, we have the convergence of the Ces\`aro averages \begin{align} \frac{1}{n}\sum_{t=0}^{n-1}\Phi^t\pi\to\mu \end{align} as $n\to\infty$~\cite{Miy79,Lin84}. The same type of convergence holds as long as $\pi$ is harmonically mixing~\cite{PivYas02}. Similar results have been obtained for a wide range of algebraic cellular automata (see e.g.~\cite{CaiLuo93,MaaMar98,FerMaaMarNey00,PivYas02, HosMaaMar03,PivYas04,PivYas06,Sob08}). In particular, the reversible cellular automaton of Example~\ref{exp:xor:transpose} has been shown to have the same randomizing effect~\cite{MaaMar99}. See~\cite{Piv09} for a survey. It is also worth mentioning a similar result due to Johnson and Rudolph~\cite{JohRud95} regarding maps of the unit circle $\xT\IsDef\xR/\xZ$. Namely, let $\pi$ be a Borel measure on $\xT$. They showed that if $\pi$ is invariant, ergodic and of positive entropy for the map $3\times: x\mapsto 3x \pmod{1}$, then it is randomized by the map $2\times: x\mapsto 2x \pmod{1}$, in the sense that $(2\times)^t\pi$ converges to the Lebesgue measure along a subsequence $J\subseteq\xN$ of density~$1$. \hfill$\ocircle$ \end{example} \begin{figure} \begin{center} \includegraphics[width=0.6\textwidth]{figures/xor-500x300-bmp-104-504.pdf} \end{center} \caption{% Randomization effect of the XOR cellular automaton (see Example~\ref{exp:xor:randomization}). Time goes downward. The initial configuration is chosen using coin flips with bias $1{\,:\,}9$. The initial density of symbol $\symb{1}$ is $0.104$. The density at time $300$ is $0.504$. } \label{fig:xor:randomization} \end{figure} Randomization behavior similar to that in the XOR cellular automaton has been observed in simulations of other (non-additive) cellular automata, but the mathematical results are so far limited to algebraic cellular automata. The uniform Bernoulli measure is the unique measure with maximum entropy on the full shift $(\xSp{X},\sigma)$ (i.e., the ``state of maximum randomness''). The convergence (in density) of $\Phi^t\pi$ to the uniform Bernoulli measure may thus be interpreted as a manifestation of the second law of thermodynamics~\cite{Piv09}. We say that a cellular automaton $\Phi:\xSp{X}\to\xSp{X}$ (asymptotically) \emph{randomizes} a probability measure $\pi\in\xSx{P}(\xSp{X})$, if there is a set $J\subseteq\xN$ of density~$1$ such that the weak limit \begin{align} \Phi^\infty\pi &\IsDef \lim_{J\ni t\to\infty}\Phi^t\pi \end{align} exists and is a shift-invariant measure with maximum entropy, that is, $h_{\Phi^\infty\pi}(\xSp{X},\sigma)=h(\xSp{X},\sigma)$. The \emph{density} of a set $J\subseteq\xN$ is defined as \begin{align} d(J) &\IsDef \lim_{n\to\infty} \frac{ J \cap \{0,1,\ldots,n-1\} }{n} \;. \end{align} Note that the limit measure $\Phi^\infty\pi$ must be invariant under $\Phi$, even if $(\xSp{X},\sigma)$ has multiple measures with maximum entropy. If $\Phi$ randomizes a measure $\pi$, the Ces\`aro averages $(\sum_{t=0}^{n-1} \Phi^t\pi)/n$ will also converge to $\Phi^\infty\pi$. The converse is also true as long as $\pi$ is shift-invariant and the limit measure is shift-ergodic: \begin{lemma}[see~\cite{JohRud95}, Corollary~1.4] \label{lem:density-vs-cesaro} Let $\xSp{X}$ be a compact metric space and $\xSx{Q}\subseteq\xSx{P}(\xSp{X})$ a closed and convex set of probability measures on $\xSp{X}$. Let $\pi_1,\pi_2,\ldots$ be a sequence of elements in $\xSx{Q}$ whose Ces\`aro averages $(\sum_{i=0}^{n-1} \pi_i)/n$ converge to a measure $\mu$ as $n\to\infty$. If $\mu$ is extremal in $\xSx{Q}$, then there is a set $J\subseteq\xN$ of density~$1$ such that $\pi_i\to\mu$ as $J\ni i\to\infty$. \end{lemma} As mentioned in Example~\ref{exp:xor:randomization}, the stronger notion of randomization fails for the XOR cellular automaton. We say that a cellular automaton $\Phi$ \emph{strongly randomizes} a measure $\pi$ if $\Phi^t\pi$ converges to a measure with maximum entropy. \begin{question} Are there examples of surjective or reversible cellular automata that strongly randomize all (say) Bernoulli measures? Is there a generic obstacle against strong randomization in surjective or reversible cellular automata? \end{question} If the cellular automaton $\Phi$ has non-trivial conservation laws, the orbit of a measure $\pi$ will be entirely on the same ``energy level''. Nevertheless, we could expect $\pi$ to be randomized within its energy level. To evade an abundance of invariant measures, let us assume that $\Phi$ has only finitely many linearly independent conservation laws. More precisely, let $F=\{f_1,f_2,\ldots,f_n\}\subseteq C(\xSp{X})$ be a collection of observables conserved by $\Phi$ such that every observable $g\in C(\xSp{X})$ conserved by $\Phi$ is physically equivalent to an element of the linear span of $F$. The measures $\Phi^t\pi$ as well as their accumulation points are confined in the closed convex set \begin{align} \{\nu\in\xSx{P}(\xSp{X}): \text{$\nu(f_i)=\pi(f_i)$ for $i=1,2,\ldots,n$} \} \;. \end{align} Let us say that $\Phi$ \emph{randomizes} $\pi$ \emph{modulo $F$} if there is a set $J\subseteq\xN$ of density~$1$ such that \begin{align} \Phi^\infty\pi &\IsDef \lim_{J\ni t\to\infty}\Phi^t\pi \end{align} exists, is shift-invariant, and has entropy $s_{f_1,f_2,\ldots,f_n}(\pi(f_1),\pi(f_2),\ldots,\pi(f_n))$, where \begin{align} s_{f_1,f_2,\ldots,f_n}(e_1,e_2,\ldots,e_n) &= \sup\{h_\nu(\xSp{X},\sigma): \text{$\nu\in\xSx{P}(\xSp{X},\sigma)$ and $\nu(f_i)=e_i$ for $i=1,2,\ldots,n$} \} \;. \end{align} \begin{question} What are some examples of non-algebraic cellular automata (with or without non-trivial conservation laws) having a randomization property? \end{question} Suitable candidates to inspect for the occurrence of a randomization behavior are those that do not have any non-trivial conservation laws. \begin{question} Do strongly transitive cellular automata randomize every Gibbs measure? \end{question} \begin{question} Does a one-dimensional reversible cellular automaton that has a positively expansive transpose randomize every Gibbs measure? \end{question} \section{Conclusions} There is a wealth of open issues in connection with the statistical mechanics of reversible and surjective cellular automata. We have asked a few questions in this article. From the modeling point of view, there are at least three central problems that need to be addressed: \begin{itemize} \item What is a good description of macroscopic equilibrium states? \item What is a satisfactory description of approach to equilibrium? \item How do physical phenomena such as phase transition appear in the dynamical setting of cellular automata? \end{itemize} By virtue of their symbolic nature, various questions regarding cellular automata can be conveniently approached using computational and algorithmic methods. Nevertheless, many fundamental global properties of cellular automata have turned out to be algorithmically undecidable, at least in two and higher dimensions. For example, the question of whether a given two-dimensional cellular automaton is reversible (or surjective) is undecidable~\cite{Kar94a}. Similarly, \emph{all} non-trivial properties of the limit sets of cellular automata are undecidable, even when restricted to the one-dimensional case~\cite{Kar94b,GuiRic10} (see also~\cite{Del11}). Whether a given cellular automaton on a full shift conserves a given local observable can be verified using a simple algorithm~\cite{HatTak91}, but whether a (one-dimensional) cellular automaton has any non-trivial local conservation law is undecidable~\cite{ForKarTaa11}. It is an interesting open problem whether the latter undecidability statement remains true when restricted to the class of reversible (or surjective) cellular automata. We hope to address this and other algorithmic questions related to the statistical mechanics of cellular automata in a separate study. Problems similar to those studied here have been addressed in different but related settings and with various motivations. Simple necessary and sufficient conditions have been obtained that characterize when a one-dimensional probabilistic cellular automaton has a Bernoulli or Markov invariant measure~\cite{TooVasStaMitKurPir90,MaiMar12}. The equivalence of parts~(b) and~(c) in Theorem~\ref{thm:conservation-invariance:1} is also true for positive-rate probabilistic cellular automata~\cite{DaiLuiRoe02}. For positive-rate probabilistic cellular automata, however, the existence of an invariant Gibbs measure implies that all shift-invariant invariant Gibbs measures are Gibbs for the same Hamiltonian! The ergodicity problem of the probabilistic cellular automata (see e.g.~\cite{TooVasStaMitKurPir90}) has close similarity with the problem of randomization in surjective cellular automata. \section*{Acknowledgements} We would like to thank Aernout van Enter, Nishant Chandgotia, Felipe Garc\'ia-Ramos, Tom Kempton and Marcus Pivato for helpful comments and discussions. \addcontentsline{toc}{section}{References} \bibliographystyle{plain}
\section*{Introduction} Electron spins in quantum dots (QDs) can be considered as promising candidates for realization of quantum computation ideas and spintronics devices. \cite{a1,a2} The main parameter indicating the applicability of a system for quantum computation is the spin coherence time. An extremely long spin lifetime is observed in zero-dimensional structures due to a strong confinement in all three dimensions. \cite{a3} An especially great potential for a long coherence time is expected in the Ge/Si system with quantum dots. In this system electrons are localized in strained Si regions, where spin-orbit (SO) coupling is very weak. However, recent investigations of spin decoherence by spin echo method in the Ge/Si QD system \cite{a4} demonstrated that the spin relaxation times are unexpectedly short ($\sim10\mu s$). It was suggested that the reason of such intensive spin relaxation consists in the appearance of effective magnetic fields during electron tunneling between quantum dots. These magnetic fields (Rashba fields)\cite{a5} originate from spin-orbit interaction and arise due to the absence of mirror symmetry of the localizing potential for an electron in the vicinity of Ge QD. Spin relaxation occurs through stochastic spin precession in effective magnetic fields during random tunneling between QDs (analog of Dyakonov-Perel mechanism for delocalized carriers\cite{a6}). Obviously, suppression of the tunneling in the array of well-separated quantum dots allows to eliminate the existence of in-plane fluctuating magnetic fields. In this case the hyperfine interaction with $^{29}$Si nuclear spins comes into force and determines the spin relaxation time. If the tunneling is suppressed not by spatial separation of QDs, but by Coulomb repulsion \cite{a7}, the anisotropic exchange interaction can also control the spin relaxation process. The efficiency of each mechanism depends in different ways on the localization degree of electrons. Changing the tunnel coupling between quantum dots (by changing their density) and correspondingly, the localization degree of electrons, it is possible to alter the relative contribution of different mechanisms. With the increase of electron localization radius the contribution of hyperfine interaction becomes smaller due to averaging-out of different orientations of nuclear spins. A related increase of the relaxation time occurs until the moment when the wave function overlapping provides the hopping between neighboring localization centers. In these conditions Dyakonov-Perel mechanism begins to control the spin relaxation process.\cite{a8} Longest spin relaxation time is expected right before the point where Dyakonov-Perel mechanism comes into force. Similar effect was detected for n-type GaAs impurity system, where threefold increase of spin relaxation time was obtained in the vicinity of metal-to-insulator transition.\cite{a9} In self-assembled tunnel-coupled QD structures it is hard to get a gradual change of the localization radius by changing the QD array density. The stochastic nucleation of QDs during growth in Stranskii-Krastanov mode \cite{a10} leads to the formation of the regions with a high local density of QDs. In such regions a strong tunnel coupling between dots results in the intensive spin relaxation through Dyakonov-Perel mechanism. However, under certain conditions, the existence of groups of closely located QDs can provide not a decrease, but an increase of the spin relaxation time. First, QD groups should be well separated from each other. In this case, the effective magnetic fields can be averaged due to electron back-and-forth motion within each QD group. Second, QDs inside the group should have strong tunnel coupling providing the effective localization radius comparable with QD group size. As a result, the averaging of local magnetic fields related to nuclear spins will take place. The present work is devoted to the electron spin resonance (ESR) study of inhomogeneous QD arrays, where the averaging of Rashba and hyperfine fields inside QD groups is expected to provide a long spin relaxation time. We succeed in creating the experimental structure containing well separated groups of QDs with a large electron localization radius. The coupling between QDs and, consequently, the electron localization radius in the structures under study turned out to be dependent on the external magnetic field orientation. A fourfold increase of the spin relaxation time as compared to the previous data for dense homogeneous QD arrays \cite{a4} has been detected at a special orientation of magnetic field, where the electron localization radius was close to the QD group size. \section {Samples and Experiment} The samples were grown by molecular-beam epitaxy on n-Si$(001)$ substrates with the resistivity of 1000 $\Omega\,$cm. To increase the response from the sample, we have grown 6 layers of the Ge nanoclusters separated by 30 nm Si-layers. Each QD layer was formed by deposition of 7 ML Ge at the temperature $T=550^{\circ}$C. On the top of the structure, a $0.3\,\mu$m epitaxial n-Si layer (Sb concentration $\geq10^{17} $cm$^{-3}$) was grown, the same layer was formed below QD layers. The scanning tunneling microscopy (STM) of the structure with a single QD layer uncovered by Si shows the bimodal distribution of QDs ($hut$- and $dome$-clusters) (Fig.~1). The density of $dome$-clusters is $\sim10^{10}$~cm$^{-2}$, the typical base width is $l=50$~nm, the height is $h=10$~nm. The $hut$-clusters are distributed between $dome$-clusters with density $\sim10^{11}$~cm$^{-2}$, their typical base width is $l=15$~nm, the height is $h=1.5$~nm. The obtained sizes of $dome$-clusters have to provide a strong localization of electrons at the apex of such type QDs with a small localization radius. To increase the localization radius, we used the temperature 500$^{\circ}$C for overgrowth of QDs allowing to transform $dome$-clusters to disk-like clusters without intensive Ge-Si intermixing inside QDs. The cross-section images obtained by TEM ( transmission electron microscopy) show that the height of disk-like dots does not exceed 3 nm in the experimental structure. These dots, as well as original $domes$, are characterized by the absence of mirror symmetry due to a difference between the smeared top and sharper bottom of QDs. So, after such overgrowth the localization radius is expected to be comparable with the lateral size of QD. STM data show a non-homogenous in-plane distribution of $dome$-clusters and the existence of groups of 2-3 closely spaced nanoclusters on the average (Fig.~1). A sufficient tunneling coupling between them allows the electron wave function to spread over the group of QDs and promotes a further increase of the electron localization radius. $Hut$-clusters in these structures can not be the centers of localization because the binding energy of electron on such type quantum dots is very small ($\sim10$meV).\cite{a11} Recently, to provide the localization of electrons on $hut$-clusters, the stacked structures with four layers of Ge quantum dots were grown \cite{a12}. The distances between QD layers were 3~nm, 5~nm, and that resulted in the effective deepening of potential well near $hut$-clusters due to accumulation of strain from different QD layers. In the structure under study the distance between QD layers is 30 nm, then the strain accumulation does not occur. \begin{figure} \includegraphics[width=2.8in]{fig1} \caption{\label{f1} Right panel: STM of the uncovered sample with two-shaped QDs ($hut$-clusters and $dome$-clusters), 1 $\mu$m$\times$1 $\mu$m image. Left panel: a schematic structure of investigated sample. The examples of quantum dot groups are indicated by dashed line loops.} \end{figure} ESR measurements were performed with a Bruker Elexsys~580 X-band EPR spectrometer using a dielectric cavity Bruker ER-4118 X-MD-5. The samples were glued on a quartz holder, then the entire cavity and the sample were maintained at a low temperature with a helium flow cryostat (Oxford CF935). The needless EPR signal from dangling bonds ($g=2.0055$) was avoided using the passivation of structures with atomic hydrogen before measurements. To increase the number of registrable spins, the sandwiched sample was prepared. Samples were thinned by acid etching up to $150-250\,\mu$m. After thinning, samples were glued together; finally, the object composed of 4-5 wafers was investigated. The spin echo measurements were carried out at temperature 4.5~K in resonance magnetic field $H=3470$~G (can be slightly varied $\pm5$~G depending on resonance conditions) with direction corresponding to the narrowest ESR line width, $\theta=30^{\circ}$, where $\theta$ is the angle between magnetic field and growth direction of the structure [001]. A two-pulse Hahn echo experiment ($\pi/2\!-\!\tau\!-\!\pi\!-\!\tau\!-$~echo) was used to measure $T_2$ (a detailed explanation can be found in Ref.~\onlinecite{a13}). In order to observe a longitudinal spin relaxation (corresponding time $T_1$), a different pulse sequence is applied ($\pi\!-\!\tau\!-\!\pi/2\!-\!T\!-\!\pi\!-\!T\!-$~echo). The first $\pi$-pulse rotates the magnetization opposite to its thermal equilibrium orientation, where the interaction with the environment causes the spins to relax back to the initial orientation parallel to $\mathbf{H}$. After time $\tau$, a $\pi/2$-pulse followed by another $\pi$-pulse is used to observe a Hahn echo. In the first and second type of experiments, the durations of $\pi/2$ and $\pi$ pulses were 60~ns and 120~ns, respectively; the interpulse time in the second experiment was kept $T=200$~ns. \section {Results} The ESR spectra measured at different directions of magnetic field are shown in Fig.~2, where $\theta=0^{\circ}$ corresponds to the magnetic field applied parallel to the growth direction $Z$. At $\theta=0^{\circ}$ the ESR line is most symmetrical and its shape is close to Gaussian. The line asymmetry becomes more pronounced with the increase of angle $\theta$ and the line shape tends to Lorentzian already at $\theta=10^{\circ}$. The line shape analysis performed at $\theta=30^{\circ}$ is shown in Fig.~3. A careful examination of ESR line shape shows that the ESR line represents the sum of an absorption line (dotted) and a dispersion line (dashed). The rotation of the sample in the magnetic field results in a change of the resonance line width and the resonance field. The orientation dependence of the ESR line width for the structure under study is demonstrated in Fig.~4 . When the external magnetic field deviates from the growth direction up to $\theta\approx30^{\circ}$, the ESR line width sharply decreases from $\Delta H$=1.9~Oe to $\Delta H$=1.4~Oe. A further tilt of the magnetic field leads to the line broadening with maximum $\Delta H$=2.4~Oe at $\theta=60^{\circ}$. For the in-plane magnetic field, the ESR line width is narrowed again down to $\Delta H$=1.8~Oe. Such nonmonotomic behavior is unusual for electrons in 2D system, and has not been observed up to now. The angular dependence of g-factor is shown in Fig.~5. At small angles (up to $30^{\circ}$) the g-factor slightly changes nearly $g=1.9994(5)$. Between $\theta=30^{\circ}$ and $\theta=40^{\circ}$ the g-factor value jumps to $g=1.9992$ and remains nearly constant up to $\theta=90^{\circ}$. \begin{figure} \includegraphics[width=3.3in]{fig2} \caption{\label{f2} ESR spectra at different orientations of magnetic field. For $\theta=0^\circ$ the magnetic field is parallel to the growth direction of the structure [001], $\theta=90^\circ$ corresponds to magnetic field applied along crystallographic direction [110].} \end{figure} \begin{figure} \includegraphics[width=2.8in]{fig3} \caption{\label{f6} Analysis of the ESR line shape for $\theta=30^{\circ}$. The solid line represents the sum of an absorption line (dotted), and a dispersion line (dashed).} \end{figure} \begin{figure} \includegraphics[width=3.2in]{fig4} \caption{\label{f64} The angular dependence of ESR line width and theoretical approximation (Eq.6) of this dependence (solid line) for the structure under study. For $\theta=0^\circ$ the magnetic field is parallel to the growth direction of the structure. } \end{figure} \begin{figure} \includegraphics[width=3.2in]{fig5} \caption{\label{f25} The angular dependence of electron g-factor for the structure under study. For $\theta=0^\circ$ the magnetic field is parallel to the growth direction of the structure.} \end{figure} The data of spin echo measurements performing at $\theta=30^{\circ}$, when the most narrow ESR line width is observed, are shown in Fig.~6,~7. \begin{figure} \includegraphics[width=3.5in]{fig6} \caption{\label{f4} Results of two-pulse spin echo experiments performed at $\theta=30^\circ$ (points) and approximation by superposition of two exponential functions, Eq.1 (solid line). Corresponding microwave pulse sequence is $\pi/2-\tau -\pi-\tau-echo$.} \end{figure} According to the results of a two-pulse Hahn echo experiment, the spin echo behavior can be described by superposition of two exponentially decaying functions: \begin{eqnarray} M(t)=M^{(1)}_{x,y}\,\exp(-2\tau/T_2^{(1)})+M^{(2)}_{x,y}\,\exp(-2\tau/T_2^{(2)}), \end{eqnarray} where $M(0)=M^{(1)}_{x,y}+M^{(2)}_{x,y}$ is the lateral (in QD plane) magnetization after $\pi/2$-pulse. The decay parameters give two times of spin dephasing: $T_2^{(1)}\approx0.26 ~\mu$s and $T_2^{(2)}\approx1.5~\mu$s. The analysis of the inversion signal recovery measured in three-pulse echo experiments shows non-exponential behavior (Fig.~7). The experimental curve can be described by the superposition of two functions: \begin{eqnarray} M(t)=M_{0z}-M^{(1)}_z \exp(-\tau/T_1^{(1)})-M^{(2)}_z \exp(-\tau/T_1^{(2)}), \end{eqnarray} where $M_{0z}$ is the equilibrium magnetization, $M_{0z}=M^{(1)}_{0z}+M^{(2)}_{0z}$, $M^{(1,2)}_z=M^{(1,2)}_{0z}-M^{(1,2)}_z(0)$, $M_z(0)=M^{1}_z(0)+M^{2}_z(0)$ is the magnetization just after applying of an inverting $\pi$-pulse. In correspondence with this equation at the beginning the samples magnetization recovers very fast. After some time the part of spins is returned to an equilibrium state and the recovery rate becomes much slower. The characteristic times obtained by fitting the experimental data are $T_1^{(1)}\!\approx\!2\,\mu s$ and $T_1^{(2)}\!\approx\!35\,\mu$s. All values of spin relaxation times were determined with error $\pm20\%$. \begin{figure} \includegraphics[width=3.5in]{fig7} \caption{\label{f47} Amplitude of inversion-recovery signal versus interpulse delay $\tau$ (symbols are experimental data, solid line is the approximation by Eq.2). Corresponding microwave pulse sequence is $\pi-\tau -\pi/2-T-\pi-T-echo$. Experiments are performed at $\theta=30^\circ$.} \end{figure} \section {Discussion} To explain the experimental results obtained in the present work, we propose the following model (see Fig.~8). Electrons are suggested to localize mainly in the groups of closely spaced QDs containing 2-3 QDs on the average (see STM data). During overgrowth QDs lose their apexes and transform to the disk-like shape QDs. As a result, the electron localization radius can become comparable with QD lateral size. Additional barriers for electrons limiting the electron motion in $XY$ directions in the Si layer arise due to the regions with a nonzero Ge content. These regions are located over the edges along the perimeter of QDs and they are formed following the mechanism of formation of SiGe rings described in Ref.\onlinecite{a14}. Thus, the electron localization radius can be taken about 50 nm for a spatially isolated single QD. In the case of a group of closely spaced QDs the separating SiGe barriers between dots inside the group are absent because of energetically unfavorable positions of Ge atoms between QDs due to a high strain.\cite{a15} SiGe barriers remain only along the external border of QD groups. Then the electron wave function can spread to the size of the QD group, $l\sim100-150$ nm. Since the confinement of electrons is not too strong, the tails of electron wave functions from different QD groups can be overlapped providing the hopping between QD groups. The external magnetic field applied along the growth direction can sufficiently change the described picture. Magnetic length $\lambda=\sqrt{c\hbar/eH} $, in our experimental set up (H=3470 Oe), is about 45 nm, that is comparable with the electron localization radius for a single QD. In these conditions, the magnetic field effectively shrinks the tails of electron wave functions, resulting in the enhancement of electron localization. Thus, the perpendicular magnetic field suppresses the transitions between QD groups and decreases the electron localization radius in the QD group down to the size of an individual QD. Nevertheless, the transitions between QDs inside the groups still persist due to a small distance between QDs. With the deviation of the magnetic field from the growth direction the probability of electron transitions between dots increases and the conductivity in local areas (QD groups) becomes higher. In experiment, this corresponds to the appearance of a noticeable dispersion signal and the enhancement of asymmetry of ESR line (ESR line becomes close to Dysonian line \cite{a16}). A similar effect was observed for SiGe/Si/SiGe structures with two-dimensional (2D) electron gas.\cite{a17} At the same time the increase of the effective electron localization radius causes the averaging of the local magnetic fields induced by nuclear spins and the smoothing of the QD parameter differences within a QD group. As a result, the narrowing of ESR line at the deviation of magnetic field from $\theta=0^{\circ}$ to $\theta=30^{\circ}$ is observed. The minimum of ESR line width at $\theta=30^{\circ}$ indicates that the electron localization radius reaches the size of QD group and the full averaging inside each group takes place. A further increase of the electron localization radius leads to the enhancement of hopping between groups and a decrease of spin lifetime through Dyakonov-Perel spin relaxation mechanism. The decreasing spin relaxation time affects the ESR line width. From $\theta=30^{\circ}$ the broadening due to spin relaxation, $\Delta H\cong1/T_2$, exceeds the nonhomogeneous broadening and further orientation dependence of ESR line width is controlled by spin relaxation time. It should be noted that there is one more mechanism which can provide the anisotropy of ESR line width, the relaxation assisted by spin-phonon interaction. This mechanism can be effective due to the lack of phonon bottleneck in our structures with large QD sizes resulting in small confinement energies of electrons. In this case the anisotropy of spin relaxation processes is defined by the shape asymmetry of disk-like QDs. Their lateral size is one order larger than their height, therefore only $k_x$- and $k_y$-phonon waves effectively influence the spin. However, the experimentally observed maximum of ESR line width at $\theta=60^{\circ}$ cannot be described in the frames of the spin-phonon interaction model,\cite{a18} which should provide a monotonic orientation dependence of ESR line width. Dyakonov-Perel mechanism allows to describe non-monotonic behavior of ESR line width on the assumption of $\tau_h$ depending on the magnetic field. This dependence, determined in the frame of the hopping model, \cite{a19} can be described as exponential: \begin{eqnarray} \tau_h=\tau_0\exp(\alpha H_z^m), \end{eqnarray} where $H_z$ is the projection of magnetic field to the growth direction; coefficient $m$ can be equal to $m=1/2$ or $m=2$ for case of strong or weak magnetic fields. For intermediate fields $\lambda\sim l$ this coefficient can take a value in the range of $\{\frac{1}{2}\div2\}$ (Ref.~\onlinecite{a19}). Spin relaxation time $T_2$ in the frames of Redfield theory \cite{a20} is given by the following expression: \begin{eqnarray} \frac{1}{T_2}=\gamma^2\delta H^{2}_{y}\sin^2\theta \tau_h+\frac{1}{2T_1}, \end{eqnarray} with \begin{eqnarray} \frac{1}{T_1}=\gamma^2(\delta H^{2}_{x}+\delta H^{2}_{y}\cdot\cos^2\theta )\frac{\tau_h}{1+\omega^{2}_{0}\tau^2_h},\nonumber\\ \end{eqnarray} where correlation time of spin-orbit field fluctuations $\tau_c$ was replaced by hopping time $\tau_h$; the $\omega_0$ is the Larmor frequency; $\delta H_x$, $\delta H_y$ are the components of effective magnetic field $\delta H$. So, using expression (3) for $\tau_h$, we obtain the following expression describing the orientational dependence of ESR line width: \begin{eqnarray} \Delta{H}=B\exp(A \cos^m \theta)\left(\sin^2\theta+ \frac{1+\cos^2\theta}{1+C\exp(2A\cos^m\theta)}\right), \end{eqnarray} where $B=\gamma\delta H^{2}_{y}\tau_0$, $A=\alpha H^m$, $C=\omega_0^2\tau_0^2$. The experimental data $\Delta H(\theta)$ in the range of $\theta\in\{30^{\circ};90^{\circ}\}$ are well approximated by this expression (Fig.~4) with $m=3/2$, $A=1.52$, $B=1.78$, $C=795.2$. Obtained coefficient $m$ corresponds to the case of intermediate magnetic fields ($\lambda\sim l$) that argues for accepted hopping model with $\tau_h$ depending on the magnetic field. \begin{figure} \includegraphics[width=3.2in]{fig8} \caption{\label{f28} Illustration of a hopping model for the structure under study. Hopping transitions inside the QD groups provide the narrowing of ESR line with the deviation of magnetic field from $\theta=0^{\circ}$ to $\theta=30^{\circ}$. Hopping transitions between QD groups provoke the spin relaxation through the spin precession mechanism and give a special orientation dependence of ESR line width in the range of $\theta\in\{30^{\circ}-90^{\circ}\}$. } \end{figure} The magnitude of the effective magnetic field $\delta H$, estimated from the $B$ coefficient turns out to be $\approx15$~Oe. This value is twice smaller than that determined in our previous work \cite{a21} for $hut$ clusters with the aspect ratio $h/l$=0.1. It is known \cite{a22} that in a QD system the effective magnetic field depends on $h/l$, the more the aspect ratio the larger the $\delta H$. For QDs under study the aspect ratio is about of $\approx0.05$, therefore the effective magnetic field proved to be smaller. The spin echo data are in a good agreement with the proposed model of the existing closed groups of quantum dots being the centers of electron localization in the sample. Experimental spin polarization behavior shows that the spin relaxation occurs in two stages, rapid and slow ones. To understand the origin of this two-stage spin dynamics we simulated the spin relaxation process in the ring-shaped group of quantum dots. Model includes the strong tunneling coupling between quantum dots in the circle. Hopping between any neighboring QDs is permitted with an equal probability for back and forth motion. Each tunneling transition is accompanied by spin rotation on a small fixed angle $\alpha=0.1$. The direction of the rotation axis is defined by product $[\mathbf{n}\times \mathbf{e_z}]$, where $\mathbf{n}$ -tunneling direction, $\mathbf{e_z}$ -growth direction of QD. The external magnetic field is applied along $\mathbf{e_z}$ and provides the Larmor precession between tunneling events. The time intervals between tunneling events are distributed exponentially with a mean value $\tau_h$. The spin relaxation caused by the interaction with phonons and nuclear spins was not included into the consideration. The transport was simulated by Monte-Carlo method for a different number of QDs in the circle. The results of simulation for the ring constructed of 10 quantum dots are demonstrated in Fig.~9. The two stage spin dynamics is clearly seen. It turned out that this effect depends on the relation between hopping time $\tau_h$ and Larmor frequency $\omega_l$. The two-stage dynamics is observed when $\omega_l \tau_h \ll 1$. For example, the data in Fig.~9 were obtained at $\omega_l \tau_h = 0.1$. The first stage of spin relaxation is related to the processes of electron spreading all over the group of QDs. At this stage the loss of spin polarization occurs due to the precession in the effective magnetic field during tunneling between dots. The spin dynamics at the second stage is defined by the phase breaking of Larmor precession during a random walk along the QD ring. Generally speaking, this stage of spin relaxation can be ruled by spin-phonon or hyperfine interaction as well, if one includes them in the consideration. The absence of two-stage dynamics in the case of $\omega_l \tau_h \geq 1$ can be understood by the simple consideration of spin behavior in the frame of reference rotating with Larmor frequency. The randomness of hopping between dots leads to averaging of effective magnetic field $(\langle \delta H\rangle=0)$ and elimination of spin relaxation at the first stage of electron extension over the group of QDs. According to the simulation results, the first rapid stage is characterized by a special relation between the longitudinal and transverse spin relaxation times $T_1$ and $T_2$, usual for 2D system with the absence of mirror symmetry $T_2\approx2T_1$. Such relation was obtained by spin echo measurements for 2D electron gas structures \cite{a23} and for dense homogeneous QD arrays,\cite{a4} and follows from the in-plane arrangement of fluctuating magnetic fields $\delta H$. Also we have verified the presence of two-stage spin dynamics in QD clusters with other spatial arrangements, for example, QD lines containing a few dots. The described general features are well preserved with changing only numerical values of $T_1$ and $T_2$. \begin{figure} \includegraphics[width=3.2in]{fig9} \caption{\label{f29} Results of spin relaxation simulation in a closed ring-shaped group of QD. Number of QDs in the ring - $n=10$. The hopping time is taken as $\tau_h=10^{-11}$s, Larmor frequency was one order smaller $\omega=10^{10}$s$^{-1}$. Two stage dynamics is observed. } \end{figure} In experiment on the $T_2$-measurements the first stage is characterized by $T_2\approx0.26 ~\mu$s. Based on the simulation results one can expect the same shortness for $T_1$. However, the three-pulse method has limitations on measurements of such short times. The difference between durations of the pulse sequences in $T_1$-experiments and $T_2$-experiments is comparable with duration of the first rapid stage of spin relaxation that makes difficult the study of the beginning of $S_z$-relaxation. The second stage of spin relaxation has the times $T_1=2\,\mu$s and $T_2=1.5\,\mu$s. Here, the special relation $T_2\approx2T_1$ is not fulfilled because of the presence of some additional spin relaxation mechanisms, for example, Larmor precession phase breaking during a random walk along the QD group or another one. The presence of long living spin polarization with characteristic time $T_1\approx35\,\mu$s is attributed to some stabilization of the $S_z$-component taking place at the electron movement within a closed QD group. According to the simulation results at a high hopping frequency ($\omega_l \tau_h <0.01$) the spin polarization after rapid stage is settled at some level depending on parameters of QD groups. In this case, Larmor precession can be neglected, and the sequence of small turnings in Rashba fields can be considered as the effective precession around growth direction $Z$. In these conditions, the $S_z$-component is stabilized. In contrast, the transverse component of spin relaxes quickly, then in the experiment we did not observe the rest transverse spin polarization with a long relaxation time. The orientational dependence of the g-factor allows us to add some details to the considered model. The value of g-factor g=1.9994, within the experimental error, coincides with typical g-factor value for electron states near the conduction band edge in Si.\cite{a24} The fact that the g-factor remains near this value up to $\theta\approx30^{\circ}$ confirms that electron is located in Si regions until this orientation of magnetic field. In other words, the electron increases its localization radius remaining in Si regions. After point $\theta=30^{\circ}$, the electron localization radius exceeds the size of QD groups and the g-factor drastically changes to value $g=1.9992$. Such behavior can be explained by the penetration of electron wave function in SiGe regions surrounding QD groups. The presence of Ge atoms can provide a decrease of the g-factor value.\cite{a21} It should be noted that the same value of electron g-factor was obtained by us in another ESR experiment for the structure with large SiGe nanodisks having diameter 100-150 nm. For this structure we use substrate with specially created nucleation sites to obtain more ordered array of quantum dots. These nucleation sites originated due to strain modulation in the surface layer induced by previously buried QDs. Large $dome$-clusters grown at previous stage at temperature $650^{\circ}$ have a good spatial ordering due to a long range elastic interaction between QDs.\cite{a15} On this strain-modulated surface we have grown 10 layers of QDs using the same temperature regime as in the structure under study ($T=550^\circ$~C for QD growth and $T=500^\circ$~C for overgrowth by Si). However, we reduce the amount of deposited Ge down to 4~ML in each QD layer, and, as result, we obtain a well-ordered array of nanodisks after overgrowth by Si. Thus, we can compare two structures: 1) a non-ordered array with groups of closely spaced QDs and 2) a well-ordered array of nanodisks, one nanodisk instead of one QD group. The average size of QD groups coincides with the characteristic size of nanodiscs. ESR data obtained on the test structure with nanodisks confirm the model proposed in this work. ESR signal has isotropic g-factor $g=1.9992\pm0.0001$ and isotropic ESR line width $\Delta H_{pp}\approx0.4$~Oe. Absolute value of g-factor is the same as in the structure with QD groups at $\theta > 30^{\circ}$. This can be explained by identical electron localization radius and identical temperature regime of QD creation. The last factor defines the GeSi intermixing and strain in the QD system, which have a high influence on g-factor value. The isotropy of ESR line is explained by the absence of tunneling transitions between nanodisks, which are well ordered in the plane and positioned at an equal ($\sim100$~nm) distance from each other. Narrowness of ESR line indicates the high efficiency of averaging of nuclear magnetic fields by the electron state with large localization radius and the high uniformity of array of nanodisks (negligible inhomogeneous broadening). In the structure with QD groups the averaging by means of tunneling between dots is not so efficient, then we observe a few times larger ESR line width. In summary, we demonstrate that the existence of closely spaced QD groups provides the increase of spin relaxation time in QD system. Changing the electron localization radius by external magnetic field allows us to catch the effect of ESR line narrowing and obtain at the special orientation of magnetic field the fourfold increased time $T_1$ as compared to the recently studied homogeneous QD arrays. \begin{acknowledgments} This work was supported by RFBR (Grants 11-02-00629-a, 13-02-12105,), SB RAS integration project No. 83 and DITCS RAS project No. 2.5. \end{acknowledgments}
\section{Introduction} For intermediate-mass (`$sd$-shell' ) nuclei, which are currently inaccessible by standard {\it ab initio} no-core shell-model (NCSM) \cite{NCSM} calculations, symmetry-based considerations are essential. In particular, we employ the no-core symplectic (NCSpM) shell model for symmetry-preserving interactions \cite{DreyfussLDDB13} with \SpR{3} the underpinning symmetry \cite{RosensteelR77}. This symmetry is inherent to the dynamics of deformed nuclear systems \cite{RosensteelR80,DraayerWR84,Rowe85,BahriR00,DytrychSBDV06}. The present study uses a schematic, but fully microscopic and physically motivated effective many-nucleon interaction, a choice that enables the use of group-theoretical methods with analytical expressions for Hamiltonian matrix elements, and which in turn makes large space solutions for $sd$-shell nuclei feasible. Recently, we successfully applied the NCSpM model to the rotational and alpha-cluster substructures of $^{12}$C, including the Hoyle state (the second $0^+$ state in $^{12}$C) and its rotational band \cite{DreyfussLDDB13}, as well as of $^{8}$Be \cite{LauneyDDTFLDMVB12}. The symplectic model has been used previously to achieve a remarkable reproduction of enhanced $E2$ transition strengths in $^{20}$Ne without effective charges and with the use of a relatively simple symmetry-breaking valence-shell interaction \cite{DraayerWR84}. In addition, it has been applied to $^{166}$Er using the Davidson potential \cite{BahriR00}. The main objective of the present study is to offer qualitative results that can provide guidance for {\it ab initio} shell model approaches by informing key features of nuclear structure and the interaction, first on the physically relevant truncation of shell-model spaces, but also on the dominant deformation and particle-hole configurations. This is especially useful for the {\it ab initio} symmetry-adapted no-core shell model (SA-NCSM) \cite{DytrychPANP2013}, which will then bring forward, with the use of a realistic nucleon-nucleon interaction, an accurate reproduction and reliable prediction of energy spectra and associated transition rates that majorly impact modeling of stellar explosions and astrophysical processes. In this study, we explore the ground-state ($g.st.$) rotational band of lower $sd$-shell nuclei, namely, $^{20}$O, $^{20,22,24}$Ne, $^{20,22}$Mg, and $^{24}$Si. These low-lying states are expected to be highly influenced by large deformation. This, together with the combinatorial growth in model space dimensionality with number of particles and the spaces in which they primarily reside, has hitherto precluded a no-core shell model description and typically, in this region, valence shell model or mean field approaches have been employed (e.g., \cite{Brown01,CaurierMP05,Hinohara11}). Many of these nuclei are in close proximity to the proton drip line and are key to understanding, e.g., novae and X-ray bursts (see, e.g., \cite{BlackmonAS06}). In particular, properties of low-lying $2^+$ and $4^+$ states in isotopes as $^{20}$Mg and $^{24}$Si are required to predict $(p,\gamma)$ reaction rates that are expected to affect the light curve for X-ray bursts. As such unstable isotopes are very hard to make experimentally and state-of-the-art radioactive-beam measurements have only recently started to provide new information \cite{BlackmonAS06}, theoretical predictions are valuable. The present approach utilizes symmetry to reduce the dimensionality of the model space through a very structured winnowing of the basis states to physically relevant subspaces. Indeed, experimental evidence supports the fact that in this mass range, the dynamics favors a dominance of low spin and high deformation, which has been demonstrated by symmetry-guided theoretical studies \cite{RosensteelR80,DraayerWR84,Rowe85} as well as through an { \it ab initio} study \cite{DytrychSBDV06,DytrychDSBV09}. The latter exploits symplectic symmetry and its deformation-related \SU{3} subgroup in an analysis of { \it ab initio} large-scale nuclear physics applications for $^{12}$C and $^{16}$O. The outcome of this study has revealed that typically only one or two symplectic many-body basis states (vertical cones) suffice to represent a large fraction -- typically in excess of about 80\% of the physics -- as measured by projecting { \it ab initio} NCSM results onto a symmetry-adapted equivalent basis. Such a symplectic pattern has been also observed in an { \it ab initio} SA-NCSM study of $^{6}$Li, $^{6}$He, and $^{8}$Be \cite{DytrychPANP2013}. These findings point to the relevance of the symplectic symmetry, first to the many-body nuclear wavefunctions, and then to the inter-nucleon interaction (as symplectic basis states appear not to mix strongly). The NCSpM builds upon these considerations, and here we offer solutions to $sd$-shell nuclei in the framework of a fully microscopic no-core shell model. This, in turn, allows us to examine the role of currently inaccessible shell-model spaces, up through 15 major shells, and of associated particle excitations to these shells for a description of large deformation. \vspace{-0.2in} \section{Symmetry-informed Approach} We employ the no-core symplectic model (NCSpM), outlined in Refs. \cite{BahriR00}, with a novel interaction that is effectively realized by an exponential dependence on the quadrupole-quadrupole ($Q.Q$) two-body interaction, the physically relevant interaction of each particle with the total quadrupole moment of the nuclear system. This introduces simple but important many-body interactions that enter in a prescribed hierarchical way given in powers of a small parameter, the only adjustable parameter in the model. The model offers a microscopic no-core shell-model description of nuclei in terms of mixed deformed configurations and allows the inclusion of higher-energy particle excitations \cite{DreyfussLDDB13} that are currently inaccessible by {\it ab initio} shell models. It reduces to the successful Elliott model \cite{Elliott58ElliottH62} in the limit of a single valence shell and a zero model parameter. The underlying symmetry of the NCSpM is the symplectic \SpR{3} group \cite{RosensteelR77} and its embedded \SU{3} subgroup \cite{Elliott58ElliottH62}. The symplectic basis (detailed in \cite{Rowe85}) utilized in NCSpM is related, via a unitary transformation, to the three-dimensional HO ($m$-scheme) many-body basis used in the NCSM (see the review \cite{DytrychSDBV08}). The conventional NCSM basis spaces \cite{NCSM} are constructed using HO single-particle states and are characterized by the $\hbar\Omega$ oscillator strength (or equivalently, the oscillator length $b=\sqrt{ \frac{\hbar }{m\Omega} }$ for a nucleon mass $m$) as well as by the cutoff in total oscillator quanta, $N_{\max} $, above the lowest HO energy configuration for a given nucleus. Indeed, the NCSpM employed within a complete model space up through $N_{\max}$, will coincide with the NCSM for the same $N_{\max}$ cutoff. The important feature of the NCSpM model is its ability to down-select to the most relevant configurations, which are chosen among all possible \SpR{3} irreducible representations (irreps) within an $N_{\max}$ model space. The \SpR{3} irreps divide the space into `vertical cones' that are comprised of basis states of definite $(\lambda\,\mu)$ quantum numbers of \SU{3} linked to the intrinsic quadrupole deformation \cite{RosensteelR77b,LeschberD87,CastanosDL88}. E.g., the simplest cases, $(0\, 0)$, $(\lambda\, 0)$, and $(0\,\mu)$, describe spherical, prolate, and oblate deformation, respectively, while a general nuclear state is typically a superposition of several hundred various $(\lambda\,\mu)$ triaxial deformation configurations. \vspace{-0.2in} \subsection{Symplectic \SpR{3} group} The translationally invariant (intrinsic) symplectic generators can be written as \SU{3} tensor operators in terms of the harmonic oscillator raising, $b_{i \alpha}^{\dagger(1\,0)}=\frac{1}{\sqrt{2}}(X_{i \alpha}-iP_{i \alpha})$, and lowering $b^{(0\,1)}$ dimensionless operators (with $\mathbf X$ and $\mathbf P$ the lab-frame position and momentum coordinates and $\alpha=1,2,3$ for the three spatial directions), \begin{eqnarray} A^{(2\,0)}_{\mathfrak{L}M}\!\!&=&\!\! \frac{1}{\sqrt{2}} \sum_{i=1}^A \left[b_{i}^{\dagger}\times b_{i}^{\dagger}\right]^{(2\,0)}_{\mathfrak{L}M} - \frac{1}{\sqrt{2}A} \sum_{s,t=1}^A \left[b^{\dagger}_{s}\times b^{\dagger}_{t}\right]^{(2\,0)}_{\mathfrak{L}M} \label{sp3RgenA}\\ C^{(1\,1)}_{\mathfrak{L}M}\!\!&=&\!\! \sqrt{2} \sum_{i=1}^A \left[b_{i}^{\dagger}\times b_{i}\right]^{(1\,1)}_{\mathfrak{L}M} \!- \frac{\sqrt{2}}{A} \sum_{s,t=1}^A \left[b^{\dagger}_{s}\times b_{t}\right]^{(1\,1)}_{\mathfrak{L}M}, \label{sp3RgenC} \end{eqnarray} together with $B^{(0\,2)}_{\mathcal{L}M}=(-)^{\mathcal{L}-M}(A^{(2\,0)}_{\mathcal{L}-M})^{\dagger}$ ($\mathcal{L}=0,2$) and $H_{00}^{(00)}= \sqrt{3} \sum_{i} \left[b_{i}^{\dagger}\times b_{i}\right]^{(00)}_{00} -\frac{\sqrt{3}}{A} \sum_{s,t} \left[b^{\dagger}_{s}\times b_{t}\right]^{(0\,0)}_{00} +\frac{3}{2}(A-1)$, where the sums run over all $A$ particles of the system. The eight operators $C^{(1\,1)}_{\mathcal{L},M}$ ($\mathcal{L}=1,2$) generate the \SU{3} subgroup of \SpR{3}. They realize the angular momentum operator: \begin{equation} L_{1M}=C^{(1\,1)}_{1M},\, M=0,\pm1, \label{Lgen} \end{equation} and the Elliott algebraic quadrupole moment tensor $\mathcal{Q}^{a}_{2M}=\sqrt{3}C^{(1\,1)}_{2M},\, M=0,\pm1,\pm2$. The mass quadrupole moment can be constructed in terms of the symplectic generators as, \begin{equation} Q_{2M}=\sqrt{3}(A^{(2\,0)}_{2M}+C^{(1\,1)}_{2M}+B^{(0\,2)}_{2M}). \label{Qgen} \end{equation} \vspace{-0.2in} \subsection{Symplectic basis} A many-body basis state of a symplectic irrep is labeled according to the group chain, \begin{equation} \begin{array}{cccccccc} \SpR{3} & \supset & U(3) & \supset & \SO{3} & \supset & SO(2) \\ \sigma & n\rho & \omega & \kappa & L & & M \end{array} \end{equation} and constructed by acting with symmetrically coupled polynomials in the symplectic raising operators, $A^{(2\,0)}$, on a unique symplectic bandhead configuration, $\ket{\sigma}$, \begin{equation} |\sigma n\rho\omega \kappa L M\rangle = \left[ \left[A^{(2\,0)}\times A^{(2\,0)} \dots \times A^{(2\,0)}\right]^{n}\times\ket{\sigma}\right]^{\rho\omega}_{\kappa L M}, \label{basis} \end{equation} where $\sigma$ $\equiv $ $N_\sigma\left(\lambda_{\sigma}\, \mu_{\sigma}\right)$ labels the \SpR{3} irrep, $n\equiv N_{n}\left(\lambda_{n}\, \mu_{n}\right)$, $\omega\equiv N_\omega\left(\lambda_{\omega}\, \mu_{\omega}\right)$, and $N_{\omega}=N_{\sigma}+N_{n}$ is the total number of HO quanta ($\rho$ and $\kappa$ are multiplicity labels). This can be generalized to include spin, $|\sigma n\rho\omega \kappa (L S_{\sigma}) JM_J \rangle = \sum_{MM_S} \CG{L M}{S_{\sigma} M_S}{J M_J} |\sigma n\rho\omega \kappa L M S_{\sigma} M_S \rangle$, and also isospin. States within a symplectic irrep have the same spin (isospin) value, which is given by the spin $S_\sigma$ (isospin $T_\sigma$) of the bandhead $|\sigma; S_{\sigma}\rangle$ \cite{DytrychSDBV08}. Symplectic basis states span the entire shell-mode space\footnote{A complete set of labels includes additional quantum numbers $\ket{\left\{\alpha \right\}\sigma}$ that distinguish different bandheads with the same $N_\sigma\left(\lambda_{\sigma}\, \mu_{\sigma}\right)$. \SpR{3}-preserving Hamiltonians render energy spectra degenerate with respect to $\left\{\alpha \right\}$. However, for all present calculations for $g.st.$ rotational bands and associated observables, $\left\{\alpha \right\}$ is unique (an only set).}. The symplectic structure accommodates relevant particle-hole (p-h) configurations in a natural way (see also Fig. 1 of Ref. \cite{DreyfussLDDB13}). According to Eq. (\ref{basis}), the basis states of an \SpR{3} irrep (vertical cone) are built over a bandhead $\ket{\sigma}$ by 2\ensuremath{\hbar\Omega}~ \ph{1} (one particle raised by two shells) monopole ($\mathfrak{L}=0$) or quadrupole ($\mathfrak{L}=2$) excitations, realized by the first term in $A^{(2\,0)}_{\mathfrak{L}M}$ of Eq. (\ref{sp3RgenA}), together with a smaller 2\ensuremath{\hbar\Omega}~\ph{2} correction for eliminating the spurious center-of-mass (CM) motion, realized by the second term in $A^{(2\,0)}_{\mathfrak{L}M}$. The symplectic bandhead $\ket{\sigma}$ is the lowest-weight \SpR{3} state, which is defined by the usual requirement that the symplectic lowering operators annihilate it -- in analogy to a $|J,M_J=-J\rangle$ state for the case of the \SU{2} group of angular momentum, that is, $J_-|J,M_J=-J\rangle=0$. The bandhead, $\ket{\sigma; \kappa_\sigma L_\sigma M_\sigma}$, is an \SU{3}-coupled many-body state with a given nucleon distribution over the HO shells and while not utilized here, can be obtained in terms of the creation operators $a^\dagger_{(\eta\, 0)}=a^\dagger_\eta$, which create a particle in the HO shell $\eta=0,1,2,\dots$. E.g., for a 0\ensuremath{\hbar\Omega}~bandhead, the nucleon distribution is a single configuration, \begin{equation} \left[ a^\dagger_{(\eta_1\, 0)} \times a^\dagger_{(\eta_2\, 0)} \times \dots \times a^\dagger_{(\eta_A\, 0)} \right]^{(\lambda_{\sigma}\, \mu_{\sigma})}_{\kappa_\sigma L_\sigma M_\sigma}\ket{0} \end{equation} with $N_ \sigma = \eta_1+ \eta_2 + \dots + \eta_A+\frac{3}{2}(A-1)$, such that $N_\sigma \ensuremath{\hbar\Omega}$ includes the HO zero-point energy and $3/2$ is subtracted to ensure a proper treatment of the CM. To eliminate the spurious CM motion, the NCSpM also uses symplectic generators constructed in ${\mathbf r}_{i}$ (${\mathbf p}_{i}$) particle position (momentum) coordinates relative to the CM. These generators are used to build the basis, the interaction, the many-particle kinetic energy operator, as well as to evaluate observables. An example for the symplectic basis states follows for $^{24}$Mg. Its lowest HO-energy configuration is given by $N_\sigma= 62.5$ or 0\ensuremath{\hbar\Omega}, while the 4\ensuremath{\hbar\Omega}~$(20\,0)$ symplectic irrep includes: \begin{enumerate} \item A bandhead ($N_n=0$) with $N_\sigma= 66.5$ (or 4\ensuremath{\hbar\Omega}) and $(\lambda_ \sigma \,\mu_ \sigma)=(20\,0)$; \item $N_n=2$ states with $N_ \omega $=68.5 and $(\lambda_ \omega \,\mu_ \omega)= (22\, 0)$, $(20\, 1)$, and $(18\, 2)$; \item and so forth for higher $N_n$. \end{enumerate} For each $(\lambda_ \omega \,\mu_ \omega)$, the quantum numbers $\kappa$, $L$ and $M$ are given by Elliott \cite{Elliott58ElliottH62}. E.g., for $(22\,0)$, $\kappa=0$, $L=0,2,4,\dots,22$, and $M=-L,-L+1,\dots,L$. \subsection{Symmetry-preserving interactions} We note that the NCSpM, as presented here, is limited to interactions that preserve the \SpR{3} symmetry. This restriction facilitates the use of a group-theoretical apparatus and analytical expressions for the Hamiltonian matrix elements, which, in turn, makes it possible to incorporate large $N_{\max}$ spaces in applications of the theory. {\it Ab initio} calculations lie beyond the scope of the current analysis, but the addition of symmetry-mixing terms in the interaction is feasible, and a logical extension of the theory to include such terms is under development. \SpR{3}-symmetric Hamiltonians appear to be particularly suitable to capture the essential characteristics of the low-energy nuclear kinematics and dynamics. The reason is that important pieces of the inter-nucleon Hamiltonian of a quantum many-body system can be expressed in terms of the \SpR{3} generators, which directly relate to the relative particle momentum and position coordinates, as well as straightforwardly account for the Pauli exclusion principle. Indeed, the many-particle kinetic energy ($\sum_i\frac{{\mathbf p}_{i}^2}{2m}$), the HO potential ($\sum_i\frac{m\Omega^2 {\mathbf r}_{i}^2}{2}$), the mass quadrupole moment operator ($Q$), and the orbital momentum ($L$) are all elements of the $\SpR{3}\supset \Un{1}\times \SU{3} \supset\SO{3}$ structure. Hence, \SpR{3}-preserving Hamiltonians can include the many-particle kinetic energy: \begin{equation} \frac{T}{\ensuremath{\hbar\Omega}}=\frac{1}{\ensuremath{\hbar\Omega}}\sum_i\frac{{\mathbf p}_{i}^2}{2m}=\frac{1}{2}H_{00}^{(00)} -\sqrt{\frac{3}{8}}(A^{(2\,0)}_{00}+B^{(0\,2)}_{00}), \end{equation} the HO potential: \begin{equation} \frac{V_{HO}}{\ensuremath{\hbar\Omega}}=\frac{1}{\ensuremath{\hbar\Omega}}\sum_i\frac{m\Omega^2 {\mathbf r}_{i}^2}{2} =\frac{1}{2}H_{00}^{(00)} +\sqrt{\frac{3}{8}}(A^{(2\,0)}_{00}+B^{(0\,2)}_{00}), \end{equation} as well as terms dependent on $L$, see Eq. (\ref{Lgen}), and $Q$, see Eq. (\ref{Qgen}). These interactions have analytical matrix elements in the \SpR{3} basis (\ref{basis}) and act within a symplectic vertical cone ($\sigma_f= \sigma_i \equiv \sigma $). For example, for the dimensionless many-particle kinetic energy, $\frac{T}{\ensuremath{\hbar\Omega}}$, the matrix elements are given as: \begin{widetext} \begin{eqnarray} &&\braketop{\sigma n_f\rho_f\omega_f \kappa_f L_f M_f}{\frac{T}{\ensuremath{\hbar\Omega}}}{\sigma n_i\rho_i\omega_i \kappa_i L_i M_i} \nonumber \\ &&=\frac{1}{2}\braketop{\sigma n_f\rho_f\omega_f \kappa_f L_f M_f}{H_{00}^{(00)}}{\sigma n_i\rho_i\omega_i \kappa_i L_i M_i} -\sqrt{\frac{3}{8}}\braketop{\sigma n_f\rho_f\omega_f \kappa_f L_f M_f}{A^{(2\,0)}_{00}+B^{(0\,2)}_{00}}{\sigma n_i\rho_i\omega_i \kappa_i L_i M_i} \nonumber \\ &&= \frac{1}{2}N_\omega \delta_{f,i} -\sqrt{\frac{3}{8}} \left (\CG{\omega_i \kappa_i L_i M_i}{(2\,0)00}{\omega_f \kappa_f L_f M_f} \RedME{\sigma n_f\rho_f\omega_f }{A^{(2\,0)}}{\sigma n_i\rho_i\omega_i} +{\rm conjugate} \right), \end{eqnarray} \end{widetext} where $\CG{\omega_i \kappa_i L_i M_i}{(2\,0)00}{\omega_f \kappa_f L_f M_f} $ is an \SU{3} Clebsch-Gordan coefficient. The matrix elements of the \SpR{3} generators, $A^{(2\,0)}$, $B^{(0\,2)}$, and $C^{(1\,1)}$, reduced with respect to \SU{3}, such as $\RedME{\sigma n_f\rho_f\omega_f }{A^{(2\,0)}}{\sigma n_i\rho_i\omega_i}$, are known exactly \cite{RosensteelR83,RoweRC84,Rowe84,Hecht85,Rosensteel90} and the steps to calculate them are outlined in the appendix. The simplest, \SpR{3}-preserving Hamiltonian, besides the HO Hamiltonian ($H_0$), of importance to nuclear dynamics is \cite{Elliott58ElliottH62,BohrMottelson69}, \begin{equation} H_{\rm E} = H_0-\frac{\chi}{2}Q.Q, \end{equation} with $\chi$ being a coupling constant and, \begin{eqnarray} && H_0 = \sum_{i=1}^A \left(\frac{{\mathbf p}_{i}^2}{2m}+\frac{m\Omega^2 {\mathbf r}_{i}^2}{2}\right) \\ && \ensuremath{\textstyle{\frac{1}{2}}} Q\cdot Q = \ensuremath{\textstyle{\frac{1}{2}}} \sum_{ij} q(i)\cdot q(j). \end{eqnarray} Here $q_{2M}(i) =\sqrt{16\pi/5b^4} r_i^2Y_{2M}(\hat {\mathbf r}_i)$ is the dimensionless single-particle mass quadrupole moment and ${\mathbf r}_{i}$ (${\mathbf p}_{i}$) is the particle position (momentum) coordinate relative to the CM. In the limit of a single, valence shell ($N_n=0$), where \SU{3} becomes the relevant symmetry, the Hamiltonian $H_{\rm E}$ was shown to effectively describe rotational features of light nuclei in the framework of the established Elliott model \cite{Elliott58ElliottH62}. The success of such an effective nuclear interaction is not unexpected, as the spherical HO potential and the $Q.Q$ interaction directly follow from the second and third term, respectively, in the long-range expansion of any two-body central force, e.g., like the Yukawa radial dependence, $V^{(2)}=\sum _{i<j} V(r_{ij}/a)=\sum _{i<j} (\xi_0+\xi_2 r^2_{ij}/a^2+\xi_2 r^4_{ij}/a^4+ \dots)$ \cite{Harvey68}, for a range parameter $a$. However, in multi-shell studies, the attractive $Q.Q$ term becomes ever stronger with increasing $N_n$ and starts to dominate the dynamics. Hence, $H_{\rm E}$ yields unphysical solutions. A successful extension to multiple shells has been achieved and applied to the $^{24}$Mg $g.st.$ rotational band \cite{PetersonH80}, where an interaction given as a polynomial in $Q$, $Q\cdot Q$, $\left[ Q\times Q \right]\cdot Q$, and $(Q\cdot Q)^2$, was employed. Furthermore, in multi-shell studies, the $Q.Q-\langle Q.Q\rangle_{N_n}$ interaction has been employed \cite{CastanosD89,RosensteelD85}, where $\langle Q.Q\rangle_{N_n}$ is the average contribution of $Q.Q$ within the subspace of $N_n$ HO excitations, that is, the trace of $Q.Q$ divided by the space dimension for a fixed $N_n$. This removes the large monopole contribution of the $Q.Q$ interaction, which, in turn, helps eliminate the considerable renormalization of the zero-point HO energy, while retaining the $Q.Q$-driven behavior of the wavefunctions. \subsection{No-core symplectic model with $H_{\gamma}$ for intermediate-mass nuclei} We consider a novel effective many-nucleon interaction \cite{DreyfussLDDB13} suitable for large-$N_{\max}$ no-core shell models, \begin{equation} H_{\gamma} = H_0 + \frac{\chi}{2} \frac{\left( e^{-\gamma(Q.Q-\langle Q.Q\rangle_{N_n})} -1 \right)}{\gamma}, \label{effH} \end{equation} that addresses the limitations of the conventional $H_{\rm E}$, while retaining the $H_{\rm E}$ important features in the limit $\gamma \rightarrow 0$, where $\gamma$ is a positive adjustable parameter. We take $\chi=\ensuremath{\hbar\Omega}/(4\sqrt{N_{\omega,f}N_{\omega,i}})$ with $N_{\omega, f(i)}$ the total HO quanta of the final (initial) many-body basis state. The decrease of $\chi$ with $N_ \omega $, to a leading order in $\lambda/N_ \omega $, has been shown by Rowe \cite{Rowe67} based on self-consistent arguments and used in an \SpR{3}-based study of cluster-like states of $^{16}$O \cite{RoweTW06}. Above all, the effective interaction (\ref{effH}) introduces hierarchical many-body interactions in a prescribed way (for $\gamma \ll 1$). $H_\gamma$ also ensures that the $Q.Q$ term tails off for large $N_n$ eliminating its ever stronger attraction with increasing $N_n$. Such an interaction directly ties to the $Q$ polynomial considered in the above-mentioned study of Ref. \cite{PetersonH80}. Indeed, while higher-order terms in $Q\cdot Q$ of Eq. (\ref{effH}) could be understood as a renormalization (as shown in Ref. \cite{LeBlancCVR86}) of the $\chi$ coupling constant of the $NN$ interaction, $-\frac{1}{2}\sum_{ij} q(i) \cdot q(j)$: \begin{equation} \frac{\chi}{2 \gamma}{\left( e^{-\gamma Q\cdot Q} -1 \right)}= -\frac{1}{2}[\chi(\sum_{k=0}^{\infty} \frac{(-\gamma )^{k}(Q \cdot Q)^{k}}{(k+1)!}) ]Q \cdot Q, \end{equation} they become quickly negligible for a reasonably small $\gamma$. E.g., we find that for $^{12}$C, besides $Q\cdot Q$, only one term is sufficient for the ground-state band, while three terms are sufficient for the Hoyle-state band \cite{DreyfussLDDB13}. \begin{figure}[t] \begin{center} (a) \hspace{1in} (b) \hspace{1in} (c) \\ \includegraphics[width=0.32\columnwidth]{Fig/enMg20_g00205.pdf} \includegraphics[width=0.32\columnwidth]{Fig/enNe20_g00205.pdf} \includegraphics[width=0.32\columnwidth]{Fig/enO20_g00205.pdf} \end{center} \caption{ NCSpM energy spectrum of (a) $^{20}$Mg and (c) $^{20}$O using the $S_p=S_n=S=0$ $48.5 (4\, 2)$ \SpR{3} irrep built over the most deformed 0\ensuremath{\hbar\Omega}~ bandhead, as well as of (b) $^{20}$Ne using the $S_p=S_n=S=0$ $48.5(8\, 0)$ \SpR{3} irrep built over the most deformed 0\ensuremath{\hbar\Omega}~ bandhead. Experimental data (``Expt.") is from \cite{Tilley98A20}. $B(E2)$ transition rates are in W.u. units. } \label{enSpectrumNe20} \end{figure} \begin{figure}[t] \includegraphics[width=0.45\textwidth]{Fig/Ne20_0hw.pdf} \caption{ $^{20}$Ne low-lying states obtained by NCSpM for an $N_{\max}=12$ model space consisting of all possible 0\ensuremath{\hbar\Omega}~ \ph{0} symplectic bandheads (for $S=0$ and $S=1$; spin-2 states are not shown). The lowest-lying $48.5(8\, 0)$ \SpR{3} irrep is selected for the $^{20}$Ne calculations (Fig. \ref{enSpectrumNe20}b). } \label{enSpectrumNe20full0hw} \end{figure} \begin{figure*}[th] \begin{center} (a) \hspace{2in} (b) \hspace{2in} (c) \\ \includegraphics[width=0.32\textwidth]{Fig/Ne20_BE2_vsNmax.pdf} \includegraphics[width=0.32 \textwidth]{Fig/Ne20_Q_vsNmax.pdf} \includegraphics[width=0.32 \textwidth]{Fig/Ne20_rmsR_vsNmax.pdf} \end{center} \caption{ NCSpM observables for $^{20}$Ne using the \ph{0} $48.5(8\, 0)$ \SpR{3} irrep as a function of the model space, $N_{\max}$: (a) $B(E2;\, 2^+_1 \rightarrow 0^+_{g.st.})$ and $B(E2;\, 4^+_1 \rightarrow 2^+_1)$ transition strengths; (b) electric quadrupole moments for $2^+_1$ and $4^+_1$; and (c) the matter rms radius of the ground state. } \label{Ne20obsNmax} \end{figure*} For the NCSpM calculations, we use the empirical estimate $\ensuremath{\hbar\Omega} \approx 41/A^{1/3}$, namely, $\ensuremath{\hbar\Omega}=18$ MeV for $A=12$ (for analysis of $^{12}$C \cite{DreyfussLDDB13}) and $\ensuremath{\hbar\Omega}=15$ MeV for $A=20$ and $A=22$ isotopes. The \ensuremath{\hbar\Omega}~ value, in turn, fixes the $\chi$ coupling strength of the $Q.Q$-term. With $\ensuremath{\hbar\Omega} N_{\omega}$, the eigenvalue of $H_0$ in the model Hamiltonian (\ref{effH}), and with $\chi$ proportional to $\ensuremath{\hbar\Omega}/4$, the eigenstates are rendered \ensuremath{\hbar\Omega}-independent. \section{Results and Discussions} The NSCpM utilizes Bahri's symplectic computational code \cite{Bahri95_Sp3r_code} that uses Draayer \& Akiyama's numerical \SU{3} package \cite{AkiyamaD73}. The model has been successfully applied to the ground-state and Hoyle-state rotational bands in $^{12}$C \cite{DreyfussLDDB13}, where both rotational features and $\alpha$-cluster substructures have been described in the fully microscopic $N_{\rm max}=20$ no-core shell-model framework, as suggested by the reasonably close agreement of the model outcome with experiment and {\it ab initio} results in smaller spaces. The present study reveals that the model is also applicable to low-lying states of other light nuclei without any parameter adjustment, namely, we use $\gamma =0.74\times 10^{-4}$, the value obtained in the NCSpM analysis for $^{12}$C. In particular, we focus on the $g.st.$ rotational band of selected $A=20, 22$ and $24$ isotopes. We note that, for the $g.st.$ band as opposed to cluster-driven excited rotational bands, comparatively lower $N_{\rm max}$ values are necessary to achieve convergence of energies, $E2$ observables and radii, with $N_{\rm max}=12$ found to be sufficient for the present calculations. Model spaces are down-selected based on findings of {\it ab initio} large-scale calculations for $^{12}$C and $^{16}$O that have revealed low-spin and high-deformation dominance \cite{DytrychDSBV09}, as well as the importance of symplectic irreps built over the most deformed 0\ensuremath{\hbar\Omega}~ bandhead (the leading \SU{3} configuration) \cite{DytrychSBDV06}. For example, the latter study has shown a preponderance of the 0\ensuremath{\hbar\Omega}~$(0\, 4)$ symplectic irrep in $^{12}$C, which is indeed the irrep built over the most deformed 0\ensuremath{\hbar\Omega}~ bandhead, that is, the spin-zero $(0\, 4)$. \subsection{$^{20}$Ne and $A=20$ isotopes} We present calculations for the $g.st.$ rotational band of $^{20}$Ne together with the short-lived $^{20}$Mg at the proton-drip line (with no measured energy spectrum) and, its mirror nucleus, the neutron-rich $^{20}$O. They are indeed well described by the NCSpM in a $N_{\rm max}=12$ model space, where convergence of results is achieved (Fig. \ref{enSpectrumNe20} and Table \ref{Observables}). For $^{20}$Mg and $^{20}$O, the model space is down-selected to only one spin-zero symplectic irrep, $(4\,2)$, for $J^\pi=0^+, 2^+$, and $4^+$, with 1299 basis states (fixed $M$). For $^{20}$Ne, the model space consists of the spin-zero symplectic irrep, $(8\,0)$, for $J^\pi=0^+, 2^+, 4^+,$ and $6^+$, with 1070 basis states. All these irreps are built over the most deformed \ph{0} bandhead and expand up through $N_{\max}=12$. To show the significance of the symmetry-based selection and the important role of the most deformed 0\ensuremath{\hbar\Omega}~ bandhead together with the \SpR{3} excitations thereof, we consider a model space for $^{20}$Ne that consists of all symplectic irreps that start at 0\ensuremath{\hbar\Omega}. The resulting NCSpM energy spectrum is displayed in Fig. \ref{enSpectrumNe20full0hw} for $S=0$ and $S=1$. Indeed, no other $0^+$ state is found to lie below the $0^+$ of $(8\,0)$ for the $\gamma$ parameter used here ($S=2$ symplectic bandheads render states higher than 8-9 MeV). The $0^+$ of $(6\,1)$ is as much as 5 MeV above the $(8\,0)$, while all other $0^+$ states lie at $\sim 10$ MeV and higher. This indicates that the $(8\,0)$ irrep expanded up through $N_{\max}=12$ is indeed suitable for a reasonable description of the ground state of $^{20}$Ne. An important feature of the NCSpM is that it provides electric observables without the need for introducing effective charges. And while $g.st.$ rotational energies converge comparatively quickly, at $N_{\max} \sim 4$, we find that larger model spaces are needed to reproduce observables sensitive to enhanced collectivity (Fig. \ref{Ne20obsNmax}). For $N_{\max} \sim 4$, observables, such as the $B(E2)$ transition strengths, electric quadrupole moments, and matter rms radii have realized only 60\% of their total increase as compared to the $N_{\max}=0$ counterparts. Indeed, additional $40-50\%$ are needed for the $B(E2)$ strengths and $20\%$ for the $Q$ moments to obtain converged values. To reach convergence and to avoid the use of effective charges, at least $N_{\max} = 10$ is necessary, which is where results are also found to compare reasonably to experiment (Fig. \ref{enSpectrumNe20} and Table \ref{Observables}). This suggests that the model successfully reproduces observables that are informative of the state structure and the long-range behavior of the wavefunctions. \begin{figure}[th] (a)\\ \includegraphics[width=0.6 \columnwidth]{Fig/Ne20_BE2_gammaDependence.pdf} \\ (b) \\ \includegraphics[width=0.6 \columnwidth]{Fig/Ne20_en_gammaDependence.pdf} \\ \caption{ $\gamma$-Dependence of the $^{20}$Ne $g.st.$ rotational band: (a) $B(E2)$ transition strengths and (b) $2^+_1$ and $4^+_1$ energies. The grey vertical arrow indicates $\gamma=0.74\times 10^{-4}$, the value obtained in NCSpM analysis for $^{12}$C and used in the present calculations with no further adjustments. } \label{Ne20_gammaDependence} \end{figure} While the model parameter $\gamma$ has not been adjusted in the present study (but was significantly limited by the three lowest-lying $0^+$ states in $^{12}$C \cite{DreyfussLDDB13}), its value has a large effect upon observables under consideration. A typical dependence on this parameter for $sd$-shell nuclei is shown for $^{20}$Ne (Fig. \ref{Ne20_gammaDependence}). As the $\gamma$ value decreases for a given nucleus, thereby increasing the tendency of the high-\ensuremath{\hbar\Omega}~ excitations to become energetically favorable, the nucleus expands spatially and the $g.st.$ rotational band stretches energetically. This is accompanied by enhancement of collectivity and by considerably larger $B(E2)$ transition strengths. It is then remarkable that without adjusting $\gamma$ for $sd$-shell nuclei, energy spectra and other observables are found in a reasonable agreement with the experimental counterparts, as shown here for $A=20$ nuclei (Fig. \ref{enSpectrumNe20} and Table \ref{Observables}) and next for heavier isotopes. \begin{table}[t] \caption{NCSpM matter rms radii $r_m $ (fm) of the ground state and quadrupole moments $Q$ ($e$\,fm$^2$) of the $2^+$, $4^+$ and $6^+$ states of the $g.st.$ rotational band for the nuclei under consideration. Experimentally deduced matter radii are summarized in Ref. \cite{OzawaST01} and each of the original references is provided in the table; measured $Q$ moments are taken from Refs. \cite{Tilley98A20,Firestone05A22} for $A=20$ and $22$, respectively. } \begin{center} \begin{tabular}{llrrrr} \hline\hline && \hspace{0.2in}$r_m (0^+_{\rm gs})$ & \hspace{0.4in}$Q_{2^+_1}$ & \hspace{0.3in}$Q_{4^+_1}$ & \hspace{0.3in}$Q_{6^+_1}$ \\ \hline $^{20}$Mg & Expt. &$ 2.88(4)\tablenote{From Ref. \cite{OzawaST01ref28}} $& -- & -- & -- \\ & NCSpM &$ 2.73 $&$ -12.67 $&$ -16.67 $& -- \\ $^{20}$Ne & Expt. &$ 2.87(3)^a $&$ -23(3) $& -- & -- \\ & NCSpM &$ 2.79 $&$ -15.69 $&$ -19.69 $&$ -21.05 $\\ $^{20}$O & Expt. &$ 2.69(3)^a $& -- & -- & -- \\ & NCSpM &$ 2.73 $&$ -8.45 $&$ -11.11 $& --\\ $^{22}$Mg & Expt. &$ 2.89(6)\tablenote{From Ref. \cite{OzawaST01ref19}} $& -- & -- & -- \\ & NCSpM &$ 2.82 $&$ -17.88 $&$ -23.07 $&$ -25.93 $\\ $^{22}$Ne & Expt. & -- &$ -17(3) $& -- & -- \\ & NCSpM &$ 2.82 $&$ -14.90 $&$ -19.22 $&$ -21.61 $\\ $^{24}$Si & Expt. & -- & -- & -- & -- \\ & NCSpM &$ 2.40 $&$ -14.38 $&$ -18.18 $&$ -19.75 $\\ $^{24}$Ne & Expt. &$ 2.79(13)\tablenote{From Ref. \cite{OzawaST01}} $& -- & -- & -- \\ & NCSpM &$ 2.40 $&$ -10.27 $&$ -12.98 $&$ -14.11 $\\ \hline\hline \end{tabular} \end{center} \label{Observables} \end{table}% \subsection{$A=22$ and $24$} We perform NCSpM $N_{\max}=12$ calculations with no parameter adjustment (using $\ensuremath{\hbar\Omega}=15$ MeV and $\gamma =0.74\times 10^{-4}$) for $^{22}$Mg and $^{22}$Ne. For these nuclei, the $N_{\max}=12$ model space is down-selected to only one spin-zero symplectic irrep, $(8\,2)$, for $J^\pi=0^+, 2^+$, $4^+$ and $6^+$, with 2900 basis states (fixed $M$). Calculations for $^{22}$Mg and $^{22}$Ne yield energy spectra in close agreement with experiment (Fig. \ref{enA22}). In addition, most of the $B(E2)$ transitions strengths (Fig. \ref{enA22}) as well as electric quadrupole moments and matter rms radii (Table \ref{Observables}) predicted by the model fall within the experimental uncertainties where measurements or experimentally deduced values exist. \begin{figure}[th] \includegraphics[width=0.44\columnwidth]{Fig/enMg22_g00205.pdf} \includegraphics[width=0.44 \columnwidth]{Fig/enNe22_g00205.pdf} \caption{ NCSpM energy spectrum of $^{22}$Mg and its mirror nucleus $^{22}$Ne using the $S_p=S_n=S=0$ $55.5(8\, 2)$ \SpR{3} irrep built over the most deformed 0\ensuremath{\hbar\Omega}~ bandhead. Experimental data (``Expt.") is from \cite{Firestone05A22}. $B(E2)$ transition rates are in W.u. units. } \label{enA22} \end{figure} NCSpM $N_{\max}=12$ calculations (Fig. \ref{enA24}) are also presented here for the short-lived $^{24}$Si, even though a slightly larger $\ensuremath{\hbar\Omega}$ value, 21 MeV, is used only in this case. This nucleus is difficult to measure and knowledge on its structure, including spin-parity assignments and deformation of states, is necessary. For comparison, we also study the mirror nucleus $^{24}$Ne, for which richer experimental data is available. For these nuclei, the model space is down-selected to only one spin-zero symplectic irrep, $(10\,0)$, for $J^\pi=0^+, 2^+$, $4^+$ and $6^+$, with 1171 basis states (fixed $M$). The results, including energy spectra, $E2$ observables and radii, are found reasonable as compared to the available experiment (Fig. \ref{enA24} and Table \ref{Observables}). The $^{24}$Si wavefunctions, calculated by NCSpM and independent of the choice for $\ensuremath{\hbar\Omega}$, are found to be dominated by the $(10\,0)$ 0\ensuremath{\hbar\Omega}-, $(12\,0)$ 2\ensuremath{\hbar\Omega}- and $(14\,0)$ 4\ensuremath{\hbar\Omega}-configurations (Fig. \ref{Si24_probability}), thereby, as discussed in the following section, carrying considerably large prolate deformation. While there are 96 (or 274) basis states in the $(10\,0)$ symplectic irrep for the $0^+$ (or $2^+$) state, only a few of them contribute to the wavefunction at a level greater than 0.1\%, as shown in Fig. \ref{Si24_probability}. We note that the slightly smaller radius calculated by the model for $^{24}$Ne suggests that additional spin-zero and spin-one irreps besides the $(10\,0)$ vertical cone are likely to influence the low-energy dynamics. However, they are expected to remain of secondary importance to $(10\,0)$. \begin{figure}[th] \includegraphics[width=0.44 \columnwidth]{Fig/enSi24_g00205hw21.pdf} \includegraphics[width=0.44\columnwidth]{Fig/enNe24_g00205hw21.pdf} \caption{ NCSpM energy spectrum of $^{24}$Si and its mirror nucleus $^{24}$Ne using the $S_p=S_n=S=0$ $62.5(10\, 0)$ \SpR{3} irrep built over the most deformed 0\ensuremath{\hbar\Omega}~ bandhead. Experimental data (``Expt.") is from \cite{Firestone07A24}. $B(E2)$ transition rates are in W.u. units. } \label{enA24} \end{figure} \begin{figure}[th] \includegraphics[width=\columnwidth]{Fig/Si24Prblty_J0_J2.pdf} \caption{ Probability distribution (in \%) for the $0^+$ ground state and the lowest $2^+$ state of $^{24}$Si ($^{24}$Ne). Only contributions greater than 0.1\% are shown. } \label{Si24_probability} \end{figure} \section{Dominant deformed configuration} Examinations of the \SU{3} content of the NCSpM wavefunctions bring forward important information on deformation and associated dominant configurations through the deformation-related $(\lambda_\omega \,\mu_ \omega)$. This is based on the mapping \cite{CastanosDL88} between the shell-model $(\lambda \,\mu)$ \SU{3} labels (microscopic) and the shape variables of the Bohr-Mottelson collective model \cite{BohrMottelson69}, which provides a description of the nuclear surface in terms of the elongation $\beta>0$ and the $0 \le \gamma \le \pi/2$ asymmetry parameter. Specifically, in the limit of large deformation, $(\lambda\ 0)$ and $(0\ \mu)$ can be associated with a prolate ($\gamma = 0 ^\circ $) and oblate ($\gamma = 60 ^\circ$) shapes, respectively, while larger $\lambda$ ($\mu$) values are linked to larger deformation, $\beta$. \begin{figure}[th] \includegraphics[width=0.64\columnwidth]{Fig/20Ne_probability.pdf} \includegraphics[height=1.5 in]{Fig/22Ne_probability.pdf} \caption{ NCSpM probability distribution (specified by the area of the circles) for the ground state of (a) $^{20}$Ne using the $(8\,0)$ \SpR{3} irrep, and (b) $^{22}$Ne using the $(8\,2)$ \SpR{3} irrep. The symplectic states are grouped according to their $(\lambda_\omega \,\mu_\omega)$ \SU{3} symmetry, which are mapped onto the $(\beta \,\gamma)$ shape variables of the collective model (see text for further details). } \label{probability} \end{figure} It is then clear that the nuclei under consideration are highly deformed and prolate in their low-lying states, as manifested in Fig. \ref{probability} for selected nuclei (this feature is the least pronounced for $^{20}$O and $^{20}$Mg). This is also confirmed by the negative large electric quadruple moments (Table \ref{Observables}) and enhanced $B(E2)$ values. Fig. \ref{probability} further reveals that the most dominant modes are observed among the ones with high $\lambda_\omega$ and low $\mu_ \omega$. Configurations in higher-\ensuremath{\hbar\Omega}~ model spaces tend to increase the deformation (larger $\beta$) and decrease nuclear triaxiality (smaller $\gamma$) as compared to the predominant 0\ensuremath{\hbar\Omega}~ configuration (the bandhead). These high-\ensuremath{\hbar\Omega}~ configurations ($N_n =6$ and beyond), while contributing only slightly to the wavefunctions, bring in substantially large deformation, thereby becoming critical for the convergence of the observables shown in Fig. \ref{Ne20obsNmax}. \section{Conclusion} We carried forward a no-core symplectic NCSpM study with a schematic long-range many-nucleon interaction that showed how highly deformed structures in intermediate-mass nuclei emerged out of a no-core shell-model framework. While previously the NCSpM has been successfully employed in a study of the $\alpha$-cluster driven Hoyle-state rotational band in $^{12}$C \cite{DreyfussLDDB13}, which has fixed the only adjustable parameter in the schematic interaction, here we show that the framework is extensible to low-lying states of other nuclei without any parameter adjustment. We focused on the $g.st.$ rotational band of $^{20}$Ne (a nucleus multiple of an $\alpha$ particle), as well as of $^{22,24}$Ne, $^{20}$O and of the proton-rich $^{20,22}$Mg, and $^{24}$Si nuclei. By utilizing the symplectic symmetry, we were able to accommodate model spaces up through 15 major HO shells and hence, to take into account essential high-\ensuremath{\hbar\Omega}~particle excitations. These excitations were found key to the description of large deformation and the convergence of electric observables without effective charges. These configurations were included in the shell-model space by considering only one symplectic irrep (vertical cone) built on the most deformed spin-zero bandhead and extended to $N_{\rm max}=12$. This further confirms the dominance of low-spin/high-deformation and the importance of the symplectic symmetry to the low-energy nuclear dynamics. Most importantly, the NCSpM has allowed us to identify, from a no-core shell-model perspective, components of the inter-nucleon interaction and type of particle excitations that appear foremost responsible for unveiling the primary physics governing highly-deformed structures, starting with rotational and alpha-clustering features in the case of $^{12}$C \cite{DreyfussLDDB13} and $^{8}$Be \cite{LauneyDDTFLDMVB12}, but also expanding to the region of the lower $sd$ shell ($A\le24$). Therefore, the NCSpM appears as a useful tool to inform properties of the inter-nucleon interaction and to suggest efficacious shell-model truncation strategies to be employed in {\it ab initio} studies. \newline \acknowledgements We thank George Rosensteel and David Rowe for useful discussions. This work was supported by the U.S. NSF (OCI-0904874), the U.S. DOE (DE-SC0005248), and SURA, and in part by U.S. DOE (DE-FG02-95ER-40934). ACD acknowledges support by the U.S. NSF (grant 1004822) through the REU Site in Dept. of Physics \& Astronomy at LSU. \newline
\section{Introduction} \labbel{intro} It is well-known that many covering properties, which in general are distinct, turn out to be equivalent for \emph{linearly ordered topological spaces} (henceforth, LOTS, for short). For example, a LOTS is pseudocompact if and only if it is countably compact, if and only if it is sequentially compact. See Gulden, Fleischman and Weston \cite{GFW} and Purisch \cite{purisch} for these and more general results, also involving uncountable cardinals and dealing with product theorems, as well. Most of the above results generalize to \emph{GO spaces}, short for \emph{generalized ordered topological spaces}; see \cite{purisch} and \cite{goic}, or below. Here we show that, for every ultrafilter $D$, $D$-compactness and $D$-\brfrt pseudocompactness are equivalent in GO spaces (Remark \ref{iff}, or Theorem \ref{thm}(1) $\Leftrightarrow $ (5)). The particular case when $D$ is an ultrafilter over $ \omega$ is due to Garc{\'{\i}}a-Ferreira and Sanchis \cite{GFS}, whose paper contains many other related theorems. More generally, we characterize $D$-converging sequences in GO spaces (Proposition \ref{single}(7)-(9)). We also show that the $D$-compactness of some GO space $X$ depends exclusively on the ``decomposability spectrum'' of $D$ and on the cardinal types of (pseudo-)gaps in $X$, equivalently, on the existence of non converging strictly monotone sequences of certain cardinal order types (Theorem \ref{thm}). Some of the results mentioned in the first paragraph can be obtained as corollaries of Theorem \ref{thm}. For example, we get a proof that any product of initially $\lambda$-compact GO spaces is still initially $\lambda$-compact (Corollary \ref{initial}). The result is quite surprising, since noncompact covering properties are usually not preserved by taking products. Just to state a significant example, the square of a Lindel\"of LOTS is not necessarily Lindel\"of. The Sorgenfrey line is an example which is a GO space. To get an example which is a LOTS, consider $\mathbb R \times \omega $ with the lexicographic order. Also, it is well-known that the product of two countably compact topological spaces is not necessarily countably compact, and many similar counterexamples are known for initial $\kappa$-compactness, when $\kappa$ is regular. See, e.~g., Stephenson \cite{ste84}, Vaughan \cite{vau84}, Nyikos and Vaughan \cite{NV} and more references there. Hence, for $\kappa$ regular, preservation of initial $\kappa$-compactness under products is really a special property of GO spaces. In passing, let us mention that, on the other hand, when $\kappa$ is a singular strong limit cardinal, every product of initially $\kappa$-compact topological spaces is still initially $\kappa$-compact, by the celebrated Stephenson and Vaughan Theorem 1 in \cite{sv74}. In the particular case of LOTS, preservation of initial $\kappa$-compactness under products had originally been proved in \cite{GFW}, for every infinite cardinal $\kappa$. Not aware of Gulden, Fleischman and Weston's result, we originally have found a proof which holds for GO spaces, too, and which uses ultrafilter convergence \cite{goic}. Later we realized that the theorem can also be proved by adapting some arguments from \cite{GFW} (see the second proof of Corollary \ref{initial} here), hence the use of ultrafilters becomes unnecessary. However, luckily for scholars fond of ultrafilters, there are indeed compactness properties whose preservation under products (in GO spaces) involves ultrafilters in an essential way. For example, this is the case when considering simultaneously countable compactness and $[ \lambda , \lambda ]$-\brfrt compactness. Considering preservation of such properties leads to statements independent from the usual axioms of set theory, statements which involve the existence of ultrafilters with an ``unusual'' descending completeness spectrum (equivalently, decomposability spectrum). See Corollary \ref{gocorbb} for the general statement, and Corollary \ref{gocorbbb} for an explicit example. Now a few words about the fate of \cite{goic}. Since all the results proved there are subsumed by the present paper, we are not going to submit \cite{goic} elsewhere. However, we shall keep it available in archived form, since it might be useful for those looking for a direct proof of Corollaries \ref{initial} and \ref{cor} here, which essentially were the main theorems of \cite{goic}. \section{Preliminaries} \labbel{prel} We now recall the relevant definitions. A \emph{LOTS} is a linearly ordered set endowed with the open interval topology. A \emph{GO space} is a linearly ordered set with a $T_2$ topology having a base of order-convex sets. GO spaces are exactly spaces which can be obtained as subspaces of LOTS; the notion is more general, since if $X$ is a LOTS, and $Y \subseteq X$, then the subset topology on $Y$ induced by the topology on $X$ might be finer than the order topology on $Y$ relative to the restriction of the order on $X$. See, e.~g., \cite{BL} for more informations and references about LOTS and GO spaces. If $D$ is an ultrafilter over some set $I$, then a topological space $X$ is said to be \emph{$D$-compact} if every $I$-indexed sequence $(x_i) _{i \in I} $ of elements of $X$ \emph{$D$-converges} to some $x \in X$, that is, $\{ i \in I \mid x_i \in U\} \in D$, for every open neighborhood $U$ of $x$. The notion of ultrafilter convergence has proved particularly useful in the study of compactness properties of topological spaces, in particular with respect to preservation under products. Classical papers on the subject are Bernstein \cite{ber70}, Ginsburg and Saks \cite{GS}, and Saks \cite{sak78}. Excellent surveys of results proved until the mid '80's are Vaughan \cite{vau84} and Stephenson \cite{ste84}. Further results can be found in Caicedo \cite{Ca} and \cite{sssr}, together with additional references. Ginsburg and Saks \cite{GS} made also a very effective use of the notion of a $D$-limit point of a sequence of subsets of a topological space, and introduced the notion of $D$-pseudocompactness. See also \cite{gf99}. If $(Y_i) _{i \in I} $ is a sequence of subsets of $X$, a point $x \in X$ is said to be a \emph{$D$-limit point} of $(Y_i) _{i \in I} $ if $ \{ i \in I \mid U \cap Y_i \not= \emptyset \} \in D $, for every neighborhood $U$ of $x$ (thus when each $Y_i$ is a singleton we get back the notion of $D$-convergence). A topological space $X$ is \emph{$D$-pseudocompact} if every sequence $(O_i) _{i \in I} $ of nonempty open sets of $X$ has a $D$-limit point. Clearly, we can equivalently relax the assumption that \emph{all} the $O_i$'s are nonempty to the assumption that $ \{ i \in I \mid O_i \not= \emptyset \} \in D $, since we can change those $O_i$'s in a set not in $D$ without affecting $D$-limit points of the sequence (since $D$ is a filter). Consider the following statement. (*) $A$ and $B$ are open subsets of a GO space $X$, $A \cup B = X$, and $a < b$, for every $a \in A$ and $b \in B$. An ordered pair $(A, B)$ satisfying (*) is called a \emph{gap} of $X$ in case that neither $A$ has a maximum, nor $B$ has a minimum. Here we are including \emph{end gaps} in the definition of a gap, that is, we allow either $A$ or $B$ to be empty. An ordered pair $(A, B)$ satisfying (*) is called a \emph{pseudo-gap} in case that both $A$ and $B$ are nonempty, and either $A$ has a maximum and $B$ has no minimum, or $A$ has no maximum and $B$ has a minimum. Clearly pseudo-gaps can occur only in GO spaces which are not LOTS, since we are asking that $A$ and $B$ are open. For a property $\mathcal P$, the expression ``$X$ has no (pseudo-)gap satisfying $\mathcal P$'' shall be used as an abbreviation for ``$X$ has neither a gap nor a pseudo-gap satisfying $\mathcal P$''. \section{$D$-convergence of a given sequence} \labbel{sing} \begin{proposition} \labbel{single} Suppose that $X$ is a linearly ordered set, $D$ is an ultrafilter over some set $I$, and $(x_i) _{i \in I} $ is a sequence of elements of $X$. Set $A= \{ x\in X \mid \{ i \in I \mid x < x_i \}\in D \}$, and $B= \{ x\in X \mid \{ i \in I \mid x_i < x \}\in D \}$. Then the following statements hold. \begin{enumerate} \item $x \in X \setminus (A \cup B) $ if and only if $\{ i \in I \mid x_i = x \} \in D$. \item $|X \setminus (A \cup B) | \leq 1$. \item If $ x \in X \setminus (A \cup B) $, then $a < x$, for every $a \in A$, and $x < b$, for every $b \in B$. \item If $a \in A$ and $b \in B$, then $ \{ i \in I \mid x_i \in (a, b) \} $ belongs to $ D$, hence it is not empty. \item If $a \in A$ and $b \in B$, then $a < b$. \item If $X= A \cup B$, then either $A$ has no maximum or $B$ has no minimum. \item If $X$ is a GO space, then $(x_i) _{i \in I} $ $D$-converges to $x \in X$ if and only if one of the following (mutually exclusive) conditions holds. \begin{enumerate} \item $\{ i \in I \mid x_i = x \} \in D$, or \item $x$ is the maximum of $A$ and $x \in \overline{B}$, or \item $x$ is the minimum of $B$ and $x \in \overline{A}$. \end{enumerate} \item If $X$ is a LOTS, then $(x_i) _{i \in I} $ $D$-converges to $x \in X$ if and only if either \begin{enumerate} \item $\{ i \in I \mid x_i = x \} \in D$, or \item $x$ is the maximum of $A$ and $A \cup B = X$, or \item $x$ is the minimum of $B$ and $A \cup B = X$. \end{enumerate} \item If $X$ is a GO space, then $(x_i) _{i \in I} $ $D$-converges to some $x \in X$ if and only if $(A,B)$ is neither a gap, nor a pseudo-gap of $X$. In particular, if $X$ is a LOTS, then $(x_i) _{i \in I} $ $D$-converges in $X$ if and only if $(A,B)$ is not a gap. \item Suppose that $X$ is a GO space and that $X=A \cup B$. For $i \in I$, put \begin{equation*} O_i=\begin{cases} \bigcup _{a \in A} (x_i, a) & \text{if $x_i \in A$}, \\ \bigcup _{b \in B} (b, x_i) & \text{if $x_i \in B$}.\\ \end{cases} \end{equation*} Then $(x_i) _{i \in I} $ $D$-converges to $x $ if and only if $x$ is a $D$-limit point of $(O_i) _{i \in I} $. \end{enumerate} \end{proposition} \begin{proof} (1) Since $D$ is an ultrafilter, it follows that, for each $x \in X$, one and exactly one of the following sets: $ \{ i \in I \mid x < x_i \}$, $\{ i \in I \mid x_i < x \}$, $\{ i \in I \mid x_i = x \}$ belongs to $D$, since they are disjoint and their union is $I$. Hence, by the definitions, $x\not\in A \cup B$ if and only if the last eventuality occurs. (2) If $x,y\not\in A \cup B$, then, by (1), both $X=\{ i \in I \mid x_i = x \} $ and $Y=\{ i \in I \mid x_i = y \} $ are in $ D$. Since $D$ is a filter, $ X \cap Y \in D$, and, since $D$ is proper, $ X \cap Y \not= \emptyset $. If $ i \in X \cap Y $, then $x = x_i = y$. (3) If $a \in A$ and $ x \not\in (A \cup B) $, then both $ \{ i \in I \mid a < x_i \}$ and $\{ i \in I \mid x_i = x \}$ belong to $D$, by (1). As above, their intersection is not empty, and this implies $a < x$. In the same way, we get that $x < b$, for every $b \in B$. (4) If $a \in A$ and $b \in B$, then $ \{ i \in I \mid a < x_i \}$ and $\{ i \in I \mid x_i < b \}$ belong to $D$. The intersection of the above two sets is $ \{ i \in I \mid x_i \in (a, b) \}$, therefore this set belongs to $D$, hence is nonempty. (5) is immediate from (4). (6) By contradiction, if $x$ is the maximum of $A$ and $y$ is the minimum of $B$, then, by (5) and since $X=A \cup B$, $y$ is the immediate successor of $x$. Both $\{ i \in I \mid x< x_i \}$ and $\{ i \in I \mid x_i< y \}$ belong to $D$, since $x \in A$ and $y \in B$. But the above two sets are disjoint, a contradiction. (7) Suppose that $(x_i) _{i \in I} $ $D$-converges to $x \in X$. By (1), (2) and the uniqueness of $D$-limits in Hausdorff spaces, either $ x \not\in A \cup B$ and (a) holds, or $X= A \cup B$. In this latter case, either $x \in A$ or $x \in B$. Suppose that $ x \in A$. If $x$ is not the maximum of $A$, there is $x' \in A$ such that $x < x'$. Then $( -\infty, x')$ is a neighborhood of $x$, and $\{ i \in I \mid x' < x_i\} \in D$, since $x' \in A$. But then $\{ i \in I \mid x_i \in ( -\infty, x') \} \not\in D$, contradicting $D$-convergence. Hence $x$ is the maximum of $A$. If $x \not\in \overline{B}$, then, by (5) and since $A \cup B=X$, $(- \infty, x]$ is a neighborhood of $x$, hence, by $D$-convergence, $\{ i \in I \mid x_i \leq x\} \in D$, but this contradicts $x\in A$. Next, notice that if $x \in A \cap \overline{B}$, then necessarily $x$ is the maximum of $A$, because of (5). Hence $x \in \overline{B}$. Symmetrically, if $x \in B$, then $x$ is the minimum of $B$ and $x \in \overline{A}$. Hence one among the conditions (a)-(c) holds. Conversely, if (a) holds, then trivially $(x_i) _{i \in I} $ $D$-converges to $x$. Suppose that (b) holds. Since $x \in A$, then, by (4), $ \{ i \in I \mid x < x_i < b\} \in D$, for every $b \in B$. Since $x \in \overline{B}$. then, for every neighborhood $U$ of $x$, there is some $b \in B$ such that $U \supseteq [x,b)$, but then $ \{ i \in I \mid x_i \in U\} \supseteq \{ i \in I \mid x < x_i < b\} \in D$, thus $(x_i) _{i \in I} $ $D$-converges to $x$. Case (c) is proved in a symmetrical way. Notice that (b) and (c) are mutually exclusive, since $A$ and $B$ are disjoint, by (5). Conditions (b) and (c) are also both mutually exclusive with (a) by (1). (8) Since every LOTS is, in particular, a GO space, then Condition (7) applies. Hence it is enough to prove that, say, in every LOTS (8)(b) implies (7)(b). Indeed, if (8)(b) holds, then, since $A \cup B=X$, by (4) and (6), $x$ is an infimum of $B$, but in a LOTS this implies $x \in \overline{B}$, thus (7)(b) holds (notice that, by the very definition of $A$, if $A$ has a maximum, then $X \setminus A=B$ is not empty). (9) We use the characterization given in (7). If one of (a), (b) or (c) in (7) holds, then trivially $(A,B)$ is neither a gap nor a pseudo-gap. Conversely, suppose that $(A,B)$ is not a (pseudo-)gap. If $A \cup B \not = X$, then (7)(a) holds, for some $x \in X$, by (1). Otherwise, $A \cup B = X$. Since we are including end gaps in our definition of a gap, both $A$ and $B$ are nonempty. By (5), and since $A \cup B = X$, if neither $A$ has a maximum, nor $B$ has a minimum, then both $A$ and $B$ are open, contradicting the assumption that $(A, B)$ is not a gap. Assume that, say, $A$ has a maximum $x$. Since $A \cup B = X$, then $x$ has no immediate successor, thus $B$ has no minimum, since $x$ is the maximum of $A$. If $x \not\in \overline{B}$, then $B$ would be clopen, and $(A, B)$ would be a pseudo-gap. Thus $x \in \overline{B}$ and 6(b) holds. (10) Suppose that $(x_i) _{i \in I} $ $D$-converges to $x $. By (1), we are either in case (7)(b) or (7)(c). Suppose, say, that we are in case (7)(b). Since $x \in \overline{B}$, then every neighborhood $U$ of $x$ contains an interval $[x,b']$, for some $b' \in B$. For every $i \in I$ such that $x_i \in B$, there is $b \in B$ such that $b > \sup \{ b', x_i \} $, by (6). Thus $b' \in [x, b'] \cap (b, x_i) \subseteq U \cap O_i$, hence $U \cap O_i \not= \emptyset $. Since $x$ is the maximum of $A$, then, by (4), $\{ i \in I \mid x_i \in B \} = \{ i \in I \mid x < x_i \} \in D$, thus $x$ is a $D$-limit point of $(O_i) _{i \in I} $. Conversely, suppose that $X= A \cup B$, and that $x$ is a $D$-limit point of $(O_i) _{i \in I} $. Since $X= A \cup B$, then either $\{ i \in I \mid x_i \in A \} \in D$ or $\{ i \in I \mid x_i \in B \} \in D$, say, the latter eventuality occurs. We first show that $x \not \in B$. Indeed, if $x \in B$, then $\{ i \in I \mid x_i < x \} \in D$ and, since $\{ i \in I \mid x_i \in B \} \in D$, then $\{ i \in I \mid x_i \in B \cap (- \infty, x) \} \in D$. In particular, the above set is not empty, hence there is $x' < x$ such that $x' \in B$. Arguing in the same way, $\{ i \in I \mid x_i \in B \cap (- \infty, x') \} \in D$. But then, letting $U= (x', \infty)$, we have that $U$ is a neighborhood of $x$ such that $\{ i \in I \mid U \cap O_i = \emptyset \}\in D$, since $ O_i \subseteq (- \infty, x') $, whenever $x_i \in B \cap (- \infty, x')$. This contradicts the assumption that $x$ is a $D$-limit point of $(O_i) _{i \in I} $, hence $x \not \in B$. Thus $x \in A$. Moreover, $x$ is the maximum of $A$. If not, there is some $x' \in A$ such that $x< x'$. Then $U=(-\infty, x')$ is a neighborhood of $x$ such that $U \cap O_i = \emptyset$, for a set of indices in $D$, namely, for $\{ i \in I \mid x_i \in B \}$, by (5), again contradicting the assumption that $x$ is a $D$-limit point of $(O_i) _{i \in I} $. In view of (7)(b), in order to finish the proof, it is enough to show that $x \in \overline{B}$. If not, $U=(-\infty, x]$ is a neighborhood of $x$, and we can get a contradiction arguing as before, since by (5) $A$ and $B$ are disjoint, $A=U=(-\infty, x]$, and $ O_i \subseteq B$, whenever $x_i \in B$. \end{proof} \begin{remark} \labbel{iff} From Proposition \ref{single}(10) it follows easily that, for every ultrafilter $D$ and every GO space $X$, $D$-compactness of $X$ is equivalent to $D$-\brfrt pseudocompactness of $X$. The only-if part is trivial. For the converse, suppose that $X$ is $D$-\brfrt pseudocompact, and that $(x_i) _{i \in I} $ is a sequence of elements of $X$. If $X \not = A \cup B$, then $(x_i) _{i \in I} $ $D$-converges, by \ref{single}(1). Otherwise, define $(O_i) _{i \in I} $ as in (10). It is enough to show that $ \{ i \in I \mid O_i \not= \emptyset \} \in D $, since then we can apply $D$-pseudocompactness to $(O_i) _{i \in I} $ and use (10). Since $X= A \cup B$, then either $\{ i \in I \mid x_i \in A \} \in D$, or $\{ i \in I \mid x_i \in B \} \in D$. Say the latter case occurs, then, by the definition, $B$ has no minimum, hence, for every $i \in I$ such that $x_i \in B$, we have $O_i= B \cap (- \infty, x_i) \not= \emptyset $. Hence the $O_i$'s are nonempty for a set of indices in $D$, namely, a set containing $\{ i \in I \mid x_i \in B \}$. \end{remark} However, much more can be said about the connections among $D$-compactness, $D$-pseudocompactness and other properties of a GO space $X$. This is the main theme of the next section. \section{$D$-compactness, $D$-pseudocompactness, gaps, products, etc.} \labbel{secth} To state the results of the present section in their full generality, we need some more definitions. An ultrafilter $D$ over some set $I$ is \emph{$\lambda$-descendingly complete} if every $ \subseteq $-decreasing sequence $(Z _ \alpha ) _{ \alpha \in \lambda } $ of sets in $D$ has intersection still in $D$. The ultrafilter $D$ is \emph{$\lambda$-decomposable} if there is a function $f:I \to \lambda $ such that $f ^{-1}(X) \not \in D $, for every $X \subseteq \lambda $ such that $|X| < \lambda $. Clearly, if $\lambda$ is an infinite cardinal, a $\lambda$-decomposable ultrafilter is not $\lambda$-descendingly complete: just consider $Z _ \alpha = f ^{-1}([ \alpha , \lambda )) $. If $\lambda$ is regular, the converse holds. If $\lambda$ is regular and $D$ is not $\lambda$-descendingly complete, as witnessed by $(Z _ \alpha ) _{ \alpha \in \lambda } $, the following function $f$ witnesses that $D$ is $\lambda$-decomposable. Define $f$ by letting $f(i)$ be the smallest $\alpha$ such that $i \not\in Z _ \alpha $, if such an $\alpha< \lambda $ exists; $f(i)$ is unimportant and can be arbitrary if $ i \in \bigcap _{ \alpha \in \lambda } Z _ \alpha $, since we are assuming that $ \bigcap _{ \alpha \in \lambda } Z _ \alpha \not \in D$. Thus, for infinite regular cardinals, $\lambda$-decomposability is equivalent to the negation of $\lambda$-descending completeness. Let us also mention that there is another notion equivalent to $\lambda$-\brfrt decomposability, for $\lambda$ an infinite regular cardinal, that is, \emph{$ ( \lambda , \lambda )$-regularity}. We shall make only a limited use of $ ( \lambda , \lambda )$-regularity here (see Sections \ref{secinit} and \ref{setth}), but we should warn the reader that theorems about $\lambda$-descending (in)completeness or $\lambda$-decomposability are frequently stated in equivalent forms in terms of regularity, or vice versa. A full discussion is given in \cite{mru}; see Section 1 there and, in particular, Properties 1.1(xi) and Consequence 1.2. For an ultrafilter $D$, let $K_D= \{ \nu \geq \omega \mid D \text{ is $\nu$-decomposable}\} $. Many problems are still open about the possible values $K_D$ can assume, and solutions to such problems are heavily dependent on the universe of set theory one is working in. For example, in certain models of set theory, $K_D$ is always an interval of cardinals, with bottom element $ \omega$; on the other hand, there are models in which $K_D$ is a rather sparse set of cardinals. We refer to the comments after Problem 6.8 in \cite{mru} for further information and details, or to the remarks before Corollary \ref{gocorbbb} here, where we show that such set theoretical problems affect the behavior under products of certain compactness properties of GO spaces. In particular, certain statements turn out to be independent from the usual axioms for set theory. In this paper we shall be mainly concerned with \emph{regular} cardinals in $K_D$, hence we shall establish the special notation $K_D ^{Reg}$ for the set of such cardinals. Namely, $K_D^{Reg}= \{ \nu \geq \omega \mid \nu \text{ is regular and } D \text{ is $\nu$-decomposable}\} $. A topological space is \emph{$[ \nu, \lambda ]$-compact} if every open cover of cardinality $\leq \lambda $ has a subcover of cardinality $<\nu$. It is a classical result by Alexandroff and Urysohn \cite{AU} that if $\nu$ is a regular cardinal, then $[ \nu, \nu ]$-compactness of a topological space $X$ is equivalent to $\CAP_ \nu$, which is the property asserting that every subset of $X$ of cardinality $\nu$ has a complete accumulation point. \emph{Initial $\lambda$-compactness} is $[ \omega , \lambda ]$-compactness, and, again by \cite{AU}, it is equivalent to $[ \nu, \nu ]$-compactness, for every cardinal $\nu \leq \lambda $, equivalently, for every \emph{regular} cardinal $\nu \leq \lambda$. As usual, by $[ \nu, \lambda ]$ we shall also denote the \emph{interval} of all cardinals $\mu$ such that $\nu\leq \mu \leq \lambda$. We hope that the partially overlapping notation will cause no confusion. A topological space is \emph{weakly $[ \nu, \lambda ]$-compact} if every open cover of cardinality $\leq \lambda $ has a subfamily of cardinality $<\nu$ with dense union. \emph{Weak initial $\lambda$-compactness} is weak $[ \omega , \lambda ]$-compactness. The above notions are frequently named using very disparate terminology. Weak $[ \nu, \lambda ]$-compactness has sometimes been called \emph{weak $\lambda$-$\nu$ compactness}, and we used still different terminology in \cite{tproc2}, where we called it $\mathcal O$-$[ \nu, \lambda ]$-compactness. When $\nu$ is a regular cardinal, weak $[ \nu, \nu ]$-compactness is equivalent to a notion called \emph{pseudo-$( \kappa , \kappa )$-compactness}. Weak initial $ \omega $-compactness is equivalent to \emph{faint compactness}, and equivalent to pseudocompactness in the class of Tychonoff spaces. Weak initial $\lambda$-compactness has been called \emph{almost $\lambda$-compactness}, or \emph{weak-$\lambda$-$\aleph_0$-compactness} by some authors. See \cite[Remark 3]{tapp2} for further details and references. If $ \lambda $ is an infinite cardinal, a sequence $(x_ \gamma ) _{ \gamma < \lambda } $ of elements of a topological space \emph{converges} to some point $x$ if, for every neighborhood $U$ of $x$, there is $\gamma < \lambda $ such that $x _{ \gamma '} \in U $, for every $\gamma' > \gamma $. If $(A,B)$ is a gap of some GO space, we say that that \emph{$\lambda$ is a type of} $(A,B)$ if either $A$ has cofinality $\lambda$, or $B$ has coinitiality $\lambda$ (thus a gap has at most two types). If $(A,B)$ is a pseudo-gap, say, $B$ having a minimum, the \emph{type of} $(A,B)$ is the cofinality of $A$, and, symmetrically, the coinitiality of $B$, if $A$ has a maximum. The above two notions are related as follows. If $(x_ \gamma ) _{ \gamma < \lambda } $ is a non converging strictly increasing sequence of elements of a GO space $X$, then, letting $A = \{ x \in X \mid x < x_ \gamma \text{ for some } \gamma < \lambda \}$, and $B=X \setminus A$, we have that $(A,B)$ is a (pseudo-)gap having type $\lambda$. A symmetrical statement holds for strictly decreasing sequences. Conversely, given a (pseudo-)gap $(A,B)$ having type $\lambda$, we obtain a strictly monotone non converging sequence $(x_ \gamma ) _{ \gamma < \lambda } $, by considering either a cofinal subset of $A$ or a coinitial subset of $B$. \begin{theorem} \labbel{thm} Suppose that $D$ is an ultrafilter over some set $I$, and recall that $K_D ^{Reg}$ is the set of those infinite regular cardinals $\nu$ such that $D$ is $\nu$-decomposable. For every GO space $X$, the following conditions are equivalent. \begin{enumerate} \item $X$ is $D$-compact. \item $X$ is $[ \nu, \nu]$-compact, for every cardinal $\nu \in K_D ^{Reg}$. \item For every cardinal $ \nu \in K_D ^{Reg}$, and every strictly increasing (resp., strictly decreasing) $\nu$-indexed sequence of elements of $X$, the sequence has a supremum (resp., an infimum) to which it converges. \item $X$ has no (pseudo-)gap of type belonging to $K_D ^{Reg}$. \item $X$ is $D$-pseudocompact. \item $X$ is weakly $[ \nu, \nu]$-compact, for every cardinal $\nu \in K_D ^{Reg}$. \item For every cardinal $ \nu \in K_D ^{Reg} $, and every sequence $(O_ \gamma ) _{\gamma < \nu}$ of open nonempty sets of $X$, there is $x \in X$ such that $|\{\gamma < \nu \mid O_ \gamma \cap U \not= \emptyset \}|= \nu$, for every neighborhood $U$ of $x$. In the above condition we can equivalently require either that the $O_ \gamma $'s are pairwise disjoint, or that $O_ \gamma \subset O _{ \gamma '} $, for $\gamma> \gamma '$. \item $X$ is $D'$-compact, for every ultrafilter $D'$ such that $K _{D'} ^{Reg} \subseteq K_D ^{Reg}$. \item $X$ is $D'$-pseudocompact, for every ultrafilter $D'$ such that $K _{D'}^{Reg} \subseteq K_D ^{Reg}$. \end{enumerate} \end{theorem} \begin{proof} (1) $\Rightarrow $ (2) It is enough to show that if $\nu$ is regular and $D$ is $\nu$-decomposable, then every $D$-compact topological space is $[ \nu, \nu]$-compact. This is essentially a classical result (which holds for every topological space), due to \cite{GS} in the case $\nu= \omega $, and to \cite{sak78} in the general case. See also \cite{Ca} and \cite{tproc2} for other versions and generalizations. Briefly, and using the equivalent reformulation of $[ \nu, \nu]$-compactness in terms of complete accumulation points, if $Y \subseteq X$ and $|Y |= \nu$, enumerate $Y$ as $\{ y _ \gamma \mid \gamma \in \nu\}$. If $f:I \to \nu$ witnesses the $\nu$-decomposability of $D$, then, by $D$-compactness, the sequence $(y _{f(i)} ) _{i \in I} $ has some $D$-limit point, which is easily seen to be also a complete accumulation point of $Y$. (2) $\Rightarrow $ (3) is almost immediate. Indeed, if, by contradiction, $(x_ \gamma ) _{ \gamma < \nu} $ is, say, strictly increasing, then $ O = X \setminus \bigcup _{\gamma < \nu} (- \infty, x_ \gamma ) $ is open. Then the family containing $ O$ and $(- \infty, x_ \gamma ) $, for $\gamma < \nu$, is an open cover by $\nu$ sets, but no subfamily by $<\nu$ sets is a cover, since $\nu$ is regular. The equivalence of (3) and (4) should be clear from the remarks shortly before the statement of the theorem. The proof of (4) $\Rightarrow $ (1) is the key argument in the proof of the theorem, and uses in an essential way the assumption that we are in a GO space. Suppose by contradiction that (1) fails and let $(x_i) _{i \in I} $ be a sequence which does not $D$-converge. Let $A$ and $B$ be defined as in the statement of Proposition \ref{single}. By condition (9) in the same proposition, $(A,B)$ is either a gap or a pseudo-gap of $X$, in particular, $A \cup B =X$. Hence either $\{i \in I \mid x_i \in A\} \in D$, or $\{i \in I \mid x_i \in B\} \in D$. Suppose, say, that the latter occurs. Then by the assumption that $\{i \in I \mid x_i \in B\} \in D$, by the definition of $B$, and since $D$ is a filter, we get that, for every $b \in B$, $X_b=\{i \in I \mid x_i \in (-\infty, b) \cap B\} \in D$. Notice that it follows that $B$ has no minimum, since each $X_b$ is nonempty. Choose a strictly decreasing coinitial sequence $(b_ \gamma ) _{ \gamma \in \nu} $ in $B$. Then $\nu$ is a type of the (pseudo-)gap $(A,B)$, and $ (X _{b_ \gamma }) _{ \gamma \in \nu} $ witnesses that $D$ is not $\nu$-descendingly complete, since $ \bigcap _{ \gamma \in \nu} X _{b_ \gamma }= \{i \in I \mid x_i \not\in B\} \not\in D $. Thus $\nu\in K_D ^{Reg}$, by the equivalence mentioned after the definition of decomposability, and since $\nu$ is regular, being the coinitiality of $B$. Thus $(A,B)$ has type $\nu \in K_D ^{Reg}$, contradicting (4). So far, we have proved the equivalence of (1)-(4). (1) $\Rightarrow $ (5) is trivial. (5) $\Rightarrow $ (6) is similar to (1) $\Rightarrow $ (2), by considering a sequence of nonempty open sets of $X$, rather than a sequence of elements of $X$. Full details can be found in \cite[Fact 6.1, Corollary 4.6 and condition (d) in Theorem 4.4]{tproc2}, by taking $\mathcal F = \mathcal O$ there. The equivalence of (6) with the first statement in (7) is true in every topological space, and is similar to the equivalence of $[ \nu, \nu]$-compactness with $\CAP_ \nu$ (recall that we are assuming that $\nu$ is regular). Details can be found again in \cite{tproc2}, by taking $\mathcal F = \mathcal O$ in Theorem 4.4(a) $\Leftrightarrow $ (c) there. For GO spaces, (7) $\Rightarrow $ (3) is a standard argument. Say, $(x_ \gamma ) _{ \gamma \in \nu} $ is strictly increasing. Put $O_ \gamma = \bigcup _{ \eta > \gamma } (x_ \gamma , x_ \eta) $. Then if $x$ is such that $|\{\gamma < \nu \mid O_ \gamma \cap U \not= \emptyset \}|= \nu$, for every neighborhood $U$ of $x$, then necessarily $(x_ \gamma ) _{ \gamma \in \nu} $ converges to $x$. Here the $O_ \gamma $'s are strictly decreasing with respect to inclusion. If we want the $O_ \gamma $'s to be pairwise disjoint, then, for $\gamma= \alpha + n$ with $\alpha=0$ or $\alpha$ limit, take $O_ \gamma = (x _{ \alpha + 2n}, x _{ \alpha + 2n+2} )$. Hence (1)-(7) are all equivalent, for every ultrafilter $D$ and every GO space $X$. (2) $\Rightarrow $ (8) Suppose that $D'$ is such that $K _{D'}^{Reg} \subseteq K_D^{Reg}$. If (2) holds, then, since $K _{D'}^{Reg} \subseteq K_D^{Reg}$, $X$ is $[ \nu, \nu]$-compact, for every cardinal $\nu \in K _{D'}^{Reg}$. Applying the equivalence of (1) and (2) \emph{in the case of the ultrafilter $D'$}, we get that $X$ is $D'$-compact. (8) $\Rightarrow $ (1) follows trivially by taking $D=D'$. The equivalence of (9) with, say, (6) and (5) is similar. \end{proof} Given a family $(X_j) _{j \in J} $ of GO spaces, the product $ \prod _{j \in J} X_j$ is not necessarily a GO space; however, it is a topological space, when endowed with the (Tychonoff) product topology. $ \prod _{j \in J} X_j$ can also be given the structure of a partially ordered set, by letting the relation $ \mathbf{x}\leq \mathbf{y}$ hold in $ \prod _{j \in J} X_j$ if and only if it holds componentwise. The next corollary deals with the above structures on $ \prod _{j \in J} X_j$. It shows that if each $X_j$ is a GO space, then many properties of $ \prod _{j \in J} X_j$ are determined by the corresponding properties of the $X_j$'s. To avoid trivial exceptions, we shall always assume that all factors in a product are nonempty. \begin{corollary} \labbel{corthm} Suppose that $D$ is an ultrafilter over some set $I$. For every family $(X_j) _{j \in J} $ of GO spaces, the following conditions are equivalent. \begin{enumerate} \item[(a)] For every $j \in J$, the GO space $X_j$ satisfies one (and hence all) of the conditions in Theorem \ref{thm}. \item[(b)] The topological space $ \prod _{j \in J} X_j$ satisfies all of the conditions (1)-(2),(5)-(9) in Theorem \ref{thm}. \item[(c)] The topological space $ \prod _{j \in J} X_j$ satisfies one of the conditions (1)-(2),(5)-(9) in Theorem \ref{thm}. \item[(d)] For every cardinal $\nu \in K_D ^{Reg}$, every monotone $\nu$-indexed sequence in $ \prod _{j \in J} X_j$ converges. \end{enumerate} \end{corollary} \begin{proof} (a) $\Rightarrow $ (b) If each $X_j$ satisfies any one of the conditions in Theorem \ref{thm}, then, by the very same theorem, $X_j$ is $D$-compact. By an easy and classical property of $D$-compactness \cite{ber70,GS,sak78}, every product of $D$-compact spaces is still $D$-compact, hence $ \prod _{j \in J} X_j$ is still $D$-compact. Now notice that the proof that (1) implies any one of the conditions (1)-(2),(5)-(7) in Theorem \ref{thm} holds for an arbitrary topological space, not only for GO spaces. Moreover, if each $X_j$ is $D$-compact, then, by Theorem \ref{thm} (1) $\Rightarrow $ (8), each $X_j$ is $D'$-compact, for every ultrafilter $D'$ such that $K _{D'} ^{Reg} \subseteq K_D ^{Reg}$. Again since $D'$-compactness is preserved under products, $ \prod _{j \in J} X_j$ is $D'$-compact, thus condition \ref{thm} (8) holds for $ \prod _{j \in J} X_j$. The proof that (9) holds for $ \prod _{j \in J} X_j$ is similar, since $D'$-pseudocompactness is preserved under products, too, \cite{GS}. (b) $\Rightarrow $ (c) is trivial. (c) $\Rightarrow $ (a) If $ \prod _{j \in J} X_j$ satisfies any one of the conditions (1)-(2),(5)-(9), then trivially each $X_j$ satisfies the same condition, hence (a) holds. (a) $\Rightarrow $ (d) By Theorem \ref{thm}, if (a) holds, then condition (3) in Theorem \ref{thm} holds, for every $j \in J$. Suppose that $(\mathbf{x}_ \gamma ) _{ \gamma \in \nu} $ is a monotone, say, increasing, sequence in $ \prod _{j \in J} X_j$. Since $\nu$ is a regular cardinal, for every $j \in J$, the projection $(x _{j, \gamma} ) _{ \gamma \in \nu} $ of $(\mathbf{x}_ \gamma ) _{ \gamma \in \nu} $ into $X_j$ is either eventually constant, or has a strictly increasing subsequence of order type $\nu$. In both cases, $(x _{j, \gamma} ) _{ \gamma \in \nu} $ converges in $X_j$. This is trivial in the former case; in the latter case, the strictly increasing subsequence converges, by condition (3) in Theorem \ref{thm}, and then also $(x _{j, \gamma} ) _{ \gamma \in \nu} $ converges (to the same point), since it is increasing. Now it is trivial to see that a sequence in a product converges if and only if each projection converges, hence $(\mathbf{x}_ \gamma ) _{ \gamma \in \nu}$ converges in $ \prod _{j \in J} X_j$. (d) $\Rightarrow $ (a) If (d) holds, then trivially each $X_j$ satisfies condition (3) in Theorem \ref{thm}, thus (a) holds. \end{proof} \section{The particular case of initial compactness} \labbel{secinit} For LOTS, the equivalence of conditions (2), (4) and (10) in the next corollary has first been proved, under different terminology and with further equivalences, in \cite[Theorem 3]{GFW}. For GO spaces the corollary appears in \cite{goic}. Recall that an ultrafilter $D$ over $\lambda$ is \emph{regular} if there is a family $(Z_ \alpha ) _{ \alpha \in \lambda } $ of members of $D$ such that the intersection of any infinite subfamily is empty (this is also called \emph{$( \omega, \lambda )$-regularity}). It is a standard application of the Axiom of Choice, or just the Prime Ideal Theorem, to show that, for every infinite cardinal $\lambda$, there exists a regular ultrafilter $D$ over $\lambda$. See, e.~g., Chang and Keisler \cite[Proposition 4.3.5]{CK}. For such an ultrafilter, it is easy to show that $ \nu \in K_D$, for every regular $\nu \leq \lambda $: see \cite[Properties 1.1(xii)]{mru} (in fact, one actually has that $K_D=[ \omega, \lambda ]$, by \cite[Properties 1.1(1) and the last paragraph in Remark 1.5(b)]{mru}, but we shall not need this here). It follows that $K _{D'} ^{Reg} \subseteq K_D^{Reg}$, for every regular ultrafilter over $\lambda$, and every ultrafilter $D'$ over any set of cardinality $\leq \lambda$. \begin{corollary} \labbel{initial} For every infinite cardinal $\lambda$, and every GO space $X$, the following conditions are equivalent. \begin{enumerate} \item $X$ is $D$-compact, for some regular ultrafilter over $\lambda$. \item $X$ is initially $\lambda$-compact. \item For every infinite regular cardinal $ \nu \leq \lambda $, and every strictly increasing (resp., strictly decreasing) $\nu$-indexed sequence of elements of $X$, the sequence has a supremum (resp., an infimum) to which it converges. \item $X$ has no (pseudo-)gap having type $\leq \lambda$. \item $X$ is $D$-pseudocompact, for some regular ultrafilter over $\lambda$. \item $X$ is weakly initially $\lambda$-compact. \item For every infinite (equivalently, every infinite regular) cardinal $ \nu \leq \lambda $, and every sequence $(O_ \gamma ) _{\gamma < \nu}$ of open nonempty sets of $X$, there is $x \in X$ such that $|\{\gamma < \nu \mid O_ \gamma \cap U \not= \emptyset \}|= \nu$, for every neighborhood $U$ of $x$. In the above condition we can equivalently require either that the $O_ \gamma $'s are pairwise disjoint, or that $O_ \gamma \subset O _{ \gamma '} $, for $\gamma> \gamma '$. \item $X$ is $D$-compact, for every ultrafilter $D$ over any set of cardinality $ \leq\lambda$. \item $X$ is $D$-pseudocompact, for every ultrafilter $D$ over any set of cardinality $ \leq\lambda$. \item $X$ is \emph{$\lambda$-bounded}, that is, every subset of cardinality $\leq \lambda$ has compact closure. \end{enumerate} \end{corollary} \begin{proof} We are first giving the proof as a consequence of Theorem \ref{thm}. As we mentioned in the introduction, a more direct proof (which nevertheless uses essentially the same ideas) can be found in \cite{goic}. We then sketch an alternative proof which goes along with some arguments in \cite{GFW}. Most of the equivalences in the corollary are the particular cases of Theorem \ref{thm} applied when $D$ is a regular ultrafilter over $\lambda$. This is the case for conditions (1), (3)-(5), (7)$_{\rm reg}$, (8) and (9), by the remarks before the statement of the corollary, and where by (7)$_{\rm reg}$ we denote condition (7) restricted to regular $\nu$'s. As far as (2) is concerned, it is standard and already proved in \cite{AU} that initial $\lambda$-compactness is equivalent to $[\nu,\nu]$-compactness for every regular $\nu \leq \lambda $, so we get the equivalence with condition (2) in Theorem \ref{thm}, in the case when $D$ is regular over $\lambda$. Alternatively, (1) $\Rightarrow $ (2) in the present corollary follows immediately from \cite[Theorem 3.4]{Ca}. Notice that \cite{Ca}, following standard use in the model theoretical setting, uses a notation in which the order of the cardinals is reversed, both in the definition of $[\mu,\nu]$-compactness and of $( \omega , \nu)$-regularity. In contrast with (2), it is not always necessarily the case that (for topological spaces in general) weak $[\nu,\nu]$-compactness for every regular $\nu \leq \lambda $ implies weak initial $\lambda$-compactness. This is Remark 30 in \cite{tapp2}, relying on an example by Garc{\'{\i}}a-Ferreira \cite{gf99}, which in turn builds on a construction by Kanamori \cite{kan86}. However, (5) implies (6) by \cite[Corollary 15]{tapp2}, (6) implies (7) by \cite[Theorem 1]{mpcap} and Retta \cite[Theorem 3(d)]{ret93}, and (7) implies (7)$_{\rm reg}$ trivially, hence, for GO spaces, they are all equivalent, since we have already proved that for GO spaces (5) and (7)$_{\rm reg}$ are equivalent. Finally, the equivalence of (8) and (10) holds for every Hausdorff regular space, \cite[Theorems 5.3 and 5.4]{sak78} (recall that it can be proved that every GO space is regular). Our first proof of the corollary is thus complete. \smallskip An alternative proof of (4) implies (10) goes as follows. Suppose that $Y \subseteq X$, and $|Y| \leq \lambda $. The closure $ \overline{Y} $ of $Y$ has no (pseudo-)gap of type $\leq \lambda$, since otherwise it would extend to a (pseudo-)gap of $X$ having the same type. Moreover, $ \overline{Y} $ has no subset of order type or reversed order type $>\lambda$, since then one could construct a subset of $Y$ having the same order type, but this is impossible, since $|Y| \leq \lambda $. In conclusion, $ \overline{Y} $ has no (pseudo-)gap at all, but this implies that it is compact, by a well-known result, e.~g., Nagata \cite[Theorem VIII.2]{nag85}. Then an alternative proof of the corollary is obtained by the following chains of implications (4) $\Rightarrow $ (10) $\Rightarrow $ (8) $\Rightarrow $ (1) $\Rightarrow $ (2) $\Rightarrow $ (3) $\Rightarrow $ (4) and (4) $\Rightarrow $ (10) $\Rightarrow $ (8) $\Rightarrow $ (9) $\Rightarrow $ (5) $\Rightarrow $ (6) $\Rightarrow $ (7) $\Rightarrow $ (4), which are either trivial, or proved before. \end{proof} \begin{corollary} \labbel{cor} Suppose that $X$ is a product of topological spaces and that all factors but at most one are GO spaces. Then the following hold. \begin{enumerate} \item $X$ is initially $\lambda$-compact if and only if each factor is initially $\lambda$-compact. \item $X$ is weakly initially $\lambda$-compact if and only if each factor is weakly initially $\lambda$-compact. \end{enumerate} \end{corollary} \begin{proof} (1) An implication is trivial. For the other direction, by the equivalence of (2) and (10) in Corollary \ref{initial}, all but at most one factor are $\lambda$-bounded. A product of regular $\lambda$-bounded spaces is still $\lambda$-bounded \cite[Lemma 4]{GFW}, or \cite[Theorem 5.7 and implications (1), (1$'$) in Diagram 3.6]{ste84}, and a product of a $\lambda$-bounded space with an initially $\lambda$-compact space is initially $\lambda$-compact \cite[Theorem 5.2 and implications (1), (2) in Diagram 3.6]{ste84}. Hence (1) follows by first grouping together the GO spaces, and then, in case, multiplying their product with the possibly non GO factor. Alternatively, one can use an argument parallel to the one we are going to give for (2). (2) Again, an implication is trivial. For the other direction, by the equivalence of (6) and (9) in Corollary \ref{initial}, all but at most one factor are $D$-pseudocompact, for every ultrafilter $D$ over any set of cardinality $\lambda$. Since $D$-pseudocompactness is preserved under products \cite{GS}, we have that all but at most one factor are $D$-pseudocompact, for every ultrafilter $D$ over any set of cardinality $\lambda$. Since the (possible) remaining factor is weakly initially $\lambda$-compact, by assumption, (2) follows from the next lemma (which, in some form or another, is probably folklore). \end{proof} \begin{lemma} \labbel{lmlm} If the topological space $X$ is $D$-pseudocompact, for every ultrafilter $D$ over $\lambda$, and the topological space $Y$ is weakly initially $\lambda$-compact, then $X \times Y$ is weakly initially $\lambda$-compact. \end{lemma} \begin{proof} Let $S_ \omega ( \lambda )$ denote the set of all finite subsets of $\lambda$. By \cite[Theorem 10]{tapp2}, a topological space is weakly initially $\lambda$-compact if and only if, for every sequence $\{ O_Z \mid Z \in S_ \omega ( \lambda ) \}$ of nonempty open sets, there exists an ultrafilter D over $S_ \omega ( \lambda ) $ such that both $\{ Z \in S_ \omega ( \lambda ) | \alpha \in Z\} \in D$, for every $\alpha \in \lambda $, and $\{ O_Z \mid Z \in S_ \omega ( \lambda ) \}$ has a $D$-limit point. So let $\{ O_Z \mid Z \in S_ \omega ( \lambda ) \}$ be a sequence of nonempty open sets in $X \times Y$: it is enough to find some $D$ as above. Without loss of generality, we can suppose that $O_Z= U_Z \times V_Z$, for every $Z \in S_ \omega ( \lambda )$. Since $Y$ is weakly initially $\lambda$-compact, by the quoted theorem, there exists an ultrafilter D as above such that $\{ V_Z \mid Z \in S_ \omega ( \lambda ) \}$ has a $D$-limit point $y$. Since $| S_ \omega ( \lambda )|= \lambda $, then $X$ is $D$-pseudocompact, hence also $\{ U_Z \mid Z \in S_ \omega ( \lambda ) \}$ has a $D$-limit point $x$. But then $(x,y)$ is a $D$-limit point of $\{ O_Z \mid Z \in S_ \omega ( \lambda ) \}$. \end{proof} See \cite[Corollary on p. 203]{GFW} for other results related to Corollary \ref{cor}. See \cite{GFS} for further results in the particular case of countable compactness, that is, $\lambda= \omega $. \section{The impact of set theory} \labbel{setth} In view of Theorem \ref{thm} (2) $\Leftrightarrow $ (4), one would be tempted to conjecture that, for every infinite regular cardinal $\nu$, a GO space $X$ is $[ \nu, \nu]$-compact if and only if $X$ ha no (pseudo-)gap of type $\nu$. However this is false: just consider $\mathbb{R}$ with the usual order, and with the discrete topology. The counterexample can be easily turned into a LOTS: let $X$ be the product $\mathbb{R} \times \mathbb{Z} $, with the lexicographic order. The induced topology on $X$ is the discrete one and, since $| X| = 2 ^{ \omega } $, then $X$ is not $[\nu, \nu]$-compact, for every $\nu \leq 2 ^{ \omega } $. On the other hand, $X$ has only gaps of type $ \omega$, thus, taking $\nu= \omega _1$ (or any $\nu$ with $ \omega <\nu \leq 2 ^{ \omega } $---of course, this is significant only if the Continuum Hypothesis fails), we get that having no gap of type $\nu$ does not necessarily imply $[\nu, \nu]$-compactness. Indeed, the equivalence holds only for a very special class of cardinals. It can be shown that a regular cardinal $\nu > \omega $ is \emph{weakly compact} if and only if, for every GO space (equivalently, LOTS) $X$, $[ \nu, \nu]$-compactness of $X$ is equivalent to $X$ having no (pseudo-)gap of type $\nu$. Put in another way, an equivalence like (2) $\Leftrightarrow $ (4) in Theorem \ref{thm} can hold only if considered \emph{simultaneously} for all $\nu $ in some appropriate class $H$ of regular cardinals. Theorem \ref{thm} shows that the equivalence holds in case $H$ can be realized as $ K_D ^{Reg} $, for some ultrafilter $D$. In this sense, the existence of a regular ultrafilter over $\lambda$ can be seen as a fact ``responsible behind the scene'' for the equivalence of initial $\lambda$-compactness with the nonexistence of (pseudo-)gaps of type $\leq\lambda$. As we mentioned, we have a proof that if $H= \{ \nu \} $ is a singleton, then the equivalence \ref{thm} (2) $\Leftrightarrow $ (4) (with $H$ in place of $ K_D ^{Reg} $ there) holds if and only if either $\nu= \omega $, or $\nu$ is weakly compact. The above considerations suggest the following definition. If $H$ a class of infinite cardinals, let us say that a topological space is \emph{$H$-compact} if it is $[\nu, \nu]$-compact, for every $\nu \in H$. Thus the equivalence (2) $\Leftrightarrow $ (4) in Theorem \ref{thm} asserts that if $H=K_D ^{Reg}$, for some ultrafilter $D$, then a GO space $X$ is $H$-compact if and only if $X$ has no (pseudo-)gap of type in $H$. If we strengthen the above conditions by asking forms of preservation under products, we get that the corresponding equivalence holds \emph{exactly} for those classes $H$ which can be expressed as unions of classes of the form $K_D ^{Reg}$. This is the content of the next corollary, which also shows that in many cases, say, when $H$ is finite, the equivalence holds if and only if $H=K_D ^{Reg}$, for some ultrafilter $D$. Thus, for GO spaces, preservation of $H$-compactness under products is highly dependent on set theory, as we shall discuss shortly after the proof of the corollary. Recall that an ultrafilter $D$ is \emph{$( \lambda ,\mu ) $-regular} if there is a set of $\mu $ members of $D$ such that the intersection of any $\lambda$ of them is empty. We refer again to \cite{mru} for more information about regularity of ultrafilters. Here regularity shall be used in connection with results from \cite{Ca}, asserting that, roughly, $( \lambda ,\mu ) $-regular ultrafilters are standard witnesses for $[ \lambda , \mu]$-compactness of products of topological spaces. Let $Reg$ denote the class of all infinite regular cardinals. An \emph{interval of regular cardinals} is a set of the form $Reg \cap [ \lambda, \mu]$, for certain cardinals $\lambda$ and $\mu $. Notice that we allow $\mu $ to be singular. If $H$ is a set of cardinals, and $F$ is a finite union of intervals of regular cardinals, we say that \emph{$H$ includes the cofinalities of the extremes of $F$} if $F$ can be represented as $F = Reg \cap \bigcup _{p \in P} [ \lambda _p, \mu _p] $, with $P$ finite, and in such a way that $ \cf \mu_p \in H$, for every $p \in P$ (of course, in case $H=F$ this is relevant only when some $\mu _p$ is singular). \begin{corollary} \labbel{gocorbb} Suppose that $H$ is a class of infinite regular cardinals. Then the following statements are equivalent. \begin{enumerate} \item Every product of a family of $H$-compact GO spaces is still $H$-compact. \item Same as (1), restricted to spaces which are regular cardinals with the order topology. \item For every $\nu \in H$ there is an ultrafilter $D_\nu$ such that $\nu \in K _{D_\nu} $ and $K _{D_\nu} ^{Reg} \subseteq H$. \item[(3$'$)] There is a class $\mathcal D$ of ultrafilters such that $H= \bigcup _{D \in \mathcal D} K _{D} ^{Reg} $. \item Every product of GO spaces without (pseudo-)gaps of type in $H$ is $H$-compact. \item More generally, if $Y$ is a product of GO spaces, and each factor of $Y$ satisfies at least one of conditions (2)-(4), (6)-(7) in Theorem \ref{thm}, with $K_D^{Reg}$ there replaced by $H$, then $Y$ is $H$-compact. \item[(5$'$)] If $Y$ is a product of GO spaces, and each factor of $Y$ satisfies all the conditions (2)-(4), (6)-(7) in Theorem \ref{thm}, with $K_D^{Reg}$ there replaced by $H$, then $Y$ is $H$-compact. \item For every finite $F \subseteq H$, there is an ultrafilter $D_F$ such that $F \subseteq K _{D_F} ^{Reg} \subseteq H$. \item More generally, for every set $F \subseteq H$ such that $F$ is a finite union of intervals of regular cardinals, and $H$ includes the cofinalities of the extremes of $F$, there is an ultrafilter $D_F$ such that $F \subseteq K _{D_F} ^{Reg}\subseteq H$. \end{enumerate} Suppose in addition that $H$ is finite, or, more generally, that $H$ is a finite union of intervals of regular cardinals and $H$ includes the cofinalities of its extremes. Then the preceding conditions are also equivalent to the following one. \begin{enumerate} \item[(8)] There is an ultrafilter $D$ such that $K_D ^{Reg}= H$. \end{enumerate} \end{corollary} \begin{proof} (1) $\Rightarrow $ (2) is trivial. (2) $\Rightarrow $ (3) Let $\mathcal T$ be the class of the regular cardinals which are not in $H$. Notice that every member of $\mathcal T$ is $[ \nu, \nu]$-compact, for every $\nu \in H$. By (2), every product of members of $\mathcal T$, too, is $[ \nu, \nu]$-compact, for every $\nu \in H$. For every $\nu \in H$, by \cite{sak78} or \cite[Theorem 3.4]{Ca}, there exists a $\nu$-decomposable (equivalently, $(\nu, \nu)$-regular, since $\nu$ is regular) ultrafilter $D_\nu$ such that every member of $\mathcal T$ is $D_\nu$-compact. It is easy to see that if $\kappa$ is a regular cardinal, $D$ is an ultrafilter, and the space $\kappa$ is $D$-compact, then $D$ is $\kappa$-descendingly complete (see the proof of \cite[Proposition 1]{tproc}, though the result is stated with different terminology), equivalently, $D$ is not $\kappa$-decomposable. In the case at hand, this shows that if $\kappa $ is regular and does not belong to $ H$, then $\kappa \not \in K _{D_\nu}$. This means exactly that $ K _{D_\nu} ^{Reg} \subseteq H$. Since $D_ \nu$ is $\nu$-decomposable, then $ \nu \in K _{D_\nu}$, hence $D_\nu$ witnesses (3). (3$'$) is a restatemetn of (3). (3) $\Rightarrow $ (4) We have to show that every product $Y$ as in (4) is $[ \nu , \nu]$-compact, for every $\nu \in H$. So let $ \nu \in H$, and let $D_ \nu$ be given by (3). Since $K _{D_\nu} ^{Reg}\subseteq H$, then no factor of $Y$ has a (pseudo-)gap in $K _{D_\nu} ^{Reg}$. By Theorem \ref{thm} (4) $\Rightarrow $ (1) applied to $D_ \nu$, each factor of $Y$ is $D_ \nu$-compact, hence also $Y$ is $D_ \nu$-compact. Hence $Y$ is $[ \nu, \nu]$-compact, since $\nu \in K _{D_\nu} ^{Reg}$ and, as we mentioned, \ref{thm} (1) $\Rightarrow $ (2) holds for every topological space. (Of course, we are repeating here some arguments from the proof of Corollary \ref{corthm}, and the implication can be also obtained as a consequence of \ref{corthm}). (4) $\Rightarrow $ (1) is trivial, since if a GO space is $[ \nu, \nu]$-compact, then it has no (pseudo-)gap of type $\nu$. Cf. the proof of \ref{thm} (2) $\Rightarrow $ (3). Hence (1)-(4) are all equivalent. (3) $\Rightarrow $ (5) is similar to (3) $\Rightarrow $ (4). (5) $\Rightarrow $ (5$'$) is trivial. (5$'$) $\Rightarrow $ (1), again, is trivial, since if a GO space is $[ \nu, \nu]$-compact, then it satisfies all the conditions (2)-(4), (6)-(7) in \ref{thm}, for that given $\nu$. Hence (1)-(5) are all equivalent. The equivalence of (3) and (6) is purely set-theoretical. (6) $\Rightarrow $ (3) is trivial since $H$ consists only of regular cardinals. For the converse, suppose that $F= \{ \nu_1, \dots, \nu_n \} $, and let $D _{\nu_1} $, \dots, $D _{\nu_n} $ be the corresponding ultrafilters given by (3). We are going to show that $D_F = D _{\nu_1} \times \dots \times D _{\nu_n}$ works in (6). Recall that, for $D$, $D'$ ultrafilters, say, over $I$, $I'$, their \emph{product} $D \times D' $ is the ultrafilter over $I \times I'$ defined by $ Z \in D \times D'$ if and only if $\{ i \in I \mid \{i' \in I' \mid (i, i') \in Z \} \in D' \} \in D$. It follows from the last statement in \cite[Proposition 7.1]{mru} that, for every regular cardinal $\nu$, the following holds: $\nu \in K _{D \times D'} $ if and only if either $\nu \in K _{D } $, or $\nu \in K _{D'} $ (the result in \cite[Proposition 7.1]{mru} is stated in terms of $(\nu,\nu)$-regularity, but, when $\nu$ is regular, this is equivalent to $\nu$-decomposability). In other words, (*) $K _{D \times D'} ^{Reg} = K _{D } ^{Reg} \cup K _{ D'} ^{Reg}$, for every pair of ultrafilters $D$ and $D'$. In particular, if ${D_F}$ is defined as above, then $K _{D_F} ^{Reg} = \bigcup _{\nu= \nu_1, \dots, \nu_n} K _{D_\nu} ^{Reg} $. By (3) we have that $K _{D_\nu} ^{Reg}\subseteq H$, for $\nu= \nu_1, \dots, \nu_n$, hence $K _{D_F} ^{Reg} \subseteq H$. Moreover, $ \nu \in K _{D_\nu}$, for $\nu= \nu_1, \dots, \nu_n$, hence $F \subseteq K _{D_F}^{Reg}$. We have proved (3) $\Rightarrow $ (6). The equivalence of (3) and (7), too, can be proved in a set-theoretical way, by a small variation on \cite[Proposition 7.6 (a) $\Leftrightarrow $ (b)]{mru}, considering the case $\chi = \omega $ there, since every ultrafilter is $ \omega$-complete. Notice that the proof of \cite[Proposition 7.6]{mru} uses $( \nu , \nu)$-regularity even for $\nu$ singular, hence a proof along those lines needs the assumption that $H$ includes the cofinalities of the extremes of $F$. In fact, the assumption is necessary, as we shall show in Remark \ref{bbrmk}. We shall give here a proof of (3) $\Rightarrow $ (7) with a stronger topological flavor, and which relies heavily on \cite{Ca}. As above, (7) $\Rightarrow $ (3) is trivial. To get the converse, we shall prove (1) $\Rightarrow $ (7). First, we need a lemma, an easy extension of \cite[Corollary 1.8(iii)]{Ca}. \renewcommand{\qedsymbol}{$\Box_{to\ be\ continued}$} \end{proof} \begin{lemma} \labbel{caiclem} Suppose that $\mathcal T$ is a class of topological spaces, and $\lambda$, $\mu $ are infinite cardinals. If every product of members of $\mathcal T$ is both $[ \cf \mu, \cf\mu ]$-compact, and $[\nu, \nu]$-compact, for every regular $\nu$ with $\lambda \leq \nu \leq \mu$, then every product of members of $\mathcal T$ is $[ \lambda, \mu ]$-compact. \end{lemma} \begin{proof} The particular case $\cf \mu \geq \lambda $ is \cite[Corollary 1.8(iii)]{Ca}, and the lemma can be proved in a similar way. Otherwise, in case $\cf \mu < \lambda $, we can apply \cite[Corollary 1.8(iii)]{Ca} for every regular $\mu' < \mu $, getting that every product of members of $\mathcal T$ is $[ \lambda, \mu' ]$-compact, for every regular $\mu' < \mu $, and this, together with $[ \cf \mu, \cf\mu ]$-compactness immediately implies $[ \lambda, \mu ]$-compactness. \end{proof} \begin{proof}[Proof of \ref{gocorbb} (continued)] We have promised a proof of (1) $\Rightarrow $ (7). So, let $F = Reg \cap \bigcup _{p \in P} [ \lambda _p, \mu _p] $, with $P$ finite, and $\cf \mu_p \in H$, for every $p \in P$. We have from (1) that every product of $H$-compact spaces is still $H$-compact. Applying Lemma \ref{caiclem} to the family $\mathcal T$ of all $H$-compact spaces, since $F $ is assumed to be a subset of $ H$ and since $H$ includes the cofinalities of the extremes of $F$, we get that, for every $p \in P$, every product of $H$-compact spaces is $[ \lambda_p, \mu_p ]$-compact. By \cite[Theorem 3.4]{Ca}, for every $p \in P$, there is a $( \lambda_p, \mu_p )$-regular ultrafilter $D_p$ such that every $H$-compact space is $D_p$-compact. As in the proof of (2) $\Rightarrow $ (3), for every regular cardinal $\kappa \not\in H$, the space $\kappa$ with the order topology is $H$-compact, hence $D_p$-compact, and then, using again \cite[Proposition 1]{tproc}, we get that $D_p$ is not $\kappa$-decomposable, thus $\kappa \not\in K _{D_p} ^{Reg} $. Since this holds for every regular $\kappa \not\in H$, then $ K _{D_p}^{Reg} \subseteq H $. Enumerating $P$ as $\{ p_1, \dots, p_n \}$, and putting $D_F = D _{p_1} \times \dots \times D _{p_n}$, we get $ K _{D_F}^{Reg} \subseteq H $, by using (*) above. Moreover, a $( \lambda, \mu )$-regular ultrafilter is $\nu$-decomposable, for every regular $\nu$ such that $\lambda \leq \nu \leq \mu$ (see, e.~g., \cite[Property 1.1(xii)]{mru}), thus $[ \lambda_p, \mu_p ] \cap Reg \subseteq K _{D_p}$, for every $p \in P$. Applying again (*), we get that $F \subseteq K _{D_F}$. In conclusion, $D_F$ witnesses (7). To complete the proof of Corollary \ref{gocorbb}, it remains to prove (7) $\Leftrightarrow $ (8) under the additional assumption, but this is trivial, by taking $F=H$ (of course, in case $H$ is finite, it is easier to use (6)). \end{proof} \begin{remark} \labbel{bbrmk} In general, the additional assumption before condition (8) in Corollary \ref{gocorbb} is necessary in order to prove the equivalence of (8) with the other conditions. First, we are going to show that (3) does not necessarily imply (8), when $H$ is infinite. Suppose that there are infinitely many measurable cardinals $\mu _1< \mu_2< \dots$, and let $H= \{ \mu_n \mid n \in \omega \}$. If $\mu $ is a measurable cardinal, there is a $\mu $-complete uniform ultrafilter $D_ \mu $ over $\mu $, thus $K _{D_\mu}= \mu=K _{D_\mu}^{Reg}$. This implies that condition (3) in \ref{gocorbb} holds for the above $H$. However, if we further assume GCH, there is no ultrafilter $D$ such that $K _{D}^{Reg}=H$. Indeed, the last paragraph in \cite[Section 7]{mru} shows that if $D$ is such an ultrafilter, then $D$ is $( \kappa , \kappa ) $-regular, for $\kappa= \sup _{n \in \omega } \mu_n $. But then \cite[Corollary 5.8]{mru} implies that $D$ is either $\cf \kappa $-decomposable, or $\kappa ^+$-decomposable, thus $K _{D}^{Reg} $ strictly contains $ H$, a contradiction (notice that $\cf \kappa = \omega $). There is also a similar counterexample in which $H$ is an interval, but $H$ does not include the cofinality of its upper extreme. Suppose that $\kappa$ is $\kappa ^{+n} $-compact, for every $ n \in \omega$. Here $\kappa ^{+n} $ is $\kappa ^{+\dots+} $ with $n$ occurrences of ``$+$''. Let $\kappa ^{+ \omega }= \sup _{n \in \omega } \kappa ^{+n} $ and $H=[ \kappa , \kappa ^{+ \omega })= Reg \cap [ \kappa , \kappa ^{+ \omega }]$. By $\kappa ^{+n} $-compactness, there is some $\kappa$-complete $( \kappa, \kappa ^{+n})$-regular ultrafilter $D_n$, thus $K _{D_n} = K _{D_n}^{Reg}=[ \kappa, \kappa ^{+n}] $. Hence $H$ satisfies condition (3) in \ref{gocorbb}. However, there is no ultrafilter $D$ such that $K _{D}^{Reg}=H$. Indeed, by \cite[Theorem 5.7 and Corollary 5.8]{mru}, such a $D$ would be either $ \kappa ^{+ \omega +1}$-decomposable, or $\cf \kappa ^{+ \omega }$-decomposable, that is, $ \omega$-decomposable; but both the above assertions contradict $K _{D}^{Reg}=H$. Thus (8) fails. Notice that this last counterexample also shows that the assumption that $H$ includes the cofinalities of the extremes of $F$ is necessary in (7): just take $F=H$ to make (7) fail. \end{remark} Corollary \ref{gocorbb} shows that certain properties of products of GO spaces depend heavily on the set theoretical universe in which we work. If there is no inner model with a measurable cardinal, then, for every ultrafilter $D$, $K_D$ is always an interval of cardinals with minimum $ \omega$, by results from Donder \cite{Do}, which extends and generalizes former results by M. Benda, C. C. Chang, R. B. Jensen, J. Ketonen, K. Prikry, J. H. Silver, among many others. See \cite[p. 363]{mru} for details. On the other hand, modulo some large cardinal consistency assumptions, it is possible to have an ultrafilter $D$ for which $K_D ^{Reg} = \{ \omega, \omega _{ \omega +1} \} $, Ben-David and Magidor \cite{BM}, Apter and Henle \cite{AH}. This is just an example concerning relatively small cardinals; many results are known, and nevertheless deep problems are still open about the possible values that the set $K_D$ can assume. See Problem 6.8 in \cite{mru}, and the comments below it. By Corollary \ref{gocorbb} (e.~g., (1) $\Leftrightarrow $ (3)), all these problems affect the behavior of GO spaces with respect to products, thus the use of ultrafilters in the results of the present section proves to be irreplaceable. We shall explicitly state just the example dealing with the smallest possible cardinals. \begin{corollary} \labbel{gocorbbb} (Assuming the consistency of a strongly compact cardinal) If $H = \{ \omega, \omega _{ \omega +1} \}$, then the preservation of $H$-compactness under products in the class of GO spaces is both relatively consistent and independent from the axioms of ZFC. \end{corollary} \begin{proof} Immediate from Corollary \ref{gocorbb}(1) $\Leftrightarrow $ (8), and the just mentioned results by Donder, Ben-David and Magidor, and Apter and Henle. \end{proof} \begin{remark} \labbel{bbbrmk} Notice that $H = \{ \omega, \omega _{ \omega +1} \}$ is the ``lowest'' set for which independence can occur. Indeed, \cite[Corollary 1.8(ii)]{Ca} proves that if every product of members of a family $\mathcal T$ is $[ \lambda ^+, \lambda ^+]$-compact, then every product of members of $\mathcal T$ is $[ \lambda , \lambda ]$-compact. Iterating, we get that if all products of members of $\mathcal T$ (whether $\mathcal T$ consists of GO spaces or not) are $[ \omega _n, \omega_n]$-compact, then all such products are $[ \omega _i, \omega_i]$-compact, for all $i \leq n$, hence also initially $ \omega_n$-compact. Moreover, if there is some infinite cardinal $\lambda$ such that every product of members of some family $\mathcal T$ is $[ \lambda , \lambda ]$-compact, then the smallest such $\lambda$ is either $ \omega$, or a measurable cardinal (hence it cannot be $ \omega _{ \omega +1} $). This is a slight generalization (with essentially the same proof) of \cite[Corollary 1.8(i)]{Ca}. \end{remark} \section{Relationships between $H$-compactness and omission of gaps in $H$} \labbel{equivcpn} The results in the present paper suggest that it is interesting to study the following property, depending on a class $H$ of infinite regular cardinals. \smallskip (**)$_H$ For every GO space $X$, $X$ is $[H]$-compact if and only if $X$ has no (pseudo-)gap of type in $H$. \smallskip In general, (**)$_H$ is false; for example, take $H=\{ \lambda \}$ with $ \omega <\lambda \leq \mathfrak c$; then $\mathbb R \times \mathbb Z $ with the lexicographic order furnishes a counterexample to (**)$_H$. However, (**)$_H$ holds in many interesting cases. The simplest affirmative case is given already by Gulden, Fleischman and Weston result in the case of LOTS, or Corollary \ref{initial} (2) $\Leftrightarrow $ (4) here for general GO spaces. In this sense, the corollary states that (**)$_H$ holds in case $H= Reg \cap [ \omega, \lambda ]$. Another example is given by Corollary \ref{gocorbb}, which implies that (**)$_H$ holds in case $H= K _{D}^{Reg} $, for some ultrafilter $D$. Notice that, for $\lambda$ regular, if a GO space is $[\lambda, \lambda ]$-compact, then, trivially, it has no (pseudo-)gap of type $\lambda$; hence, in order to prove an instance of (**)$_H$, it is enough to prove the non trivial implication. In order to give some results about (**)$_H$ we need some further definitions. Suppose that $\lambda$ is an infinite cardinal, $X$ is a linearly ordered set, and $Y \subseteq X$. If $Y$ has no maximum, we say that $Y$ is \emph{$\lambda$-full (in $X$) on the right} if, for every $y \in Y$, $ \bigcup _{y' \in Y} (y, y')_X $ has cardinality $\geq \lambda$. The subscript $_X$ in $(y, y')_X$ is intended to mean that the interval is evaluated in $X$, that is, $(y, y')_X= \{x \in X \mid y < x < y'\} $. When $X$ is clear from the context, we shall omit it. Symmetrically, if $Y$ has no minimum, $Y$ is \emph{$\lambda$-full (in $X$) on the left} if, for every $y \in Y$, $ \bigcup _{y' \in Y} (y', y)_X $ has cardinality $\geq \lambda$. For sake of brevity, if $\alpha$ is an ordinal, we say that a linear order $Y$ is \emph{$ \alpha \pm$ordered} in case $Y$ is order-isomorphic either to $ \alpha $, or to $\alpha \breve \ $, i.e., $ \alpha $ with the reversed order. The next lemma shows that $\lambda$-full subsets always exist, for linear orders of cardinality $\geq\lambda$. \begin{lemma} \labbel{omega} Suppose that $\lambda$ is an infinite regular cardinal. Then every linearly ordered set $X$ of cardinality $\lambda$ either has a $\lambda\pm$ordered subset, or a $\lambda$-full subset of order type $ \omega$. \end{lemma} \begin{proof} Suppose that there is $Y \subseteq X$ such that $|Y| = \lambda $ and, for every $y \in Y$, either $(-\infty, y)_Y$ or $(y, \infty)_Y$ has cardinality $<\lambda$. Then an easy induction of length $\lambda$, using the regularity of $\lambda$, shows that $Y$ (and hence $X$) has a $\lambda\pm$ordered subset. Otherwise, for every $Y \subseteq X$ of cardinality $\lambda $, there is $y \in Y$ such that both $(-\infty, y)_Y$ and $(y, \infty)_Y$ have cardinality $\lambda$. Then an induction of length $ \omega$ produces a strictly increasing sequence $(y_n) _{n \in \omega } $ such that $\{ y_n \mid n \in \omega \}$ is $\lambda$-full in $X$. \end{proof} \begin{proposition} \labbel{pair} Suppose that $\nu \leq \lambda$ are regular cardinals. Then the following conditions are equivalent. \begin{enumerate} \item Every linearly ordered set of cardinality $ \lambda$ has either a $ \lambda \pm$ordered subset, or a $\nu\pm$ordered $\lambda$-full subset. \item For every GO space $X$, if $X$ has no (pseudo-)gap of type $\nu$ or $\lambda$, then $X$ is $[ \lambda, \lambda ]$-compact. \item For every LOTS $X$, if $X$ has no gap of type $\nu$ or $\lambda$, then $X$ is $[ \lambda, \lambda ]$-compact. \end{enumerate} In Conditions (2) and (3) we can equivalently restrict ourselves to those $X$'s such that $|X| \leq \lambda ^\nu$. \end{proposition} \begin{proof} (Sketch) Recall that if $\lambda$ is regular, then $[ \lambda, \lambda ]$-compactness is equivalent to $\CAP_ \lambda $. We shall use this equivalent condition throughout the proof. (1) $\Rightarrow $ (2) Let $X$ be a GO space, $Z\subseteq X$, and $|Z|= \lambda $. Apply (1) to $Z$, say, $Z$ has subset $Y$ of order type $\nu$ and $\lambda$-full in $Z$. Since $X$ has no gap of type $\nu$, there is in $X$ a supremum $y$ of $Y$, and $y$ belongs to the closure of $Y$. Then every neighborhood of $y$ contains $\lambda$-many elements from $Z$, since $Y$ is $\lambda$-full in $Z$; in particular, $\lambda$-many elements from $X$, since $X \supseteq Z$. The other cases are treated in a similar way. (2) $\Rightarrow $ (3) is trivial. (3) $\Rightarrow $ (1) We shall sketch a proof of the contrapositive. So, suppose that (1) fails, thus there exists a linear order $L$ of cardinality $\lambda$ without $ \lambda \pm$ordered subsets and without $\nu\pm$ordered $\lambda$-full subsets. By Lemma \ref{omega}, necessarily $\nu> \omega $. Extend $L$ to another linearly ordered set $X$ in the following way. Fill each gap having both left and right type $\nu$ by adding an element in the middle of the gap. If the gap has left type $\nu$ but a different right type, ``fill'' the gap by adding a copy of $ \omega$. In the symmetric situation, add a copy of $ \omega \breve \ $. This procedure destroys all gaps of type $\nu$. We also want to ``forbid'' the existence (in $X$) of complete accumulation points of $L$. A candidate for such an accumulation point is some $ \ell \in L$ such that, say, $(\ell, \infty)_L$ is $\lambda$-full in $L$ on the left. Then replace $\ell$ with a copy of $ \omega$ or, according to the situation, with a copy of $ \mathbb Z $. In the symmetric situation, it might be necessary to replace $\ell$ with a copy of $ \omega \breve \ $, instead. Formally, $X$ can be considered as a subset of $L^+ \times \mathbb Z$ with the lexicographical order, where $L^+$ is the completion of $L$. It can be checked that the above constructions introduce no new gap of type $\nu$ or $\lambda$ ($L$ has no gap of type $\lambda$ from the beginning), and that no point of $X$ is a complete accumulation point of $L$ (here we need the assumption that L has no $\nu\pm$ordered $\lambda$-full subset otherwise the above ``fillings'' of gaps could introduce some complete accumulation point). Since $|L| = \lambda $, then (3) fails. In order to prove the last statement it is enough to check that the $X$ constructed in the last part of the proof has cardinality $ \leq \lambda ^ \nu$. \end{proof} \begin{corollary} \labbel{proph} If $H$ is a set of regular cardinals, then the following holds. \begin{enumerate} \item If $H= \{ \lambda \} $ is a singleton, then (**)$ _{H} $ holds if and only if $\lambda$ is either $ \omega$, or a weakly compact cardinal. \item If (**)$ _{H} $ holds, then $\inf H$ is either $ \omega$ or a weakly compact cardinal. \item If $\inf H= \omega $, then (**)$ _{H} $ holds. \end{enumerate} \end{corollary} \begin{proof} (1) is the particular case $\nu= \lambda $ of Proposition \ref{pair} (2) $\Leftrightarrow $ (1), since if $\lambda$ is regular, then every $ \lambda \pm$ordered subset is necessarily $\lambda$-full. Recall that, by a well-known characterization, an uncountable cardinal $\lambda$ is weakly compact if and only if every linear order of cardinality $\lambda$ has a $\lambda\pm$ordered subset. Notice that $ \omega$, too, satisfies the above property. (2) Let $\lambda= \inf H$. If $\lambda = \omega $, we are done. Otherwise, suppose by contradiction that $\lambda$ is not weakly compact, thus there is linear order $L$ without a $\lambda\pm$ordered subset. Consider the LOTS $X=L \times \mathbb Z$, with the lexicographic order. Then $X$ has no gap of type $\lambda$, since $\lambda> \omega $ ; moreover, it has no gap of type $\nu \in H \setminus \{ \lambda \} $, since $|X| = \lambda = \inf H< \nu$. By (**)$ _{H} $, $X$ is $[ \lambda, \lambda ]$-compact, a contradiction, since $X$ has cardinality $\lambda$ and the discrete topology. (3) Suppose that a GO space $X$ has no (pseudo-)gap of type in $H$. By Lemma \ref{omega}, for every $\lambda \in H$, condition (1) in Proposition \ref{pair} is satisfied, when $\nu= \omega $. By Proposition \ref{pair} (1) $\Rightarrow $ (2), $X$ is $[ \lambda, \lambda ]$-compact. This happens for every $\lambda \in H$, hence $X$ is $[H]$-compact. \end{proof} \section*{Disclaimer} \labbel{disc} Though the author has done his best efforts to compile the following list of references in the most accurate way, he acknowledges that the list might turn out to be incomplete or partially inaccurate, possibly for reasons not depending on him. It is not intended that each work in the list has given equally significant contributions to the discipline. Henceforth the author disagrees with the use of the list (even in aggregate forms in combination with similar lists) in order to determine rankings or other indicators of, e.~g., journals, individuals or institutions. In particular, the author considers that it is highly inappropriate, and strongly discourages, the use (even in partial, preliminary or auxiliary forms) of indicators extracted from the list in decisions about individuals (especially, job opportunities, career progressions etc.), attributions of funds, and selections or evaluations of research projects. \bibliographystyle {amsalpha} \def\cprime{$'$} \def\cprime{$'$} \def\cprime{$'$} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2] \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
\section{Introduction} In fluid--structure interaction applications, the underlying geometry of the computational domain may change significantly due to displacement of the structure. In order to deal with this situation in a standard setting with conforming elements, a mesh motion algorithm must be used. If the displacements are significant, the deformation of the mesh may lead to deteriorating mesh quality which may ultimately require re-meshing of the computational domain. Alternative, more flexible, techniques are therefore of significant practical interest. In this paper, we consider a combination of standard moving meshes and so-called CutFEM technology~\cite{BurmanClausHansboEtAl2014}. Essentially, the structure or elastic solid is first embedded into a boundary-fitted fluid mesh which moves along with the deformation of the solid to keep the fluid--structure interface intact. The motion of the fluid mesh surrounding the structure is obtained by solving an elasticity problem with given displacement at the fluid--structure interface. The boundary-fitted fluid mesh is then embedded into a fixed background mesh where we allow for an arbitrary overlap of the fluid meshes in order to facilitate the repositioning of the moving fluid mesh within the fixed background mesh. The fluid is then discretized on both the moving overlapping domain, using an Arbitrary-Lagrange-Eulerian (ALE) type approach\cite{DoneaGiulianiHalleux1982,DoneaHuertaPonthotEtAl2004}, and on the fixed background mesh, using a standard discretization posed in an Eulerian frame. The coupling at the fluid--fluid interface between the overlapping and underlying fluid meshes is handled using a stabilized Nitsche method developed for the Stokes problem in~\cite{MassingLarsonLoggEtAl2013}. The stabilization is constructed in such a way that the resulting scheme is inf-sup stable and the resulting stiffness matrix is well-conditioned independent of the position of the overlapping fluid mesh relative to the fixed background fluid mesh. As a result, optimal order error estimates are also established. In order to deal with the cut elements arising at the interface, we compute the polyhedra resulting from the intersections between the overlapping and background meshes. These polyhedra may then be described using a partition into tetrahedra; this partition may in turn be used to perform numerical quadrature. We refer to~\cite{Massing2012a} for a detailed discussion of the implementation aspects of cut element techniques in three spatial dimensions. We remark that Nitsche-based formulations for Stokes boundary and interface problems where the surface in question is described independently of a single, fixed background mesh were proposed in~\cite{BurmanHansbo2013,MassingLarsonLoggEtAl2013a,HansboLarsonZahedi2014a,BurmanClausMassing2014a}. A Nitsche-based composite mesh method was first introduced for elliptic problems in~\cite{HansboHansboLarson2003}. One may also consider formulations, where the structure is described via its moving boundary which is immersed into a fixed background fluid mesh. Prominent examples are Cartesian grid methods, e.g.~\cite{MurmanAftosmisBerger2003}, the classical immersed boundary method introduced by~\citet{Peskin1977,Peskin2003}, its finite element pendant proposed in \citep{BoffiGastaldi2003,ZhangGerstenbergerWangEtAl2004,ZhangGay2007}, formulations based on Lagrange multipliers, cf.~\citep{ZhangGerstenbergerWangEtAl2004,Yu2005,GerstenbergerWall2008,GerstenbergerWall2008a,PusoKokkoSettgastEtAl2014} and on Nitsche's method~\cite{HansboHermansson2003}. However, the use of an additional boundary-fitted fluid mesh as in the current work is attractive since it allows for the resolution of boundary layers and computation of accurate boundary stresses. Often, the construction of the surrounding fluid mesh can easily be generated by extending the boundary mesh in the normal direction. We plan to further investigate the properties of the fluid--structure coupling in future work. As our proposed scheme combines an ALE-based discretization on the fluid mesh surrounding the structure with an Eulerian-based discretization on the fixed background fluid mesh, it can be classified as hybrid Eulerian-ALE or Chimera approach. Such hybrid schemes are built upon the concept of overlapping meshes introduced for finite differences and finite volume schemes in the early works of~\citet{Volkov1968}, \citet{Starius1977,Starius1980}, \citet{Steger1983} and later by~\citet{ChesshireHenshaw1990} and~\citet{AftosmisBergerMelton1998}, where the primary concern was to ease the burden of mesh generation by composing individually meshed, static geometries. The idea of gluing meshes together was then explored for finite element methods by~\citet{CebralLoehner2005,LoehnerCebralCamelliEtAl2007,LoehnerCebralCamelliEtAl2008} to study the flow around independently meshed complex objects such as cars, collection of buildings or stents in aortic vessels. In these works, relatively simple interpolation schemes were used to communicate the solution between overlapping meshes. To achieve a physically more consistent coupling between the solution parts presented on different domains, Schwarz-type domain iteration schemes using Dirichlet/Neumann and Robin coupling on overlapping domains have been proposed for the Navier-Stokes equations in~\cite{HouzeauxCodina2003}. A completely different route was taken by~\citet{DayBochev2008} who reformulated elliptic interface problems as suitable first-order systems augmented with least-square stabilizations to enforce the interface conditions between the mesh domains to be tied together. Introducing special interpolation stencils close to the fluid--fluid interface, a finite volume based Chimera method for flow problems involving multiple moving rigid bodies was formulated in~\cite{Wang2000,EnglishQiuYuEtAl2013} and~\cite{HenshawSchwendeman2006}, where higher-order Godunov fluxes where used. This method was then extended by~\citet{BanksHenshawSchwendeman2012} to deal with (linearly) elastic solids in two space dimensions, and thus represents an instance of a hybrid ALE-fixed grid method. This approach has barely been explored in the context of finite-element methods for fluid--structure interaction problems: \citet{WallGerstenbergerGamnitzerEtAl2006} and later \citet{Shahmiri2011} used interpolation between fluid meshes and extended finite element techniques to couple fluid--fluid meshes, \citet{BaigesCodina2009} introduced an auxiliary ALE step to convect information on the fixed background mesh between two consecutive time-steps. In contrast to these contributions, our method is based on a variational finite element approach that leads to a monolithic and physically consistent coupling between the overlapping and underlying fluid meshes, which eliminates the need of introducing inconsistent interpolation operators. In addition, opposed to similar finite element based approaches presented e.g. in~\cite{WallGerstenbergerGamnitzerEtAl2006,Shahmiri2011}, our scheme used for the fluid problem is proven stable and optimally convergent, even for higher-order elements, independent of the location of the interface as shown in~\cite{MassingLarsonLoggEtAl2013}. Thus, the new scheme for the fluid--structure interaction problem proposed in this work exhibits the necessary robustness that is essential for developing reliable hybrid ALE-fixed mesh methods. In the current work, we consider the steady state deformation of a hyperelastic solid immersed into a viscous fluid governed by the Stokes equations. We solve for the steady state solution using a fixed point iteration where in each iteration the fluid, solid, and mesh motion problems are solved sequentially. We present two numerical examples in three dimensions, including one example with a manufactured reference solution. The outline of the remainder of this paper is as follows: in Section 2, we summarize the governing equations of the fluid--structure interaction (FSI) problem; in Section 3, we describe the overlapping mesh method; in Section 4, we present an algorithm for the solution of the stationary fluid--structure interaction model problem; in Section 5, we present three-dimensional numerical examples; before drawing some conclusions in Section 6. \section{A stationary fluid--structure interaction problem} We consider a fluid--structure interaction problem posed on a domain $\Omega = \OF \cup \OS$ where $\OF$ is the domain occupied by the fluid and $\OS$ is the domain occupied by the solid. We assume that both $\OF$ and $\OS$ are open and bounded and that they are such that $\OF \cap \OS = \emptyset$. Furthermore, we decompose the fluid domain into two disjoint subdomains $\OF_1$, $\OF_2$ such that $\OF = \OF_1 \cup \OF_2$. Here, $\OF_2$ represents a part of the fluid domain surrounding the solid domain $\OS$; more precisely, we assume that $\partial \OF_1 \cap \partial \OS = \emptyset$. The fluid--structure interface is denoted by $\IFS = \partial \OF_2 \cap \partial \OS$ and the interface between the two fluid domains is denoted by $\IFF = \partial \OF_1 \cap \partial \OF_2$. Here, the topological boundary $\partial X$ for any given set $X$ is defined by $\partial X = \overline{X} \setminus \overset{\circ}{X}$ where $\overline{X}$ and $\overset{\circ}{X}$ denotes the closure and interior of $X$, respectively. For simplicity, we assume that the fluid domain boundary consists of two disjoint parts: $\partial \OF = \IFS \cup \partial \OF_D$, and that the solid domain boundary decomposes in a similar manner: $\partial \OS = \IFS \cup \partial \OS_D$. This notation is summarized in Figure~\ref{fig:ref-phy-domains}. \begin{figure} \begin{center} \includegraphics[width=0.45\textwidth]{pdf/flap-channel-physical-domains-dd-version-2d.pdf} \caption{Fluid and structure domains for the stationary fluid--structure interaction problem.} \label{fig:ref-phy-domains} \end{center} \end{figure} We assume that the fluid dynamics are governed by the Stokes equations of the following form: find the fluid velocity $\uF : \OF \to \mathbb{R}^3$ and the fluid pressure $\pF: \OF \to \mathbb{R}$ such that \begin{alignat}{3} - \nabla \cdot (\nu^f \nabla \uF - \pF \mathbf{I}) &= \boldsymbol{f}^f & & \quad \text{in } \OF, \label{eq:fluid-momentum-pde} \\ \nabla \cdot \uF &= 0 & &\quad \text{in } \OF, \label{eq:fluid-compressibility-pde} \end{alignat} where $\boldsymbol{f}^f$ is a given body force and $\nu^f$ is the fluid viscosity. Next, we assume that the velocity is prescribed on both the fluid--structure interface and on the remainder of the fluid boundary: \begin{alignat}{3} \uF &= 0 && \quad \text{on } \IFS , \label{eq:fluid-bc1} \\ \uF &= \boldsymbol{g}^f && \quad \text{on } \partial \OF_D . \label{eq:fluid-bc2} \end{alignat} Moreover, we enforce the continuity of the fluid velocity and of the fluid ``stress'' on the fluid--fluid interface by the following conditions: \begin{alignat}{3} \jump{\uF} &= 0 & & \quad \text{on } \IFF, \label{eq:fluid-interface-jump-1} \\ \jump{ \left (\nu^f \nabla \uF - \pF \mathbf{I} \right ) \cdot \boldsymbol{n}} &= 0 & &\quad \text{on } \IFF . \label{eq:fluid-interface-jump-2} \end{alignat} Here $\jump{v} = v_1 - v_2$ denotes the jump in a function (or each component of a vector field) $v$ over the interface $\IFF$ where $v_i = v|_{\OF_i}$ denotes the restriction of $v$ to $\OF_i$ for $i=1, 2$. Furthermore, $\boldsymbol{n}$ is the unit normal of $\IFF$ directed from $\OF_2$ into $\OF_1$ . Correspondingly, we assume that the structure deforms as an elastic solid satisfying the equations: find $\uS : \OS \rightarrow \mathbb{R}^3$ such that \begin{equation} - \nabla \cdot \sigmaS(\uS) = \boldsymbol{f}^s \quad \text{in } \OS, \label{eq:structure-pde} \end{equation} where $\sigmaS$ is the (Cauchy) stress tensor and $\boldsymbol{f}^s$ is a given body force. The precise form of the Cauchy stress tensor will depend on the choice of the elastic constitutive relation. In later sections, we will consider both linearly elastic and hyperelastic constitutive equations relating the displacement to the stress. As boundary conditions, we assume that the displacement of the structure is given on part of the boundary and that the structure experiences a boundary traction $\boldsymbol{t}^s_N$ on the fluid--structure interface: \begin{alignat}{3} \uS &= \boldsymbol{g}^s_D && \quad \text{on } \partial \OS_D, \label{eq:structure-bc1} \\ \sigmaS(\uS) \cdot \boldsymbol{n} &= \boldsymbol{t}^s_N && \quad \text{on } \IFS . \label{eq:structure-bc2} \end{alignat} The coupling between the fluid and the structure problems requires the fluid and solid stresses and velocities to be in equilibrium at the interface $\IFS$. In the stationary case considered here, these kinematic and kinetic continuity conditions are taken care of by ensuring that~\eqref{eq:fluid-bc1} and \begin{equation} \label{eq:kinematic-matching-conditions} \boldsymbol{t}^s_N = \sigmaF(\uF) \cdot \boldsymbol{n} \end{equation} hold, where $\sigmaF$ is the fluid stress tensor: $\sigmaF(\uF, \pF) = 2 \nu^f \epsilon(\uF) - \pF \mathbf{I}$ and $\epsilon(\uF)$ is the symmetric gradient $\epsilon(\uF) = \tfrac{1}{2}(\nabla \uF + \nabla ({\uF})^{\top})$. In summary, the stationary fluid--structure interaction problem considered in this work is completely described by the set of equations~\eqref{eq:fluid-momentum-pde}--\eqref{eq:kinematic-matching-conditions}. \section{An overlapping finite element discretization of the FSI problem} \label{sec:chimera-method} The nonlinear nature of the fluid--structure interaction problem~\eqref{eq:fluid-momentum-pde}--\eqref{eq:kinematic-matching-conditions} mandates a nonlinear solution scheme such as a Newton-type or fixed-point method. A classical and well-studied approach is to decompose the coupled problem into separate systems of equations via a Dirichlet--Neumann fixed-point iteration~\cite{Nobile2001,LeTallec2001,Kuttler2009}. This is also the route taken here. Alternatively, more sophisticated iteration schemes based on a Robin-type reformulation of the interface conditions~\eqref{eq:fluid-bc1},\eqref{eq:structure-bc2}, and \eqref{eq:kinematic-matching-conditions} might be employed, see for instance \cite{Badia2008,Badia2009,BadiaNobileVergara2008}. The basic idea of the Dirichlet--Neumann fixed-point iteration is to start with solving the fluid problem~\eqref{eq:fluid-momentum-pde}--\eqref{eq:fluid-interface-jump-2} on a given starting domain. The resulting fluid boundary traction acting on the fluid--structure interface then serves as Neumann data for the structure problem~\eqref{eq:structure-pde}--\eqref{eq:kinematic-matching-conditions}. The structure deformation dictates a displacement of the fluid domain boundary, and in turn, a new configuration of the fluid domain. This sequence of steps is repeated until convergence. Each of the three subproblems (the fluid problem, the structure problem and the domain deformation) will be solved numerically using separate finite element discretizations. Overall, we will employ an overlapping mesh method in which a fixed background mesh is used for part of the fluid domain and a moving mesh is used for the combination of the structure domain and its surrounding fluid domain. We note that methods based on overlapping meshes (as the one considered here) are sometimes also called Chimera methods. Before describing each of the discretizations, we here present an overview of the set-up of the computational domains. For simplicity, we assume that the computational domain $\Omega$ is fixed throughout the fixed-point iteration while the fluid and structure subdomains will be updated in each iteration step. In each step, we consider the following set-up, illustrated in Figure~\ref{fig:chimera-mesh-configuration}, of the computational domains. First, we assume that $\Omega$ is tessellated by a background mesh $\mesh_0$. Second, we assume that the current representation of the subdomains $\OF_2$ and $\OS$ are tessellated by meshes $\mesh^f_2$ and $\mesh^s$, respectively, and that these meshes match at their common interface. As a result, $\mesh^{fs} = \mesh^f_2 \cup \mesh^s$ defines an admissible and conforming mesh of the combined domain $\Omega^{fs} = (\overline{\OF_2} \cup \overline{\OS})^{\circ}$. All meshes are assumed to be admissible and to consist of shape-regular simplices. \begin{figure}[htb] \begin{center} \includegraphics[width=0.45\textwidth]{pdf/flap-channel-meshes-dd-version-2d.pdf} \hspace{1ex} \includegraphics[width=0.45\textwidth]{pdf/flap-channel-reduced-background-mesh-2d.pdf} \caption{Chimera-mesh configuration of the computational domain in the starting step of the fixed-point iteration. Left: Fixed fluid background mesh $\mathcal{T}_0$ overlapped by the structure mesh $\widehat{\mathcal{T}}^s$ and a surrounding fitting fluid mesh $\widehat{\mathcal{T}}^f_2$. Right: Reduced fluid background mesh $\mathcal{T}^{\ast}_1$ and fluid overlap region $\Omega_O$.} \label{fig:chimera-mesh-configuration} \end{center} \end{figure} We further note that the background tessellation $\mesh_0$ may be decomposed into three disjoint subsets: \begin{equation} \label{p3:eq:mesh-splitting} \mesh_0 = \mathcal{T}_{0,1} \cup \mathcal{T}_{0,2} \cup \mathcal{T}_{0, \Gamma}. \end{equation} Here $\mathcal{T}_{0,1}, \mathcal{T}_{0,2}, \mathcal{T}_{0,\Gamma}$ are defined with reference to $\Omega^{fs}$ and denote the sets of elements in $\mesh_0$ that are \emph{not}, \emph{completely} or \emph{partially} overlapped by $\Omega^{fs}$. More precisely, $\mathcal{T}_{0,1} = \{ T \in \mathcal{T}_0 : T \subset \overline{\OF_1} \}$, $\mathcal{T}_{0,2} = \{ T \in \mathcal{T}_0 : T \subset \overline{\Omega^{fs}} \}$ and $\mathcal{T}_{0,\Gamma} = \{ T \in \mathcal{T}_0 : | T \cap \OF_1 | > 0 \text{ and } |T \cap \Omega^{fs}| > 0 \}$. In addition, we assume that $\mesh_0$ is sufficiently fine near the fluid--fluid interface in the sense that $T \cap \OS = \emptyset$ for all $T \in \mathcal{T}_{0, \Gamma}$. In other words, the elements in the fluid background mesh have to be small enough close to $\Gamma_{ff}$ such that a single element does not stretch from the fluid--fluid interface to the fluid--structure interface. Next, we introduce the \emph{reduced} background mesh $\mesh_1^{\ast}$, consisting of the elements in $\mesh_0$ that are either not or only partially overlapped by $\Omega^{fs}$, and associated domain $\Omega_1^{\ast}$: \begin{align} \mesh_1^{\ast} &= \mathcal{T}_{0,1} \cup \mathcal{T}_{0, \Gamma}, \quad \Omega_1^{\ast} = \bigcup_{T \in \mesh_1^{\ast}} T. \label{p3:eq:meshast-1} \end{align} Note that $\Omega_1^{\ast}$ contains (but is generally larger than) $\OF_1$. We further define the so-called \emph{fluid overlap region} $\Omega_O = \OF_2 \cap \Omega_1^{\ast}$. In general, for each overlapping mesh configuration described by some (background) mesh and some overlapping domain, the procedure described above defines what we shall refer to as the reduced (background) mesh. \subsection{An overlapping mesh method for the fluid problem} \label{ssec:fluid-discretization} In this section, we present a finite element discretization of~\eqref{eq:fluid-momentum-pde}--\eqref{eq:fluid-interface-jump-2} posed on a pair of overlapping meshes, first proposed in~\citet{MassingLarsonLoggEtAl2013}. The pair of meshes consist of an overlapped mesh and an overlapping mesh: in our case the reduced background mesh $\mesh_1^{\ast}$ plays the role of the overlapped mesh, while $\mesh^f_2$ is the overlapping mesh. For any given mesh $\mathcal{T}$, let $V_h(\mathcal{T})$ be the space of continuous piecewise linear vector fields and let $Q_h(\mathcal{T})$ be the space of continuous piecewise linears, both defined relative to $\mathcal{T}$. We define the composite finite element spaces $V_h$ and $Q_h$ for the overlapping fluid meshes by \begin{equation} \label{p3:eq:elementspaces} V^f_h = V_h(\mesh_1^{\ast}) \bigoplus V_h(\mesh^f_2), \quad Q^f_h = Q_h(\mesh_1^{\ast}) \bigoplus Q_h(\mesh^f_2). \end{equation} Moreover, we denote by $V^f_{h, \boldsymbol{g}^f}$ the subspace of $V^f_h$ that satisfies the boundary conditions~\eqref{eq:fluid-bc1}--\eqref{eq:fluid-bc2} and by $V^f_{h, \boldsymbol{0}}$ the corresponding homogeneous version. The overlapping mesh discretization of~\eqref{eq:fluid-momentum-pde}--\eqref{eq:fluid-interface-jump-2} then reads: find $(\uF_h, \pF_h) \in V^f_{h, \boldsymbol{g}^f} \times Q^f_h$ such that \begin{equation} \label{eq:stokes-olm} A^f_h(\uF_h, \pF_h ; \boldsymbol{v}, q) = L^f_h(\boldsymbol{v}, q) \quad \forall\, (\boldsymbol{v}, q) \in V^f_{h, \boldsymbol{0}} \times Q^f_h, \end{equation} where $A^f_h$ is defined for all $\boldsymbol{u}, \boldsymbol{v} \in V^f_h$ and all $p, q \in Q^f_h$ by \begin{equation} \begin{split} \label{eq:Ah-olm} A^f_h(\boldsymbol{u}, p; \boldsymbol{v}, q) = a^f_h(\boldsymbol{u}, \boldsymbol{v}) + b^f_h(\boldsymbol{v}, p) + b^f_h(\boldsymbol{u}, q) + i^f_h(\boldsymbol{u}, \boldsymbol{v}) - j^f_h(p, q), \end{split} \end{equation} and the forms $a^f_h$, $b^f_h$, $i^f_h$ and $j^f_h$ are given by \begin{align} \label{eq:ah-olm} a^f_h(\boldsymbol{u}, \boldsymbol{v}) &= (\nabla \boldsymbol{u} , \nabla{\boldsymbol{v}})_{\OF_1 \cup \OF_2} - (\meanvalue{\nablan \boldsymbol{u}}, \jump{{\boldsymbol{v}}} )_{\IFF} - (\meanvalue{\nablan \boldsymbol{v}}, \jump{{\boldsymbol{u}}} )_{\IFF} + \gamma (h^{-1} \jump{{\boldsymbol{u}}}, \jump{{\boldsymbol{v}}} )_{\IFF}, \\ \label{eq:bh-olm} b^f_h(\boldsymbol{v}, q) &= - (\nabla \cdot \boldsymbol{v}, q)_{\OF_1 \cup \OF_2} + (\jump{\boldsymbol{v}} \cdot \boldsymbol{n}, \meanvalue{q})_{\IFF}, \\ \label{eq:ih-olm} i^f_h(\boldsymbol{u}, \boldsymbol{v}) &= (\nabla(\boldsymbol{u}_1 - \boldsymbol{u}_2), \nabla(\boldsymbol{v}_1 - \boldsymbol{v}_2))_{\Omega_O}, \\ \label{eq:jh-olm} j^f_h(p, q) &= \delta \sum_{T \in \mesh_1^{\ast} \cup \mesh^f_2 } h_T^2 (\nabla p, \nabla q)_T, \end{align} for $\delta > 0$. Here and throughout, $(\cdot, \cdot)_{K}$ denotes the $L^2(K)$ inner product over some domain $K$, while $\meanvalue{v}$ denotes a convex combination $\meanvalue{v} = \alpha_1 v_1 + \alpha v_2$ with $\alpha_1 + \alpha_2 = 1$ of $v$ across the interface $\IFF$. In particular, we choose $\meanvalue{v} = v_2$ in accordance with~\citet{HansboHansboLarson2003}. Finally, the linear form $L^f_h$ is defined by \begin{equation} L^f_h(\boldsymbol{v}, q) = (\boldsymbol{f}^f, \boldsymbol{v}) - \delta \sum_{T\in \mesh_1^{\ast} \cup \mesh^f_2} h_T^2 (\boldsymbol{f}^f, \nabla q)_T \label{eq:Lh-olm} \end{equation} for all $\boldsymbol{v} \in V^f_h$ and all $q \in Q^f_h$. A major strength of the employed scheme for the fluid problem is that the extension of the stabilization term~\eqref{eq:jh-olm} from the physical domain $\Omega^f_1$ to the overlap region $\Omega_O$ in combination with the least-square stabilization~\eqref{eq:ih-olm} results in a well-conditioned and optimally convergent scheme, independent of the location of the overlapping mesh with respect to the fixed background mesh. Thereby, typical difficulties arising from potentially small cut cells where $|T \cap \Omega^f_2 | \ll | T |$ for $T \in \mathcal{T}_{0,\Gamma}$ are completely eliminated. Consequently, for a continuous solution $(\uF, \pF)$ satisfying of~\eqref{eq:fluid-momentum-pde}--\eqref{eq:fluid-interface-jump-2} and a discrete solution $(\uF_h, \pF_h)$ satisfying~\eqref{eq:stokes-olm}, the following optimal error estimate holds independently of the fluid--fluid interface position~\citep{MassingLarsonLoggEtAl2013}: \begin{equation} \tn ( \uF - \uF_h , \pF - \pF_h ) \tn \leqslant C h \left( | \uF |_{2, \OF} + |\pF|_{1, \OF} \right). \label{eq:a-priori} \end{equation} Here, $\tn \cdot \tn$ is an appropriate version of the standard norm on $H^1(\OF) \times L^2(\OF)$ accounting for the fluid overlap region $\Omega_O$; see~\citep{MassingLarsonLoggEtAl2013} for more details. \subsection{A finite element discretization of the structure problem} The structure problem is described by~\eqref{eq:structure-pde}--\eqref{eq:structure-bc2} in the current solid domain. As the current solid domain is actually unknown, a standard approach to discretizing such problems is to map the governing equations back to a fixed reference (Lagrangian) frame. We choose a reference domain $\refe{\Omega}^s$ with coordinates $\refe{\boldsymbol{x}}$ and denote the deformation map from the reference to the current solid domain by $\boldsymbol{\phi}^s$: \begin{equation} \boldsymbol{x} = \boldsymbol{\phi}^s(\refe{\boldsymbol{x}}) \quad \text{for } \refe{\boldsymbol{x}} \in \refe{\Omega}^s . \end{equation} In general, the notation for all domains and quantities pulled back to the Lagrangian framework will be endowed with a $\refe{\phantom{u}}$; for instance $\refe{\Omega}^s$ and $\refe{\boldsymbol{u}}^s$ denote the solid reference domain and solid displacement in the reference frame, respectively. In particular, $\boldsymbol \phi^s = \mathbf{I} + \refe{\boldsymbol{u}}^s$. In the Lagrangian frame, the problem reads: find the solid displacement $\refe{\boldsymbol{u}}^s: \refe{\Omega}^s \to \mathbb{R}^3$ such that \begin{alignat}{3} \label{eq:structure-pde-ref} - \nabla \cdot \refe{\Pi} (\refe{\boldsymbol{u}}^s) &= \refe{\boldsymbol{f}}^s && \quad \text{in } \refe{\Omega}^s, \\ \label{eq:structure-bc1-ref} \refe{\boldsymbol{u}}^s &= \refe{\boldsymbol{g}}^s_D && \quad \text{on } \partial \refe{\Omega}^s_D, \\ \label{eq:structure-bc2-ref} \refe{\Pi} (\refe{\boldsymbol{u}}^s) \cdot \refe{\boldsymbol{n}} &= \refe{\boldsymbol{t}}^s_N && \quad \text{on } \refe{\Gamma}^{fs} . \end{alignat} Here, the displacement $\refe{\boldsymbol{u}}^s$ and the boundary displacement $\refe{\boldsymbol{g}}^s_D$ result from the standard affine pull-back of the corresponding quantities in the current domain, for instance $\refe{\boldsymbol{u}}^s(\refe{\boldsymbol{x}}) = \boldsymbol{u}^s(\boldsymbol{x})$, and $\refe{\boldsymbol{n}}$ is the outward normal of the fluid--structure interface in the reference frame. Further, let $\boldsymbol{F}^s = \nabla \boldsymbol{\phi}^s$ and $J^s = \det \boldsymbol{F}^s$. We let $\refe{\boldsymbol{f}}^s(\refe{\boldsymbol{x}}) = J^s \boldsymbol{f}^s(\boldsymbol{x})$. Moreover, $\refe{\Pi}(\refe{\boldsymbol{u}}^s)$ denotes the first Piola-Kirchhoff stress tensor, resulting from a Piola transformation of the Cauchy stress tensor $\sigmaS$: \begin{equation} \refe{\Pi}(\refe{\boldsymbol{u}}^s) (\refe{\boldsymbol{x}}) = J^s(\refe{\boldsymbol{x}}) \, \sigmaS (\boldsymbol{\phi}^s( \refe{\boldsymbol{x}})) (\boldsymbol{F}^s)^{-\top} (\refe{\boldsymbol{x}}) . \end{equation} In view of~\eqref{eq:kinematic-matching-conditions}, we will enforce that the boundary traction acting on the solid in the reference domain is the Piola transform of the fluid traction exerted on the fluid--structure interface by the fluid in the current or physical configuration. This will be detailed in Section~\ref{sec:algorithm}. The governing equations~\eqref{eq:structure-pde-ref}--\eqref{eq:structure-bc2-ref} must be completed by a constitutive equation relating the stress to the strain. In the case of a hyperelastic material, by definition, there exists a strain energy density $\Psi$ such that \begin{equation} \refe{\Pi}(\boldsymbol{F}) = \frac{\partial \Psi}{\partial \boldsymbol{F}} . \end{equation} One example is the St.~Venant--Kirchhoff material model, in which \begin{equation} \Psi(\boldsymbol{F}) = \mu \tr \boldsymbol{E}^2 + \frac{\lambda}{2} (\tr \boldsymbol{E})^2, \quad \text{where } \boldsymbol{E} = \frac{1}{2} (\boldsymbol{F}^{\top} \boldsymbol{F} - \mathbf{I}), \label{eq:venant-kirchhoff-material} \end{equation} for Lam\'e constants $\mu, \lambda > 0$. In the special case of a linearly elastic material, we assume that the reference and physical configurations coincide, so that~\eqref{eq:structure-pde}--\eqref{eq:structure-bc2} hold over $\refe{\Omega}^s$ directly with $\sigmaS(\uS) = 2 \mu \epsilon(\uS) + \lambda \tr (\epsilon(\uS)) \mathbf{I}$. To solve~\eqref{eq:structure-pde-ref}--\eqref{eq:structure-bc2-ref} numerically, let $\refe{\mathcal{T}}^s$ be a tessellation of $\refe{\Omega}^s$ such that $\mesh^s = \boldsymbol{\phi}^s(\refe{\mathcal{T}}^s)$ and introduce the finite element approximation space \begin{equation} \refe{V}^s_{h, \boldsymbol{g}} = \{ \boldsymbol{v} \in V_h(\refe{\mathcal{T}}^s) \quad \text{such that} \quad \boldsymbol{v}|_{\partial \refe{\Omega}^s_D} = \boldsymbol{g} \}, \end{equation} where $V_h(\refe{\mathcal{T}}^s)$ is the space of continuous piecewise linear vector fields defined relative to $\refe{\mathcal{T}}^s$ as before. The finite element formulation of~\eqref{eq:structure-pde-ref}--\eqref{eq:structure-bc2-ref} then reads: find $\refe{\boldsymbol{u}}^s_h \in \refe{V}^s_{h, \refe{\boldsymbol{g}}^s_D}$ such that \begin{align} (\refe{\Pi}(\refe{\boldsymbol{u}}^s_h), \nabla \boldsymbol{v})_{\refe{\Omega}^s} - (\refe{\boldsymbol{t}}^s_N, \boldsymbol{v})_{\refe{\Gamma}^{fs}} - (\refe{\boldsymbol{f}}^s, \boldsymbol{v})_{\refe{\Omega}^s} &= 0 \quad \forall\, \boldsymbol{v} \in \refe{V}^s_{h, \boldsymbol{0}}. \label{eq:structure-problem-weak-form} \end{align} Note that the generally nonlinear constitutive relation and the geometric nonlinearity mandate a nonlinear solution scheme, such as a Newton method or an inner fixed-point iteration for~\eqref{eq:structure-problem-weak-form}. \subsection{Deformation of the surrounding fluid domain} \label{ssec:mesh-motion} The overlapping mesh method relies on keeping the background part of the fluid domain $\OF_1$ fixed while moving the part of the fluid domain $\OF_2$ surrounding the structure. This movement ensures that the mesh $\mesh^f_2$ of the latter part of the fluid domain and the structure mesh $\mesh^s$ match at the fluid--structure interface. The movement is dictated by the structure deformation only at the fluid--structure interface: the motion of the interior of the fluid domain $\OF_2$ is subject to numerical modeling. Standard approaches for the domain motion include mesh smoothing via diffusion-type equations or treating the fluid domain as a pseudo-elastic structure. Here, we choose the latter approach and model the deformation of the fluid domain as a linearly elastic structure. This approach allows for typically larger deformations than a simple diffusion equation based mesh smoothing, while avoiding unnecessary complexity. We start with a fixed reference domain $\refe{\Omega}^f_2$ and consider the following mesh deformation problem over this domain: find the mesh displacement $\refe{\boldsymbol{u}}^m : \refe{\Omega}^f_2 \rightarrow \mathbb{R}^3$ such that \begin{alignat}{3} - \nabla \cdot \refe{\boldsymbol \sigma}^m (\refe{\boldsymbol{u}}^m) &= 0 \quad && \text{in } \refe{\Omega}^f_2, \label{eq:mesh-motion-strong} \\ \refe{\boldsymbol \sigma}^m (\refe{\boldsymbol{u}}^m) \cdot \refe{\boldsymbol{n}} &= 0 \quad &&\text{on } \refe{\Gamma}^{ff}, \label{eq:mesh-motion-nbc} \\ \refe{\boldsymbol{u}}^m &= \refe{\boldsymbol{u}}^s \quad &&\text{on } \refe{\Gamma}^{fs}, \label{eq:mesh-motion-dbc} \end{alignat} where the stress tensor $\refe{\boldsymbol \sigma}^m$ is given by \begin{equation} \refe{\boldsymbol \sigma}^m(\refe{\boldsymbol{u}}^m) = 2 \mu_m \epsilon(\refe{\boldsymbol{u}}^m) + \lambda_m \tr(\epsilon (\refe{\boldsymbol{u}}^m)) \mathbf{I} \label{eq:hookes-law} \end{equation} for chosen Lam\'e constants $\mu_m, \lambda_m > 0$. Let now $\refe{\mathcal{T}}^f_2$ be a tessellation of $\refe{\Omega}^f_2$. We define the finite element space $\refe{V}^m_{h, \boldsymbol{g}}$ by \begin{equation} \refe{V}^m_{h, \boldsymbol{g}} = \{ \boldsymbol{v} \in V_h(\refe{\mathcal{T}}^f_2) \quad \text{such that} \quad \boldsymbol{v}|_{\refe{\Gamma}^{fs}} = \boldsymbol{g} \}. \end{equation} The corresponding finite element formulation of the mesh problem~\eqref{eq:mesh-motion-strong}--\eqref{eq:mesh-motion-dbc} is then: find $\refe{\boldsymbol{u}}^m_h \in \refe{V}^m_{h, \refe{\boldsymbol{u}}^s}$ such that \begin{equation} (\refe{\boldsymbol \sigma}^m (\refe{\boldsymbol{u}}^m_h), \boldsymbol{v})_{\refe{\Omega}^f_2} = 0 \quad \forall\, \boldsymbol{v} \in \refe{V}^m_{h, \boldsymbol{0}}. \end{equation} Finally, we define $\mathcal{T}^{f}_2 = \boldsymbol{\phi}^m_h(\refe{\mathcal{T}}^f_2)$ with the discrete mesh deformation $\boldsymbol{\phi}^m_h = \mathbf{I} + \refe{\boldsymbol{u}}^m_h$. The current surrounding fluid domain is then defined accordingly: $\Omega^{f}_2 = \boldsymbol{\phi}^m_h(\refe{\Omega}^f_2)$. The use of boundary condition~\eqref{eq:mesh-motion-nbc} ensures that the fluid--structure interface is preserved in the sense that \begin{align} \Gamma^{fs} = \partial \Omega^{f}_2 \cap \partial \Omega^{s} = \boldsymbol{\phi}^m_h(\refe{\Gamma}^{fs}) = \boldsymbol{\phi}^s_h(\refe{\Gamma}^{fs}). \end{align} where $\boldsymbol{\phi}_h^s$ is the solid deformation given by the discrete solution $\refe{\boldsymbol{u}}^{s}_h$ of problem~\eqref{eq:structure-problem-weak-form}. \section{Solution algorithm for the discretized FSI problem} \label{sec:algorithm} We are now in a position to give a detailed description of the overall solution scheme for the fully coupled fluid--structure interaction problem. We start with reviewing the formulation of the fluid--structure coupling in the discrete setting. For the discrete formulation, a third interface condition~\eqref{eq:mesh-motion-dbc} needs to be added to the two interface conditions~\eqref{eq:fluid-bc1} and ~\eqref{eq:structure-bc2}, due to the additional mesh deformation problem described in Section~\ref{ssec:mesh-motion}. The mesh deformation allows to express the fluid stress tensor acting on $\IFS$ in the reference configuration $\refe{\Gamma}^{fs}$ via a Piola transformation. Consequently, the stress equilibrium condition~\eqref{eq:structure-bc2} at the fluid--structure interface can be reformulated in the Lagrangian frame according to~\eqref{eq:structure-bc2-ref}. In summary, the discrete formulation of the fluid--structure interface conditions reads: \begin{align} \label{eq:final-fsi-condition-1} \uF &= 0 \quad \;\;\;\, \text{on } \IFS, \\ \label{eq:final-fsi-condition-2} \refe{\boldsymbol{u}}^s &= \refe{\boldsymbol{u}}^m \quad \text{on } \refe{\Gamma}^{fs}, \\ \label{eq:final-fsi-condition-3} \refe{\Pi}(\refe{\boldsymbol{u}}^s) (\refe{\boldsymbol{x}}) \cdot \refe{\boldsymbol{n}}(\refe{\boldsymbol{x}}) &= J^m(\refe{\boldsymbol{x}}) \, \sigmaF (\boldsymbol{\phi}^m( \refe{\boldsymbol{x}})) (\boldsymbol{F}^m)^{-\top} (\refe{\boldsymbol{x}}) \cdot \refe{\boldsymbol{n}}(\refe{\boldsymbol{x}}) \quad \text{on } \refe{\Gamma}^{fs}. \end{align} As outlined in Section~\ref{sec:chimera-method}, we employ a classical Dirichlet--Neumann fixed-point iteration approach to ensure that the interface conditions~\eqref{eq:final-fsi-condition-1}--\eqref{eq:final-fsi-condition-3} are approximately satisfied by the computed solution within a user provided tolerance. The iteration scheme is presented in detail in Algorithm~\ref{alg:fixed-point-iteration}, where the relaxation parameter $\omega^{i}$ was chosen dynamically to accelerate the convergence of the fixed-point iteration. Moreover, the fluid boundary traction is incorporated as Neumann data in the weak formulation of the structure problem by a properly chosen functional representing the boundary traction weighted with some given test function. A thorough explanation of both of these intermediate steps will be given in the next sections. \begin{algorithm}[htb] \begin{tabbing} \\ $\refe{\boldsymbol{u}}^{s, k} := 0$ \\ $\refe{\boldsymbol{u}}^{m, k} := 0$ \\ \textbf{do } \\ \hspace*{2em} \textbf{Update overlapping fluid meshes} \\ \hspace*{2em} $\Omega^{s,k+1} := (\mathbf{I} + \refe{\boldsymbol{u}}^{s, k})(\refe{\Omega}^s)$ \\ \hspace*{2em} $\Omega^{f,k+1}_2 := (\mathbf{I} + \refe{\boldsymbol{u}}^{m, k})(\refe{\Omega}^f_2)$ \\ \hspace*{2em} $\Omega^{fs,k+1} := \Omega^{s,k+1} \cup \Omega^{f,k+1}_2$ \\ \hspace*{2em} Compute reduced background mesh $(\mathcal{T}^{f,k+1}_1)^{\ast}$ with respect to $\Omega^{fs,k+1}$ \\ \hspace*{2em} $\mesh^{f,k+1} := (\mathcal{T}^{f,k+1}_1)^{\ast} \cup \mesh^{f,k+1}_2$ \\ \\ \hspace*{2em} \textbf{Solve fluid problem} \\ \hspace*{2em} Find $(\boldsymbol{u}^{f, k+1}_h,p^{f, k+1}_h)$ such that $\forall\, (\boldsymbol{v}^{f, k+1}_h,q^{f, k+1}_h) \in V_h^{f,k+1} \times Q_h^{f,k+1}$ \end{tabbing} \[ A^{f,k}_h(\boldsymbol{u}^{f, k+1}_h,p^{f, k+1}_h;\boldsymbol{v}^{f, k+1}_h, q^{f, k+1}_h) = L^{f,k+1}(\boldsymbol{v}^{f, k+1}_h,q^{f, k+1}_h) \] \begin{tabbing} \\ \hspace*{2em} \textbf{Update boundary traction functional} \\ \hspace*{2em} Define $L^{fs,k+1}(\cdot)$ by \end{tabbing} \[ L^{fs,k+1}(\refe{\boldsymbol{v}}^{s, k+1}_h) := R^{f,k+1}(\boldsymbol{u}^{f, k+1}_h,p^{f, k+1}_h;\boldsymbol{v}^{f, k+1}_h) \] \begin{tabbing} \\ \hspace*{2em} \textbf{Solve structure problem} \\ \hspace*{2em} Find $\refe{\boldsymbol{u}}^{s, k+1}_h$ such that $\forall\, \refe{\boldsymbol{v}} \in \widehat{V}_h^{s}$ \end{tabbing} \[ A^{s}_h(\refe{\boldsymbol{u}}^{s, k+1}_h, \refe{\boldsymbol{v}}) = L^{s}(\refe{\boldsymbol{v}}) + L^{fs,k+1}(\refe{\boldsymbol{v}}) \] \begin{tabbing} \\ \hspace*{2em} \textbf{Dynamic relaxation} \\ \hspace*{2em} Compute $\omega^{k+1}$ according to~\eqref{eq:compute-relax-parameter} \\ \hspace*{2em} $\refe{\boldsymbol{u}}^{s, k+1}_h := \omega^{k+1} \refe{\boldsymbol{u}}^{s, k+1}_h + (1-\omega^{k+1}) \refe{\boldsymbol{u}}^{s, k}_h$ \\ \\ \hspace*{2em} \textbf{Solve mesh problem} \\ \hspace*{2em} Find $\refe{\boldsymbol{u}}^{m, k+1}_h$ such that $\forall\, \refe{\boldsymbol{v}} \in \widehat{V}_h^{m}$ \end{tabbing} \begin{align*} A^{m}_h(\refe{\boldsymbol{u}}^{m, k+1}_h, \refe{\boldsymbol{v}}) &= L^{s}(\refe{\boldsymbol{v}}) \\ \refe{\boldsymbol{u}}^{m, k+1}_h &= \refe{\boldsymbol{u}}^{s, k+1}_h \text{on } \refe{\IFS} \end{align*} \begin{tabbing} \textbf{while } $\| \refe{\boldsymbol{u}}^{s, k+1}_h - \refe{\boldsymbol{u}}^{s, k}_h \| \leqslant $ \texttt{TOL} \end{tabbing} \caption{Fixed-point iteration.} \label{alg:fixed-point-iteration} \end{algorithm} \subsection{Dynamic relaxation} Let $U_{_S}^k$ denote the coefficient vector of the finite element approximation $\refe{\boldsymbol{u}}^{s,k}_h$ of \eqref{eq:structure-problem-weak-form} computed in the $k$-th iteration step. To accelerate the convergence of the iteration scheme, a relaxation step is introduced: \begin{align} U_{_S}^{k+1} := \omega_k U_{_S}^{k+1} + (1-\omega_k) U_{_S}^{k}, \label{eq:def:relaxation} \end{align} where the relaxation parameter $\omega_k$ is dynamically chosen in each iteration step. Here, we employed Aiken's method \cite{Kuttler2008,Kuttler2009} which is a simple scheme, yet it can greatly improve the convergence rate compared to a fix choice of $\omega_k$, as demonstrated by \citet{Kuttler2008,Kuttler2009}. Introducing the residual displacement $\Delta^k U_{_S}$ by \begin{align} \Delta^k U_{_S} := U_{_S}^k - U_{_S}^{k-1}, \label{eq:def-displacement-residual} \end{align} the new relaxation parameter $\omega_{k+1}$ is then computed by \begin{align} \omega_{k} = \max \left\{\omega_{\max}, \omega_{k-1} \left( 1 - \dfrac{\Delta^{k+1} U_{_S}}{\| \Delta^{k+1} U_{_S} - \Delta^{k} U_{_S}\|^2} \right) \right\}, \label{eq:compute-relax-parameter} \end{align} where $\omega_{\max}$ is a safety parameter chosen to avoid too large over-relaxation. The convergence of the fixed-point iteration might be accelerated further by employing more sophisticated schemes based on Robin--Robin coupling~\cite{Badia2008,Badia2009} or vector extrapolation~\cite{Kuttler2009}. \subsection{Computation of the boundary traction} Given the solution $\uF$ and a pressure solution $p^f$ of the fluid subproblem~\eqref{eq:fluid-momentum-pde}--\eqref{eq:fluid-bc2}, the incorporation of the fluid boundary traction into the weak formulation of the structure problem~\eqref{eq:structure-problem-weak-form} requires the evaluation of the so-called weighted fluid boundary traction on $\IFS$ defined by \begin{align} L^{fs}(\boldsymbol{v}) &= (\sigmaF(\uF,p^f)\cdot \boldsymbol{n}, \boldsymbol{v})_{\IFS}, \label{eq:fluid-traction-strong} \end{align} for test functions $\boldsymbol{v} \in V^{s}$. The functional~\eqref{eq:fluid-traction-strong} possesses various equivalent representations in the continuous case which are no longer equivalent when fluid velocity $\uF$ and pressure $\pF$ and test function $\boldsymbol{v}$ are replaced by their discrete counterparts $\uF_h$, $\pF_h$ and $\boldsymbol{v}_h \in V^s_h(\Omega)$, respectively. It has been observed by \citet{Dorok1995,John1997,Giles1997} that using~\eqref{eq:fluid-traction-strong} directly might lead to an inaccurate evaluation of the weighted boundary traction. In our work, we therefore employ an alternative formulation of the weighted boundary traction in the form \begin{align} L^{fs}(\boldsymbol{v}_h) &= (\sigmaF(\uF_h,\pF_h),\Ext(\boldsymbol{v}_h))_{\OF} - (\boldsymbol{f}^f, \Ext(\boldsymbol{v}_h))_{\OF}, \label{eq:fluid-traction-functional} \end{align} which was proposed and investigated by \citet{Giles1997} in the context of a posteriori error estimation. Here, $\Ext(\boldsymbol{v})$ is any function in $H^1(\Omega^{fs})$ such that $\Ext(\boldsymbol{v}_h)|_{\IFS} = \boldsymbol{v}_h$. Compared to the naive evaluation using~\eqref{eq:fluid-traction-strong}, the formulation~\eqref{eq:fluid-traction-functional} was shown to compute the weighted boundary traction more accurately and to greatly improve the convergence of stress related quantities such as the lift and drag coefficients. \section{Numerical results} We conclude this paper with two numerical tests, both in three spatial dimensions. The numerical experiments were carried out using the \rm{DOLFIN-OLM} library (\url{http://launchpad.net/dolfin-olm}). We first study the convergence rates for the finite element approximations of the fluid velocity, fluid pressure and structure displacement by constructing an artificial fluid--structure interaction problem possessing an analytical solution. Second, we consider the flow around an elastic flap immersed in a three-dimensional channel. \subsection{Software for overlapping mesh variational formulations} The assembly of finite element tensors corresponding to standard variational formulations on conforming, simplicial meshes, such as~\eqref{eq:structure-problem-weak-form}, involves integration over elements and possibly, interior and exterior facets. In contrast, the assembly of variational forms defined over overlapping meshes, such as~\eqref{eq:ah-olm}-\eqref{eq:jh-olm} and~\eqref{eq:Lh-olm}, additionally requires integration over cut elements and cut facets. These mesh entities are of polyhedral, but otherwise arbitrary, shape. As a result, the assembly process is highly non-trivial in practice and requires additional geometry related preprocessing, which is challenging in particular for three-dimensional meshes. As part of this work, the technology required for the automated assembly of general variational forms defined over overlapping meshes has been implemented as part of the software library \rm{DOLFIN-OLM}. This library builds on the core components of the FEniCS Project~\citep{LoggMardalEtAl2011,Logg2007}, in particular DOLFIN~\citep{LoggWells2010a}, and the computational geometry libraries \rm{CGAL}~\citep{cgal} and \rm{GTS}~\citep{gts}. \rm{DOLFIN-OLM} is open source and freely available from \url{http://launchpad.net/dolfin-olm}. There are two main challenges involved in the implementation: the computational geometry and the integration of finite element variational forms on cut cells and facets. The former involves establishing a sufficient topological and geometric description of the overlapping meshes for the subsequent assembly process. To this end, \rm{DOLFIN-OLM} provides functionality for finding and computing the intersections of triangulated surfaces with arbitrary simplicial background meshes in three spatial dimensions; this functionality relies on the computational geometry libraries \rm{CGAL} and \rm{GTS}. These features generate topological and geometric descriptions of the cut elements and facets. Based on this information, quadrature rules for the integration of fields defined over these geometrical entities are produced. The computational geometrical aspect of this work extends, but shares many of the features of, the previous work~\citep{Massing2012a}, and is described in more detail in the aforementioned reference. Further, by extending some of the core components of the FEniCS Project, in particular FFC~\citep{KirbyLogg2006,LoggOelgaardEtAl2011a} and UFC~\citep{AlnaesLoggEtAl2012a}, this work also provides a finite element form compiler for variational forms defined over overlapping meshes. Given a high-level description of the variational formulation, low-level C++ code can be automatically generated for the evaluation of the cut element, cut facet and surface integrals, in addition to the evaluation of integrals over the standard (non-cut) mesh entities. The generated code takes as input appropriate quadrature points and weights for each cut element or facet; these are precisely those provided by the \rm{DOLFIN-OLM} library. As a result, one may specify variational forms defined over finite element spaces on overlapping meshes in high-level \rm{UFL} notation \citep{AlnaesEtAl2012,Alnaes2011a}, define the overlapping fluid meshes $\{\mathcal{T}_0, \mathcal{T}^f_2\}$ and then invoke the functionality provided by the \rm{DOLFIN-OLM} library to automatically assemble the corresponding stiffness matrix. In particular, the numerical experiments presented below, employing the variational formulation defined by~\eqref{eq:stokes-olm}, have been carried out using this technology. \subsection{Convergence test} While numerical studies presented in \cite{MassingLarsonLoggEtAl2013} confirmed the theoretically predicted convergence rates for the overlapping mesh method for the pure flow problem presented in Section~\ref{ssec:fluid-discretization}, we here conduct a convergence study of the coupled FSI problem to verify the overall solution algorithm as described in Algorithm~\ref{alg:fixed-point-iteration}. To examine the convergence rates for the finite element approximations of the fluid velocity, fluid pressure and structure displacement, we construct a stationary FSI problem with a known analytical solution by employing the method of manufactured solutions as outlined in the following. The detailed analytical derivation of the fluid and structure related quantities are not included here to keep the presentation at an appropriate length, but can be obtained as an IPython based notebook available at \url{http://nbviewer.ipython.org/6291921}. In the reference configuration, the fluid domain $\refe{\Omega}^f$ consists of a straight tube of length $L = 1.0$ and diameter $R^f = 0.4$. We decompose $\refe{\Omega}^f$ into into a tube of radius $R^f_1 = 0.3$ and a cylinder annulus satisfying $0.3 \leqslant r \leqslant 0.4 = R^f_2$. The solid domain $\refe{\Omega}^s$ is given by a cylinder annulus of thickness $H^s = 0.1$ surrounding the fluid domain $\refe{\Omega}^f$. Using cylinder coordinates, the displacement $\refe{\boldsymbol{u}}^s$ of the solid domain is prescribed by a purely radial, $z$-dependent translation \begin{equation} \refe{\boldsymbol{u}}^s(r,\varphi,z) = H(z)\boldsymbol{e}_r, \label{eq:solid-radial-translation} \end{equation} where $H(z) = H^s 2 z (1-z)$. Correspondingly, the deformation of the fluid domain is determined by a radial stretching of the form \begin{equation} \refe{\boldsymbol{u}}^m(r,\varphi,z) = \rho(1 + H(z)/R^f)\boldsymbol{e}_r. \label{eq:fluid-radial-stretching} \end{equation} The reference and physical configuration of the various domains are depicted in Figure~\ref{fig:analytical-fsi-example-domains}. \begin{figure}[htb] \centering \includegraphics[width=0.45\textwidth]{pdf/analytical-fsi-reference-domains.pdf} \includegraphics[width=0.45\textwidth]{pdf/analytical-fsi-pysical-domains.pdf} \caption{Cross-section through the cylinder-symmetric reference (left) and physical (right) domains for the analytical FSI reference problem.} \label{fig:analytical-fsi-example-domains} \end{figure} To obtain a divergence-free velocity field in the final physical configuration, the fluid velocity is defined as a simple parabolic channel flow on the reference domain and then mapped to the physical domain via the Piola transformation induced by the fluid domain deformation~\eqref{eq:fluid-radial-stretching}. For the pressure, we simply choose $p(x,y,z) = 1-z$. Since the interface condition~\eqref{eq:final-fsi-condition-3} is not satisfied exactly, we introduce an auxiliary traction $\boldsymbol{t}_a$ given by the non-vanishing jump in the normal stresses: \begin{equation} \boldsymbol{t}_a = \bigl( \refe{\Pi}(\refe{\boldsymbol{u}}^s) (\refe{\boldsymbol{x}}) - J^m(\refe{\boldsymbol{x}}) \, \sigmaF (\boldsymbol{\phi}^m( \refe{\boldsymbol{x}})) (\boldsymbol{F}^m)^{-\top} (\refe{\boldsymbol{x}}) \bigr)\cdot \refe{\boldsymbol{n}}^s \quad \text{on } \refe{\Gamma}^{fs}. \label{eq:final-fsi-condition-3-with-auxiliary-traction} \end{equation} Regarding the remaining boundary parts, the solid displacement is uniquely determined by imposing the given displacement $\refe{\boldsymbol{u}}^s$ as a Dirichlet boundary condition on $\partial\refe{\Omega}^s \setminus \refe{\Gamma}^{fs}$. For the fluid problem, we prescribed the velocity profile on the inlet and impose the zero pressure on the outlet. In the reference configuration, a discretization of the solid domain $\refe{\Omega}^s$ and the fluid domain $\refe{\Omega}^f_2$ is provided by two fitted and conforming meshes $\refe{\mathcal{T}}^s$ and $\refe{\mathcal{T}}^f_2$, respectively, while the fluid domain $\refe{\Omega}^f_1$ is represented by a structured Cartesian mesh $\refe{\mathcal{T}}^f_1$ overlapped by the mesh $\refe{\mathcal{T}}^f_2$, see Figure~\ref{fig:analytical-fsi-example-velocity-pressure}. The numerical approximation of the fluid velocity, fluid pressure and structure displacement are then computed on a sequence of $4$ overlapping meshes. The mesh sizes of the initial meshes $\refe{\mathcal{T}}^f_1$, $\refe{\mathcal{T}}^f_2$, and $\refe{\mathcal{T}}^s$ are $0.246$, $0.14$ and $0.212$, respectively and each of the subsequent meshes is generated from the previous one by uniformly refining each mesh. Based on the manufactured exact solution, the experimental order of convergence (EOC) is then computed by \begin{align*} \text{EOC}(k) = \dfrac{\log(E_{k-1}/E_{k})}{\log(2)} \end{align*} where $E_k$ denotes the error of the numerical solution computed at refinement level $k$. The numerical experiment was conducted using $\nu^f = 0.001$ for the fluid viscosity and Lam\'e parameters given by \begin{equation} \mu = E/(2 + 2\nu), \quad \lambda = E \cdot \nu /((1 + \nu)(1 - 2\nu)) \label{eq:lame-constants} \end{equation} in $\Omega^s$ with $E = 10$ and $\nu = 0.3$. \begin{figure}[htb] \centering \includegraphics[width=0.45\textwidth]{png/analytical_fsi_velocity_bg_mesh_ref_2.png} \includegraphics[width=0.45\textwidth]{png/analytical_fsi_velocity_all_meshes_ref_2.png} \hspace{1ex} \includegraphics[width=0.45\textwidth]{png/analytical_fsi_pressure_bg_mesh_ref_2.png} \includegraphics[width=0.45\textwidth]{png/analytical_fsi_pressure_all_meshes_ref_2.png} \caption{Computed velocity (top) and pressure (bottom) solution on the fixed fluid background mesh $\mathcal{T}^f_1$ (left) and entire overlapping fluid mesh $\{\mathcal{T}^f_1, \mathcal{T}^f_2\}$ (right) for the analytical FSI problem.} \label{fig:analytical-fsi-example-velocity-pressure} \end{figure} \begin{figure}[htb] \label{fig:analytical-fsi-example-displacements} \centering \includegraphics[width=0.45\textwidth]{png/analytical_fsi_structure_displacement_mesh_ref_2.png} \includegraphics[width=0.45\textwidth]{png/analytical_fsi_structure_mesh_displacement_mesh_ref_2.png} \caption{Displacements for the analytical FSI reference problem. Left: Structure displacement of the solid tube. Right: Displacement of the fluid mesh added.} \end{figure} For the penalty parameters in the stabilized overlapping mesh method for the fluid problem, we pick $\gamma=10$ and $\delta = 0.5$. Since the overall computational time is dominated by the assembly and solution of the fluid system, the displacement field is conveniently solved using a direct solver, while the linear system arising from the fluid problem is solved by applying a transpose-free quasi-minimal residual solver with an algebraic multigrid preconditioner. Using continuous piecewise linear functions for the approximation of both the fluid velocity, fluid pressure and the structure displacement, the theoretically predicted convergence rate for a corresponding uncoupled problem is at least $1.0$ when measuring the velocity and displacement error in the $H^1$-norm and the pressure error in the $L^2$-norm. Note that it is common to observe a higher experimental order of convergence of $\sim 1.5$ for the pressure approximation when stabilized, equal-order interpolation elements are used to discretize the flow problem. Assuming at most quadratic convergence of the displacement solution in the $L^2$-norm, the $L^2$-error will be reduced by approximately $0.5^{2\cdot3} \approx 0.016$ after 3 uniform mesh refinements. To not pollute the overall convergence rate by the iteration error, we therefore chose $tol = 0.001$ for the relative $L^2$-error between two consecutive displacement solutions computed in the iteration loop. With the given tolerance, the Dirichlet-Neumann iteration converged after $5$-$7$ iteration for each refinement level. The resulting errors for the sequence of refined meshes are summarized in Table~\ref{tab:convergence-rates}. For the fluid velocity and fluid pressure the observed convergence rates are in agreement with the theoretical error decrease expected from an uncoupled problem. For the solid displacement the observed convergence rates between $1.46$-$1.9$ for the $H^1$-error are better than the theoretically expected rate of $\sim 1$. \begin{table}[htb] \centering \begin{tabular}{l|lr|lr|lr} \toprule Refinement & $\| u^f_h - u^f \|_{1} $ & EOC & $\| p^f_h - p^f \|_{0} $ & EOC & $\| u^s_h - u^s \|_{1} $ & EOC \\ \midrule 0 & $1.01188$ & - & $3.61948e-03$ & - & $3.87181e-04$ & - \\ 1 & $0.51000$ & 0.99 & $1.55216e-03$ & 1.22 & $1.40771e-04$ & 1.46 \\ 2 & $0.21912$ & 1.22 & $3.70746e-04$ & 2.06 & $4.39062e-05$ & 1.68 \\ 3 & $0.12485$ & 0.81 & $1.29430e-04$ & 1.52 & $1.17800e-05$ & 1.9 \\ \bottomrule \end{tabular} \vspace{1ex} \caption{Convergence rates of the overlapping mesh finite element method for the analytical FSI problem.} \label{tab:convergence-rates} \end{table} \subsection{Flow around an elastic flap} In the second numerical example, we consider a channel flow around an elastic flap for different orientations of the flap with respect to the channel geometry. Here, the developed method and techniques can play out their full strength as the overlapping mesh approach handles large deformation within a single simulation easily. As an additional benefit, our proposed scheme allows to seamlessly reposition the flap for a series of numerical experiments and thus has great potential for future applications in design and optimization processes which involve fluid--structure interaction problems in their forward simulation, see for instance ~\cite{LundMollerJakobsen2003, EguzkitzaHouzeauxCalmetEtAl2014}. Within the channel domain $\Omega = [0,L] \times [0,W] \times [0,H]$ with $L = 2.5$, $W = H = 0.41$, the bottom side of the flap of dimensions $L^s = 0.06$, $W^s=0.2$, $H^s=0.24$ is centered around the point $(L/2, W/2, 0)$. In the first numerical experiment, the flap is clamped on the boundary $[(L-L^s)/2,(L+L^s)/2],[(W-W^s)/2,(W+W^s)/2] \times \{0\}$, while the flap is rotated $65^{\circ}$ around the $z$-axis in a second experiment. For the numerical experiment, we assume that the flow can be described by the Stokes equations with fluid viscosity $\nu^f = 0.001$, while the flap is modeled as an hyperelastic material satisfying the St.~Venant--Kirchhoff constitutive equation~\eqref{eq:venant-kirchhoff-material} with the Lam\'e constants $\mu,\,\lambda$ defined by~\eqref{eq:lame-constants} for $E^s = 15$ and $\nu^s = 0.3$. Finally, we prescribe the inflow profile $\boldsymbol{u}^f = (16\cdot0.45y(W-y)z(H-z), 0, 0)$ at the inlet $\{0\} \times [0,W] \times [0,H]$, a ``do-nothing'' boundary condition given by $\nu \partial_{\boldsymbol{n}} u - p\boldsymbol{n} = 0$ at the outlet $\{L\} \times [0,W] \times [0,H]$ and a no-slip condition $\boldsymbol{u} = \boldsymbol{0}$ elsewhere on the boundary. The numerical results for aligned and rotated flaps are shown in Figure~\ref{fig:flap-channel-solutions-angle-0}. We especially note the smooth transition of the velocity and pressure solutions from fluid background $\mathcal{T}^f_1$ to the solid-surrounding fluid mesh $\mathcal{T}^f_2$; the interface is not visible. The meshes used for simulation of the rotated flap are shown in Figure~\ref{fig:meshes}. \begin{figure}[htb] \begin{center} \includegraphics[width=0.90\textwidth]{png/flap_channel_angle_15_meshes.png} \caption{Background fluid mesh, structure mesh and its surrounding fluid mesh in the reference configuration.} \label{fig:meshes} \end{center} \end{figure} \begin{figure} \begin{center} \includegraphics[width=0.51\textwidth]{png/flap_channel_angle_0_velocity_cropped.png} \includegraphics[width=0.48\textwidth]{png/flap_channel_angle_0_deformation_pressure_cropped.png} \\ \includegraphics[width=0.52\textwidth]{png/flap_channel_angle_15_velocity_cropped.png} \includegraphics[width=0.47\textwidth]{png/flap_channel_angle_15_deformation_pressure_cropped.png} \caption{Flow around an elastic flap for two different flap orientations. Left: Magnitude and streamlines of the velocity approximation in an $x$-$z$ (top) and $x$-$y$ (bottom) cross-section. The transparent block around the gray-colored flap visualizes the fluid mesh $\mathcal{T}^f_2$ surrounding the structure. The streamlines within the $\mathcal{T}^f_2$ are drawn slightly thicker to illustrate the smooth transition of the velocity approximation from the outer to the inner fluid domain. Right: Pressure distribution and magnitude of the structure displacement.} \label{fig:flap-channel-solutions-angle-0} \end{center} \end{figure} \section{Conclusions} We presented a Nitsche-based cut and composite mesh method for fluid--structure interaction problems. The method utilizes a Nitsche type coupling between two fluid meshes: one fixed background mesh and one moving overlapping fluids mesh which is fitted to the boundary of a hyperelastic object and deforms with the object. The fluid--fluid coupling is monolithic in the sense that it manufactures a coupled system involving both the underlying and overlapping degrees of freedom. In previous work, \cite{MassingLarsonLoggEtAl2013}, we have shown that the coupling is stable and that the solution has optimal order convergence for a stationary model problem. To solve for the steady state solution of a fluid--structure interaction problem with large elastic deformations, we consider a fixed-point iteration where we solve for the fluid, compute a boundary traction for the solid, solve for the solid, solve for the mesh motion of the overlapping fluid mesh, and finally update the geometry. This involves computing new intersections between underlying and overlapping meshes. Employing a provably stable overlapping mesh method for fluid-fluid coupling, the proposed scheme for the fluid--structure problem is guaranteed to be robust and insensitive to the overlap configuration. We verified the expected convergence rates for a model problem with a manufactured solution and demonstrated the flexibility of our approach by computing the steady state solution for an elastic flap in a channel at two different orientations. It should be noted that the overlapping mesh method allows the flap to be repositioned in the channel without requiring the generation of a single \emph{conforming} fluid mesh for each configuration. Only an element-wise, local representation of the cut cells near the interface together with some appropriate quadrature schemes are required, see for instance~\cite{Massing2012a}. Future work involves extending our method to fully time-dependent flow governed by the incompressible Navier-Stokes equations. We note that the nonlinear convection term can be handled in our setting using a discontinuous Galerkin coupling with up-winding and that, from a computational point of view, taking a time step is closely related to taking one step in our fixed-point iteration algorithm. Another area of interest is the direct coupling between fluids and solids. \section*{Acknowledgments} This work is supported by an Outstanding Young Investigator grant from the Research Council of Norway, NFR 180450. This work is also supported by a Center of Excellence grant from the Research Council of Norway to the Center for Biomedical Computing at Simula Research Laboratory. The authors would like thank the anonymous referee for the valuable comments and suggestions which helped to improve the presentation of this work. \bibliographystyle{plainnat}
\section{Introduction} Let $(X,\omega)$ be a Riemann surface with a holomorphic $1$-form on it. The set of all such pairs forms an algebraic variety $\cH(\kappa)$ called a stratum, where $\kappa$ encodes the multiplicities of the zeros of $\omega$. The stratum carries a natural action of the group $\SL_2\bR$ which is of transcendental nature. A stratum $\cH(\kappa )$ has natural charts to complex affine spaces. The coordinates are the periods of $\omega$ on $X$, thus in $\bC$. After identifying $\bC$ with $\bR^2$, the action of $\SL_2\bR$ is the standard one on each coordinate individually. In particular, the statum carries a natural Lebesgue-class measure which is invariant under $\SL_2\bR$. The finiteness of the measure was proved by Masur \cite{Masur} and Veech \cite{Veech}. The action of $\SL_2\bR$ and knowledge of invariant measures can be applied to study other dynamical systems. For interval exchange transformations this started in the work of Masur and Veech \cite{Masur, Veech}. A starting point for applications to polygonal billiards was by Kerckhoff-Masur-Smillie in \cite{KMS}. Some recent applications involve a detailed analysis of the wind-tree model by Hubert-Leli{\`e}vre-Troubetzkoy \cite{windtree}. For a comprehensive introduction to the subject, see the survey of Zorich \cite{Zorich_survey}. For more precise applications, especially to concrete examples, one needs to understand all possible invariant measures. For instance, polygonal billiards correspond to a set of Masur-Veech measure zero. Recent results of Eskin and Mirzakhani \cite{EM} show that finite ergodic invariant measures are rigid and in particular are of Lebesgue class and supported on smooth manifolds. In further work with Mohammadi \cite{EMM} they show that many other analogies with the homogeneous setting and Ratner's theorems hold. The $\SL_2\bR$-invariant measures give rise to \emph{affine invariant manifolds}. These are complex manifolds which are given in local period coordinates by linear equations. It was shown by Wright \cite{Wright_field} that the linear equations can be taken with coefficients in a number field. Note that \emph{finite} $\SL_2\bR$-invariant measures are supported on real codimension $1$ hypersurfaces inside affine manifolds, the issue arising from the action of scaling by $\bR^\times$. The affine invariant manifolds are then closed $\GL_2\bR$-invariant sets. In fact, by \cite{EMM} the closure of any $\GL_2\bR$-orbit is an affine manifold. Period coordinates are transcendental and so apriori affine manifolds, which are given by linear equations, are only complex-analytic submanifolds. In this paper, we prove the following result (see \autoref{thm:algebraic}). \begin{theorem} Affine invariant manifolds are algebraic subvarieties of the stratum $\cH(\kappa)$, defined over $\conj{\bQ}$. \end{theorem} I am grateful to Curtis McMullen for suggesting the next result (see also \cite[Theorem 1.1]{McMullen_classification} for a much more precise result in genus $2$). \begin{corollary}[see \autoref{rmk:Teich_disk}] \label{cor:Teich_disk} Let $\bH\to \cM_g$ be a \Teichmuller disk. Then its closure (in the standard topology) inside $\cM_g$ is an algebraic subvariety of $\cM_g$. \end{corollary} The lowest-dimensional affine manifolds are \Teichmuller curves. It was proved they are algebraic by Smillie and Weiss \cite[Proposition 8]{Smillie_Weiss}; a different sketch of proof (attributed to Smillie) is in \cite{Veech_closure}. That they are defined over $\conj{\bQ}$ was proved by McMullen \cite{McMullen_def_Q}. It was proved by M{\"o}ller in \cite{Moller_VHS} that they are defined over $\conj{\bQ}$ after embedding into a moduli space of abelian varieties. \Teichmuller curves and higher-dimensional $\SL_2\bR$-invariant loci also have interesting arithmetic properties. McMullen has related in \cite{McMullen_Hilbert_mod_surF} \Teichmuller curves in genus $2$ with real multiplication. He also gave further constructions using Prym loci \cite{McMullen_prym}. In genus $2$ algebraicity follows from a complete classification of invariant loci by McMullen \cite{McMullen_classification}. In the stratum $\cH(4)$ algebraicity is known by results of Aulicino, Nguyen and Wright \cite{ANW_H^4, NW}. Lanneau and Nguyen have also done extensive work on Prym loci in genus $3$ and $4$ \cite{LN_periodicity, LN_Prym_g3_G4}. Techniques from variations of Hodge structures were introduced by M{\"o}ller, starting in \cite{Moller_VHS}. In particular, he showed that \Teichmuller curves always parametrize surfaces with Jacobians admitting real multiplication on a factor. He also showed that over a \Teichmuller curve, the Mordell-Weil group of the corresponding factor is finite \cite{Moller_torsion}. In particular, zeros of the $1$-form are torsion under the Abel-Jacobi map. See also \cite{Moller_Linear_mnfds} for further results. The results in \cite{ssimple} show that on affine manifolds, the topological decomposition of cohomology (e.g. the local systems from \cite{Wright_field}) are compatible with the Hodge structures. As a consequence, affine manifolds also parametrize Riemann surfaces with non-trivial endormophisms, typically real multiplication on a factor (see \cite[Thm. 7.3]{ssimple}). This paper extends M{\"o}ller's torsion result to affine manifolds. The precise statement and definitions are in \autoref{subsec:combinig_splittings} and \autoref{thm:torsion}. \begin{theorem} Let $\cM$ be an affine invariant manifold, parametrizing Riemann surfaces with real multiplication by the order $\cO$ on a factor of the Jacobians. Then $\cM$ carries a natural local system $\Lambda$ of $\cO$-linear combinations of the zeros of the $1$-form (see \autoref{eqn:Lambda_iota_def}) and a twisted Abel-Jacobi map (see \autoref{def:twisted_cycle}) $$ \nu : \Lambda \to \Jac_\bZ\left(\oplus_\iota H^1_\iota\right) $$ The range is the factor of Jacobians admitting real multiplication (see \autoref{eqn:H1_decomp}). Then the image of $\nu$ always lies in the torsion of the abelian varieties. \end{theorem} \begin{remark} \leavevmode \begin{itemize} \item[(i)] The expression ``real multiplication by $\cO$'' is used in a rather loose sense. It means that the ring $\cO$ maps to the endomorphisms of a factor of the Jacobian. The factor of the Jacobian is always nontrivial, as it contains at least the part coming from the $1$-form $\omega$. The ring $\cO$ could be $\bZ$, and the factor could be the entire Jacobian. \item[(ii)] The local system $\Lambda$ can be trivialized on a finite cover of the stratum, and is defined as follows. The tangent space to the stratum contains the relative cohomology classes that vanish on absolute homology, denoted $W_0$. The tangent space $T\cM$ of the affine manifold intersects it in a sublocal system (over $\cM$), denoted $W_0\cM$. The dual of $W_0$, denoted $\check{W_0}$, is canonically identified with linear combinations of the zeros of the $1$-form with zero total weight. Then $\Lambda$ is an $\cO$-submodule of $(W_0\cM)^\perp\subset \check{W_0}$, i.e. of the annihilator of $W_0\cM$. \noindent Indeed, by results of Wright \cite{Wright_field}, $W_0\cM$ and thus $(W_0\cM)^\perp$ are defined over $k$ - the field giving real multiplication. Since $\check{W}_0$ carries a $\bZ$-structure, extending scalars to $\cO$, define $\Lambda:=\check{W}_0(\cO)\cap (W_0\cM)^\perp(k)$, where $A(R)$ denotes the $R$-points of $A$. \item[(iii)] When $W_0\cM$ is empty, e.g. for \Teichmuller curves, $\Lambda$ coincides with $\check{W}_0(\cO)$. In particular, it contains (up to finite index) all the $\bZ$-linear combinations of zeroes of the $1$-form, with total weight zero; on them $\nu$ is the usual Abel-Jacobi map. The extension of $\nu$ to $\cO$-linear combinations of zeros uses the $\cO$-action on the Jacobian factor. \end{itemize} \end{remark} \begin{remark} The result on torsion can be described concretely using periods of $1$-forms. Note that it refers, in particular, to $1$-forms other than $\omega$; describing them using the flat structure does not seem immediate. Let $\{r_j\}_j$ be formal integral combinations of the zeroes of $\omega$, such that the coefficients of each $r_j$ add up to $0$. They can be lifted to actual relative cycles on the surface, denoted $r_j'$ (these are now actual curves that connect zeroes of $\omega$). Let also $\{a_i,b_i\}_{i=1}^g$ denote an integral basis of the first homology of the surface. Suppose $\sum_j c^j r_j$ is an element of $\Lambda$, where $c_j$ are elements of $\cO$. The condition that its image is torsion under the twisted cycle map $\nu$ is equivalent to the following: There exist $\alpha_i,\beta_i\in \bQ$ such that whenever $\omega_l\in H^{1,0}_{\iota_l}$ is a holomorphic $1$-form, we have \begin{align} \sum_j \iota_l (c^j)\int_{r_j'}\omega_l = \sum_i \left(\alpha_i\int_{a_i} \omega_l + \beta_i \int_{b_i}\omega_l\right) \end{align} In other words, the absolute and relative periods of $\omega_l$ satisfy some linear relations. The coefficients $\iota_l(c^j)$ vary with the embedding $\iota_l$ corresponding to the subspace in which $\omega_l$ lives. \end{remark} \begin{remark} \label{rmk:Teich_disk} The algebraicity result also applies to strata of quadratic differentials. Indeed, these embed via the double-covering construction to strata of holomorphic $1$-forms. An affine invariant submanifold of a stratum of quadratic differentials can thus be viewed as one in a stratum of $1$-forms. In particular, for \autoref{cor:Teich_disk}, it suffices to restrict to \Teichmuller disks coming from holomorphic $1$-forms. Over the moduli space of curves $\cM_g$ we have the Hodge bundle $\cH_g$ whose fibers are the holomorphic $1$-forms. We can consider the projectivization of $\cH_g$ and therefore the proper projection $\bP\cH_g\to \cM_g$. The stratum $\cH(\kappa)$ is a subvariety of $\cH_g$, and we can also quotient by the $\bC^\times$-action to obtain $\bP\cH(\kappa)\subseteq \bP\cH_g$. Now a \Teichmuller disk $f:\bH\to \cM_g$ as in \autoref{cor:Teich_disk} lifts to $\tilde{f}:\bH \to \bP\cH(\kappa)\subseteq \bP\cH_g$. Combining Theorem 2.1 from \cite{EMM} and \autoref{thm:algebraic} we find that the (Zariski and usual) closure of $\tilde{f}(\bH)$ in $\cH(\kappa)$ is an algebraic variety $\bP\cM$. We can further take its (Zariski and usual) closure inside $\bP\cH_g$ to find that it is also an algebraic variety $\bP\overline{\cM}\subseteq \bP\cH_g$. Note that the Zariski and usual closure of a quasi-projective set coincide. Now, the topological closure of $f(\bH)$ will agree with the projection of the topological closure of $\tilde{f}(\bH)$ inside $\bP\cH_g$, which is $\bP\overline{\cM}$. This follows from the properness of $\bP\cH_g\to \cM_g$. Properness also ensures that the projection of an algebraic variety is still a variety, so \autoref{cor:Teich_disk} follows. \end{remark} \paragraph{Outline of the paper.} \autoref{sec:alg_special_case} proves in a special case that affine invariant manifolds are algebraic. This special case occurs when the tangent space of the affine manifold contains all relative cohomology classes. The proof only uses results from \cite{ssimple}. \autoref{sec:MHS} contains basic definitions and constructions about mixed Hodge structures. We only describe the small part of the theory that is necessary for our arguments. The proofs in later sections use this formalism, and we also include some concrete examples throughout. One does not need to be acquainted with the full theory to follow the arguments. \autoref{sec:splittings} contains the main technical part. It proves that certain sequences of mixed Hodge structures are \emph{split}, i.e. as simple as possible. This uses the negative curvature properties of Hodge bundles. \autoref{sec:algebraicity_torsion} combines the previous results to deduce the Torsion \autoref{thm:torsion}. This is then used to prove the Algebraicity \autoref{thm:algebraic}. \begin{remark}[\textbf{Self-intersections}] \label{rmk:self_inters} Affine manifolds are only immersed in a stratum (see \cite[Def.~1.1]{EM}), and could have locally finitely many self-intersecting sheets. Thus, any affine manifold $\cM$ can be written as the union $\cM=\cM_0\coprod \cM'$ where $\cM_0$ is a smooth open subset of $\cM$ and $\cM'$ is a lower-dimensional proper closed $\GL_2^+(\bR)$-invariant affine manifold (possibly disconnected or with self-intersections). Moreover the topological closure of $\cM_0$ contains $\cM'$. Since $\dim_\bC\cM' < \dim_\bC \cM$, it follows that $\cM_0$ is connected if $\cM$ is. The arguments in \autoref{sec:alg_special_case} and \autoref{sec:algebraicity_torsion} about algebraicity apply locally on $\cM_0$ and identify it (locally) with a quasi-projective variety. Thus, assuming by induction that $\cM'$ is quasi-projective, they show that $\cM_0$ is quasi-projective inside $\cH\setminus \cM'$. Again, since $\cM'$ is quasi-projective and contained in the (topological) closure of $\cM_0$, it follows that $\cM$ is quasi-projective inside $\cH$. \end{remark} \paragraph{Orbifolds.} All the arguments are made in some finite cover of a stratum, to avoid orbifold issues. In particular, period coordinates are well-defined and the zeros of the $1$-form are labeled. The results are invariant under passing to finite covers. \paragraph{Acknowledgments.} I would like to thank my advisor Alex Eskin, who was very helpful and supportive at various stages of this work, and in particular about the paper \cite{ssimple} whose methods are used here. I have also benefited a lot from conversations with Madhav Nori, especially on the topic of mixed Hodge structures. I also had several conversations on the topic of algebraicity with Alex Eskin and Alex Wright. In particular, Alex Wright explained his (unpublished) result that the torsion and real multiplication theorems of M\"{o}ller characterize \Teichmuller curves and suggested that finding and proving some generalization of the torsion theorem could imply algebraicity. I have also discussed and received very useful feedback on an earlier draft of this paper from both Alex Eskin and Alex Wright. I am very grateful for their feedback and the numerous insights they shared with me. I am also grateful to Curtis McMullen for suggesting \autoref{cor:Teich_disk}. \section{Algebraicity in a particular situation} \label{sec:alg_special_case} In this section, we prove a special case of algebraicity. It only requires results of \cite{ssimple}. In this special case the location of the tangent space to an affine manifold can be described precisely. \paragraph{Setup.} Consider an affine invariant manifold $\cM$ in some stratum $\cH(\kappa)$. We omit $\kappa$ from the notation, and refer to the stratum as $\cH$. Let $T\cH$ be the tangent bundle of the stratum, and let $W_0\subset T\cH$ be the subbundle corresponding to the purely relative cohomology classes. The survey of Zorich \cite[Section 3]{Zorich_survey} provides a clear and detailed exposition of these objects. The purpose of this section is to prove the following result. \begin{proposition} \label{prop:alg_special_case} Suppose that everywhere on $\cM$ we have $W_0\subset T\cM$, where $T\cM$ denotes the tangent bundle of $\cM$. Then $\cM$ is a quasi-projective algebraic subvariety of $\cH$. \end{proposition} Before proceeding to the proof, we recall some facts about the local structure on a stratum $\cH$. In particular, we discuss the way in which the relative cohomology groups $H^1(X,(\omega)_{red};\bC)$ provide local coordinates. These arise from the Gauss-Manin connection and the tautological section, which assigns to $(X,\omega)$ the cohomology class of $\omega$. \paragraph{Some preliminaries.} As explained in \autoref{rmk:self_inters}, it suffices to argue locally in the open part of $\cM$ where there are no self-intersections. Focus on a small neighborhood in $\cH$ of some $(X_0,\omega_0)\in\cM$, denoted $N_\epsilon(X_0,\omega_0)$. Introduce the identification coming from parallel transport (i.e. the Gauss-Manin connection) on relative cohomology $$ GM_{(X,\omega)}:H^1(X,(\omega)_{red};\bC)\toisom H^1(X_0,(\omega_0)_{red};\bC) $$ This is defined for all $(X,\omega)\in N_\epsilon(X_0,\omega_0)$ and $(\omega)_{red}$ denotes the zeroes of $\omega$, forgetting the multiplicities (i.e. the reduced divisor). Recall that for $(X,\omega)$ we have the natural element $\omega\in H^1(X,(\omega)_{red};\bC)$ viewing $\omega$ as a relative cohomology class. Call this the \emph{tautological section} $\omega:\cH\to T\cH$. Period coordinates are then described by the tautological section: \begin{align} \Pi:N_\epsilon(X_0,\omega_0)&\to H^1(X_0,(\omega_0)_{red};\bC) \label{eqn:per_coord}\\ (X,\omega) &\mapsto GM_{(X,\omega)}\omega\notag \end{align} Recall that we have a short exact sequence of vector bundles over $\cH$ $$ 0\to W_0\to T\cH \overset{p}{\to}H^1\to 0 $$ Here $H^1$ is the bundle of absolute first cohomology of the underlying family of Riemann surfaces, and $W_0$ is the purely relative part as before. It is proved by Wright in \cite[Theorem 1.5]{Wright_field} that over $\cM$ the local system of cohomology decomposes as \begin{align} \label{eqn:H1_decomp} H^1=\left(\bigoplus_{\iota} H^1_{\iota}\right)\oplus V \end{align} The summation is over embeddings $\iota$ of a number field $k$ in $\bC$, and there is a distinguished real embedding $\iota_0$. We then have $p(T\cM)=H^1_{\iota_0}$. \begin{proof}[Proof of \autoref{prop:alg_special_case}] In period coordinates in the neighborhood $N_\epsilon(X_0,\omega_0)$ the property $p(T\cM)=H^1_{\iota_0}$ translates to the statement \begin{align} \label{eqn:loc_def_M} \Pi(\cM\cap N_\epsilon) \subseteq p^{-1}(H^1_{\iota_0})_{(X_0,\omega_0)}\subseteq T\cH_{(X_0,\omega_0)} \end{align} Because $T\cM$ contains all the purely relative cohomology classes, locally $\cM$ coincides with an open subset of the middle space above. An equivalent way to state the above local description of $\cM$ is to say \begin{align} \label{eqn:cM_subset_eigenform_locus} \cM\cap N_\epsilon =\lbrace{(X,\omega)\; \vert \;p(GM_{(X,\omega)}\omega)\in H^1_{\iota_0} (X_0,\omega_0)\rbrace} \end{align} So over $\cM$ the section $p(\omega)$ must belong to the local system $H^1_{\iota_0}$. Note the local system $H^1_{\iota_0}$ cannot be globally defined on $\cH$. However, in the neighborhood $N_\epsilon$ one can still define it using the Gauss-Manin connection. \noindent\textbf{Checking algebraicity.} We also know by \cite[Thm. 1.6, Thm. 7.3]{ssimple} that each $H^1_\iota$ carries a Hodge structure. Moreover, for each $a\in k$ we have the operator \begin{align} \rho(a)=\left(\bigoplus_\iota \iota(a)\right)\oplus 0 \end{align} which acts by the corresponding scalar on each factor of the decomposition \eqref{eqn:H1_decomp}. These operators give real multiplication on the Jacobians, with $\omega$ as an eigenform. As a consequence, there is an order $\cO\subset k$ which acts by genuine endomorphisms of the Jacobian factor obtained via \eqref{eqn:H1_decomp}. Moreover, the isomorphism class of the corresponding $\bZ$-lattice, viewed as an $\cO$-module, is constant on $\cM$. Recall that we are working on the open subset of $\cM$ where there are no self-intersections, and this is still connected if $\cM$ is (see \autoref{rmk:self_inters}). Define $\cN'$ to be the locus in $\cH$ of $(X,\omega)$ which admit real multiplication by $\cO$ on a factor of the Jacobian, with the same $\cO$-module structure on the integral lattice of the factor, and with $\omega$ as an eigenform. This is a finite union of algebraic subvarieties of $\cH$, since it is the preimage of such a collection under the period map to $\bP(H^{1,0})$, which is the projectivization of the Hodge bundle over $\cA_g$. See \autoref{rmk:finiteness_in_A_g} for an explanation. The discussion above, in particular \autoref{eqn:cM_subset_eigenform_locus}, gives that $\cM\subseteq \cN'$. In the neighborhood $N_\epsilon$ of $(X_0,\omega_0)$ let $\cN$ be one of the irreducible components of $\cN'$ which contains $\cM$. We will check this component is unique, and coincides with $\cM$. Recall the defining conditions of $\cN'$ in $N_\epsilon$: \begin{align*} N_\epsilon\cap \cN' = \{(X,\omega)\vert \forall a \in k &\textrm{ we have } \rho(a) \textrm{ is of Hodge type }(0,0)\\ &\textrm{ and } \rho(a)\omega = \iota_0(a)\omega\} \end{align*} The local systems $H^1_\iota$ are defined on $\cN$, but are extended to $N_\epsilon$ using the flat connection. They serve as ``eigen-systems'' for the action of $\rho(a)$, which itself is defined in $N_\epsilon$ using the flat connection. We then have the containment $$ N_\epsilon \cap \cN' \subseteq \{(X,\omega)\; \vert\; \forall a\in k \textrm{ we have }\rho(a)\omega = \iota_0(a)\omega\} $$ However, we also have the (local) equality which follows from the identifications via parallel transport $$ \{(X,\omega)\vert \forall a\in k, \rho(a)\omega=\iota_0(a)\omega\} = \{(X,\omega)\vert p\left(\Pi(X,\omega)\right)\in H^1_{\iota_0, (X_0,\omega_0)} \} $$ Looking back at the local definition of $\cM$ given by the inclusions \eqref{eqn:loc_def_M}, we see that this locus is exactly $\cM$. So we found that locally near $(X_0,\omega_0)$ we have $$ \cM\subseteq \cN\subseteq \cM $$ This finishes the proof that $\cM$ is an algebraic subvariety of $\cH$. \end{proof} \begin{remark} \label{rmk:loc_syst_per_coord} In local period coordinates, requiring the section $\omega:\cH\to T\cH$ to be in some local system is the same as restricting to a (local) affine manifold. Algebraicity in the above theorem followed because we could identify the tangent space to $\cM$ with $p^{-1}(H^1_{\iota_0})$. In general, we need to know the precise location of the tangent space in relative cohomology. The next few sections deal with this question. \end{remark} \begin{remark} \label{rmk:finiteness_in_A_g} We now discuss the finiteness of the components of the eigenform locus for real multiplication. Assume the type of real multiplication is fixed, in other words the order $\cO$ and the isomorphism of $\cO$-lattice with symplectic form corresponding to the factor of the Jacobian. To prove finiteness of the eigenform locus, it suffices to prove finiteness of the real multiplication locus in $\cA_g$. Indeed, the eigenform locus is a projective space bundle over the real multiplication locus. Finiteness of the real multiplication locus will follow from the following general theorem of Borel and Harish-Chandra \cite{BHC}. Let $\Gamma$ be an arithmetic lattice in a $\bQ$-algebraic group $G$. Consider some representation $V$ of $G$, with a $\bZ$-structure compatible with $\Gamma$, and an integral vector $v\in V(\bZ)$ such that $G(\bR)\cdot v\subset V(\bR)$ is closed (equivalently, the stabilizer of $v$ is reductive). Then the set of integral points in the orbit $G(\bR)\cdot v$ form \emph{finitely many} classes under the action of $\Gamma$. Consider the decomposition of absolute cohomology given in \autoref{eqn:H1_decomp}, and abbreviate it as $H^1=M\oplus V$ where $M$ is the factor with real multiplication. Each factor contains a lattice, denoted $M(\bZ)$ and $V(\bZ)$ respectively. Now consider the abstract $\cO\oplus\bZ$-module $M\oplus V$, equipped with the compatible symplectic form (the $\bZ$-factor in $\cO\oplus\bZ$ acts on $V$ only). Associated to it is the period domain $\bH_{M,V}$ parametrizing pairs of abelian varieties $(A_M,A_V)$, with markings $M \xrightarrow{\sim} H^1(A_M)$ and $V\xrightarrow{\sim} H^1(A_V)$. Moreover, the markings should respect the symplectic forms and $\cO$ should act by endormorphisms of $A_M$. Consider possible embeddings $\phi: M(\bZ)\oplus V(\bZ) \into H^1_\bZ$, respecting the symplectic form. Choosing a basis of $M(\bZ)$ and $V(\bZ)$, such $\phi$ are determined by a collection of vectors in $H^1_\bZ$ subject to constraints coming from the symplectic pairing between the vectors. Equivalently, this is determined by a single vector in the direct sum of several $H^1_\bZ$'s, with the appropriate constraints. The set of real vectors subject to the same constraints forms a single orbit under $\Sp(H^1)(\bR)$. Therefore, by the Borel--Harish-Chandra theorem there are finitely many possible $\phi$ up to the action of $\Sp(H^1)(\bZ)$ on the target. Finally, each $\phi$ determines an embedding $I_\phi$ of the period domain $\bH_{M,V}$ into the Siegel space corresponding to $H^1$. The stabilizer of the image of $I_\phi$ acts with co-finite volume on this image. Since there are only finitely many $\Sp(H^1)(\bZ)$-equivalence classes of $\phi$'s, there are finitely many corresponding subvarieties in $\cA_g$. These subvarieties of $\cA_g$ parametrize abelian varieties with real multiplication by $\cO$ on a factor and $\cO$-module structure as the one corresponding to the affine manifold $\cM$. \end{remark} \section{Mixed Hodge structures and their splittings} \label{sec:MHS} This section contains background on mixed Hodge structures and their properties. The monograph of Peters and Steenbrink \cite{PetersSteenbrink} provides a thorough treatment. We include examples relevant to our situation. The full machinery, as developed e.g. by Carlson in \cite{Carlson}, is not strictly necessary. However, using this language streamlines some of the arguments. Throughout this section, we fix a ring $k$ such that $\bZ\subseteq k\subseteq \bR$. Most often, $k$ will be a field. \subsection{Definitions} First recall some standard definitions. \begin{definition} A $k$-Hodge structure of weight $w$ is a $k$-module $V_k$ and the data of a decreasing filtration by complex subspaces $F^\bullet$ on $V_\bC:=V_k\otimes_k \bC$ $$ \cdots\subseteq F^p\subseteq F^{p-1}\subseteq\cdots\subseteq V_\bC $$ The filtration is called the Hodge filtration and is required to satisfy $$ F^p\oplus \conj{F^{w+1-p}}=V_\bC $$ \end{definition} \begin{definition} A $k$-Mixed Hodge structure is a $k$-module $V_k$ with an increasing filtration $W_\bullet$ defined over $k\otimes_\bZ \bQ$ $$ \cdots W_n\subseteq W_{n+1}\subseteq \cdots \subseteq \left(V_k\otimes_\bZ \bQ\right) $$ and a decreasing filtration $F^\bullet$ on $V_\bC$ such that $F^\bullet$ induces a $k$-Hodge structure of weight $n$ on $$ \gr_n^{W} V = W_n/W_{n-1} $$ The filtration $W_\bullet$ is called the weight filtration. \end{definition} \begin{remark} We can take duals of (mixed) Hodge structures and overall, they form an abelian category. Negative indexing in the filtration is allowed. See \cite[Section 3.1]{PetersSteenbrink} for more background. For future use, we recall the definition of the dual Hodge structure. To describe the Hodge and weight filtrations on the dual of $V$, denoted $V^\vee$, let \begin{equation} \begin{aligned} \label{eqn:def_dual} F^{p+1} V^\vee &= \{\xi\in V^\vee\vert \xi(F^{-p} V)=0\}\\ W_{n-1} V^\vee &= \{\xi \in V^\vee\vert \xi(W_{-n})=0\} \end{aligned} \end{equation} \end{remark} \begin{remark} In the definition of mixed Hodge structures, the weight filtration was defined only after allowing $\bQ$-coefficients. However, if it came from a $\bZ$-module, the position of the lattice will be relevant. \end{remark} \begin{example} \label{eg:mhs01} Let $C$ be a compact Riemann surface and $S\subset C$ a finite set of points. On the relative cohomology group $H^1(C,S)$ we have a natural mixed Hodge structure with weights $0$ and $1$. This is the same as the compactly supported cohomology of the punctured surface $C\setminus S$. We have the exact sequence $$ 0\to W_0\into H^1(C,S)\onto H^1(C)\to 0 $$ The sequence is valid with any coefficients, so we consider it over $\bZ$. We have the canonical identification $W_0=\widetilde{H^0(S)}$ which is the reduced cohomology of the set $S$. Here is the mixed Hodge structure on $H^1(C,S)$. The weight filtration has $W_0$ defined by the exact sequence, and $W_1$ is the entire space. The holomorphic $1$-forms on the Riemann surface $C$ give also relative cohomology classes, so form a subspace $$ F^1\subset H^1_\bC(C,S) $$ This subspace maps isomorphically onto the holomorphic $1$-forms on $H^1_\bC(C)$. This describes the mixed Hodge structure, and according to Carlson \cite[Theorem A]{Carlson} it recovers the punctured curve in many cases. \paragraph{The dual picture.} We shall often work with duals, because the constructions are more natural. Dualizing the above sequence we find $$ 0\from W_0^\vee \from \check{H}^1(C,S)\from \check{H}^1(C)\from 0 $$ On the space $\check{H}^1(C,S)$ we have a mixed Hodge structure of weights $-1$ and $0$ (see \autoref{eqn:def_dual}). The space $W_{-1}$ is the image of $\check{H}^1(C)$ which is pure of weight $-1$. The space $W_0$ is everything. The Hodge filtration has $F^0\check{H}^1(C,S)$ equal to the annihilator of $F^1H^1(C,S)$. In particular it contains the image of $F^0\check{H}^1(C)$, which is the annihilator of $F^1H^1(C)$. Moreover, we have the (natural) isomorphism over $\bC$ $$ W_0^\vee\from F^0\check{H}^1(C,S)/F^0\check{H}^1(C) $$ \end{example} \subsection{Splittings} The concepts below were first analyzed by Carlson \cite{Carlson}, which provides more details. We work exclusively with mixed Hodge structures of weights $\{0,1\}$ and their duals, with weights $\{-1,0\}$. They are viewed as extensions of pure Hodge structures of corresponding weights. \autoref{eg:mhs01} describes the mixed Hodge structures that occur in later sections. \begin{definition} \label{def:split} Fix a ring $L$ with $k\subseteq L \subseteq \bR$. A $k$-mixed Hodge structure $$ 0\to W_0 \to E \to H^1\to 0 $$ is \emph{$L$-split} if the sequence, after extending scalars to $L$, is isomorphic as a sequence of $L$-mixed Hodge structures to $$ 0\to W_0 \to W_0\oplus H^1\to H^1\to 0 $$ The mixed Hodge structure in this sequence is the direct sum of the pure structures. The isomorphism is required to be defined over $L$, but it must map the weight and Hodge filtrations isomorphically. \end{definition} \begin{remark} \label{remark:splittings} \leavevmode \begin{enumerate} \item[(i)] The definition for splittings of duals is analogous. A mixed Hodge structure is $L$-split if and only if its dual is. \item[(ii)] Giving a splitting is the same as giving a map defined over $L$ $$ \sigma: H^1\to E $$ which is the identity when composed with projection back to $H^1$. It is required to map $F^1H^1$ isomorphically to $F^1E$. \item[(iii)] Mixed Hodge structures as above are always $\bR$-split. In the dual picture, we have the sequence $$ 0\from W_0^\vee \from E^\vee \from \check{H}^1\from 0 $$ Then $F^0E^\vee\cap \conj{F^0E^\vee}$ is a real subspace which maps isomorphically onto $W_0^\vee (\bR)$. Over $\bR$ we can thus lift $W_0^\vee$ inside $E^\vee$ using this subspace and this provides the splitting. \end{enumerate} \end{remark} \begin{example} \label{eg:splitting_ell_curve} It is more convenient to describe splittings of dual sequences, and below is the simplest example. Consider a pure weight $-1$ Hodge structure $\check{H}^1_\bZ=\langle a,b\rangle$ with filtration $F^0\check{H}^1=\langle a+\tau b\rangle$, where $\Im \tau>0$. This defines an elliptic curve \[ \Jac(\check{H}^1):=\check{H}^1_\bC/\left(F^0\check{H}^1 + \check{H}^1_\bZ \right) \cong \raisebox{0.1em}{$\bC$} / \raisebox{-0.1em} {$(\bZ \oplus\bZ\tau)$} \] Note that $\check{H}^1_\bC=F^0\check{H}^1 \oplus \conj{F^0\check{H}^1}$ (since $\tau\neq \conj{\tau}$) and in particular $\check{H}^1_\bZ\cap F^0\check{H}^1=\{0\}$ (since integral elements are invariant under complex conjugation). Consider now possible extensions of the form $$ 0\from W_0\from E \from \check{H}^1\from 0 $$ Assume $W_0$ is of $\bZ$-rank $1$, generated by $c$. Lift it to some $c_1\in E_\bZ$. It gives a map $$ \sigma_\bZ:W_0(\bZ) \to E_\bZ $$ Note that the lift $c_1$ is ambiguous, up to addition of terms $xa+yb$ with $x,y\in \bZ$. Here, the generators of $\check{H}^1$ are identified with their image inside $E$. The extra data on $E$ is a subspace $F^0E$ which contains $F^0\check{H}^1$, is complex two-dimensional, and maps surjectively onto $W_0$. Pick a vector $v\in F^0E$ which maps to $c$. It must be of the form $$ v=c_1 + \lambda a + \mu b $$ where $\lambda,\mu \in \bC$. Note that the lift $v$ is ambiguous, up to the addition of complex multiples of $a+\tau b$ (which generate $F^0\check{H}^1=F^0E\cap \ker (E\to W_0) $). This provides a second lift $$ \sigma_\bR:W_0(\bZ)\to E_\bC/F^0 \check{H}^1 $$ We can take the difference of $\sigma_\bZ$ and $\sigma_\bR$. Because projecting $\sigma_\bZ-\sigma_\bR$ back to $W_0$ is the zero map, their image must land in $\check{H}^1_\bC$. Taking into account also the ambiguity in the definition of $\sigma_\bZ$, we get a \emph{canonical} map $$ \sigma_\bZ-\sigma_\bR:W_0(\bZ)\to \check{H}^1_\bC/\left(F^0\check{H}^1 + \check{H}^1_\bZ \right) $$ So we get a canonical element of the elliptic curve associated to the Hodge structure $\check{H}^1$. This element is zero if and only if the sequence is $\bZ$-split. Indeed, the vanishing of this element means we could choose the lift $v$ above to be integral. The element is torsion in the elliptic curve if and only if the sequence is $\bQ$-split. It means we could choose $v$ above with rational coefficients in $a$ and $b$. \end{example} \subsection{Extension classes and field changes} \label{subsec:exts_fields} This section contains a discussion of algebraic facts needed later. The details of the constructions are available in \cite{Carlson} and \cite[Chapter 3.5]{PetersSteenbrink}. \paragraph{Jacobians and extensions} For a $\bZ$-Hodge structure $H^1$ of weight $1$, its Jacobian is defined using the dual Hodge structure $\check{H}^1$ by $$ \Jac_\bZ H^1 := \check{H}^1(\bC)/\left(F^0\check{H}^1+\check{H}^1(\bZ)\right) $$ This is a compact complex torus, again since $\check{H}^1_\bC=F^0\check{H}^1 \oplus \conj{F^0\check{H}^1}$ and the $\bZ$-lattice doesn't intersect $F^0$. As in \autoref{eg:splitting_ell_curve}, extensions of $H^1$ by a weight $0$ Hodge structure $W_0$ $$ 0\to W_0 \to E \to H^1\to 0 $$ are classified by elements in $\Hom_\bZ(\check{W}_0(\bZ),\Jac_\bZ H^1)$. Rather, it is dual extensions that are classified by such elements. Let now $K$ be a larger field or ring $\bZ\subseteq K\subseteq \bR$ and $H^1$ be a $K$-Hodge structure of weight $1$. We can also define a Jacobian $$ \Jac_K H^1 := \check{H}^1(\bC)/\left(F^0\check{H}^1+\check{H}^1(K)\right) $$ It is an abelian group (even a $K$-vector space, usually of infinite dimension) with no structure of manifold. Extensions are still classified by elements of $\Hom_K(\check{W}_0(K),\allowbreak \Jac_K H^1)$, where the Jacobian is viewed as a $K$-vector space. \paragraph{Morphisms} If $H^1$ has a $\bZ$-structure and we extend scalars to $K$, then we have a natural map of abelian groups $$ \Jac_\bZ H^1\to \Jac_K H^1 $$ For example $\Jac_\bQ H^1=\Jac_\bZ H^1/\langle\textrm{torsion}\rangle$ and an extension class is torsion in the usual Jacobian if and only if the extension splits over $\bQ$. More generally, suppose we have (after extending scalars to $K$) an inclusion of Hodge structures $H^1_\iota \into H^1$. The dual map becomes $\check{H}^1\onto \check{H}^1_\iota$ and we have an induced surjection on Jacobians $$ \Jac_\bZ H^1 \onto \Jac_K H^1_\iota $$ From an extension class $\xi\in \Hom_\bZ(\check{W}_0(\bZ),\Jac_\bZ H^1)$ we get another one $\xi_\iota\in \Hom_K(\check{W}_0(K),\allowbreak \Jac_K H^1)$ by composing with the above map. This gives a corresponding extension of $K$-Hodge structures. Given a $K$-subspace $S\subseteq \check{W}_0(K)$ we can restrict the extension class $\xi_\iota$ to it and get another such extension. Note that a subspace of $\check{W}_0(K)$ corresponds to a quotient of $W_0(K)$. \section{Splittings over Affine Invariant Manifolds} \label{sec:splittings} In this section we identify variations of mixed Hodge structure that naturally exist on affine invariant manifolds. The main result is that they split after an appropriate extension of scalars, in the sense of the previous section. \subsection{Setup.} Consider an affine invariant manifold $\cM$ inside a stratum of flat surfaces $\cH$. It was proved in \cite[Theorem 7.3]{ssimple} that the variation of Hodge structure over $\cM$ given by the first cohomology $H^1$ splits as $$ H^1=\left(H^1_{\iota_0}\oplus\cdots\oplus H^1_{\iota_{r-1}}\right)\oplus V $$ Each term above gives a variation of Hodge structure. Moreover, the direct sum in the parenthesis comes from a $\bQ$-local system. Each summand $H^1_\iota$ corresponds to an embedding $\iota$ of a totally real number field $k$, and comes from a local system defined over that embedding. The embedding $\iota_0$ is the distinguished embedding. Recall the tangent space $T\cH$ to the stratum is given, via period coordinates, by the relative cohomology of the underlying surfaces. Restrict the tangent bundle $T\cH$ to $\cM$. It projects to $H^1$ and we can take the preimage of the summands coming from the totally real field. This yields an exact sequence of $\bQ$-local systems \begin{align} \label{eqn:big_ses} 0\to W_0 \to E \overset{p}{\rightarrow} \left(H^1_{\iota_0}\oplus\cdots \oplus H^1_{\iota_{r-1}}\right) \to 0 \end{align} Here $W_0$ is the local system of purely relative cohomology classes. It coincides with the (reduced) cohomology of the marked points (i.e. zeroes of the $1$-form). This provides a variation of $\bQ$-mixed Hodge structures in the following sense. Above each point in $\cM$ we have an induced mixed Hodge structure. The Hodge filtrations $F^\bullet$ give holomorphic subbundles of the corresponding local systems. The Griffiths transversality conditions are empty in this case. \begin{remark} Below we discuss local systems defined over a particular field, for example $\iota_0(k)$ from above. To define this notion, fix a normal closure $K$ containing all the embeddings of $k$, with Galois group over $\bQ$ denoted $G$. Given a $\bQ$-local system $V_\bQ$, we can extend scalars to $K$ and denote it $V_K$. A sublocal system $W\subset V_K$ is ``defined over $\iota_l(k)$" if it is invariant by the subgroup $G_{\iota_l}\subset G$ stabilizing $\iota_l(k)$. We will omit the explicit extension of scalars to $K$ below, and just say that $W\subset V$ is a sublocal system defined over $\iota_l(k)$. \end{remark} \paragraph{The tangent space of the affine manifold.} According to results of Wright \cite[Thm. 1.5]{Wright_field} the tangent space $T\cM$ to the affine submanifold gives a local subsystem $T\cM\subset E$, which is defined over $k$ (rather, $\iota_0(k)$). It has the property that $p(T\cM)=H^1_{\iota_0}$. Define the kernel of the map $p$ (see \autoref{eqn:big_ses}) by $$ W_0\cM:=W_0\cap T\cM $$ Define also $T\cM_{\iota_l}$ to be the Galois-conjugate of the local system $T\cM$ corresponding to the embedding $\iota_l$. Analogously define $$ W_0\cM_{\iota_l}:=W_0\cap T\cM_{\iota_l} $$ Note that $T\cM_{\iota_l}$ surjects onto $H^1_{\iota_l}$ with kernel $W_0 \cM_{\iota_l}$. \subsection{The splitting} The space $H^1_{\iota_l}$ contains holomorphic $1$-forms denoted $H^{1,0}_{\iota_l}$. These also give \emph{relative} cohomology classes, i.e. a natural map $H^{1,0}_{\iota_l}\to H^1_{rel}$. The main theorem of this section (below) is the compatibility of this map with $T\cM_{\iota_l}$. Note that in the case when $W_0$ is contained in $T\cM$ (i.e. the setup of \autoref{sec:alg_special_case}) the statement below holds trivially. \begin{theorem} \label{thm:iota_l_splitting} Consider the variation of $\iota_l(k)$-mixed Hodge structures \begin{align} \label{eqn:ses_iota_l} 0\to W_0/W_0\cM_{\iota_l} \into p^{-1}(H^1_{\iota_l})/W_0\cM_{\iota_l} \onto H^1_{\iota_l}\to 0 \end{align} It is an exact sequence of local systems defined over $\iota_l(k)$, and each space carries compatible (mixed) Hodge structures. Then this sequence is (pointwise) split over $\iota_l(k)$ in the sense of \autoref{def:split}. The splitting is provided by the local system $T\cM_{\iota_l}/W_0\cM_{\iota_l}$, i.e. the surjection in the sequence \eqref{eqn:ses_iota_l} can be split by an $\iota_l(k)$-isomorphism \begin{align} \label{eqn:H^1_inverse} H^1_{\iota_l} \toisom T\cM_{\iota_l}/W_0\cM_{\iota_l} \end{align} \end{theorem} Note that there are two ways to lift a cohomology class in $H^{1,0}_{\iota_l}$ to $H^1_{rel}$. One uses the map \eqref{eqn:H^1_inverse}, and the other is by viewing a holomorphic $1$-form as a relative cohomology class. This theorem claims that these two ways in fact agree. \begin{proof} The proof is in three steps. In Step 1 we dualize the sequence \eqref{eqn:ses_iota_l} and use the Galois-conjugate tangent space to produce a flat splitting of the local systems. We also find the $\bR$-splitting coming from the underlying Hodge structures. Their difference is a (holomorphic) section of a bundle with negative curvature. In Step 2, we show the section must have constant Hodge norm. In Step 3 we use this to show that the section must come from a flat one, and thus must in fact be zero. This concludes the proof, since it shows that the $\bR$-splitting was in fact defined over $\iota_l(k)$. The next three sections deal with each step. \end{proof} \subsection{Proof of Step 1} Because in the exact sequence \eqref{eqn:ses_iota_l} we quotient out $W_0\cM\iota_l$ we deduce the bundle $T\cM_{\iota_l}/W_0\cM_{\iota_l}$ maps isomorphically onto $H^1_{\iota_l}$. Dualizing the sequence \eqref{eqn:ses_iota_l} we obtain \begin{align} \label{eqn:ses_iota_l_dual} 0\from \left(W_0/W_0\cM_{\iota_l}\right)^\vee \twoheadleftarrow \left(p^{-1}\left(H^1_{\iota_l}\right)/W_0\cM_{\iota_l}\right)^\vee \hookleftarrow \check{H}^1_{\iota_l}\from 0 \end{align} Denote the annihilator of $T\cM_{\iota_l}/W_0\cM_{\iota_l}$ by $\left(T\cM_{\iota_l}/W_0\cM_{\iota_l}\right)^\perp$. By the remark above, it maps isomorphically onto $\left(W_0/W_0\cM_{\iota_l}\right)^\vee$. The inverse isomorphism defines a canonical flat map of $\iota_l(k)$ local systems, which is a splitting of the left surjection in the exact sequence \eqref{eqn:ses_iota_l_dual} $$ \sigma_{\iota_l}:\left(W_0/W_0\cM_{\iota_l}\right)^\vee \to \left(p^{-1}\left(H^1_{\iota_l}\right)/W_0\cM_{\iota_l}\right)^\vee $$ We now construct another splitting using the Hodge structures (see \autoref{remark:splittings} (iii)). Consider the $F^0$ piece of the middle term in the sequence \eqref{eqn:ses_iota_l_dual}. It maps surjectively onto $\left(W_0/W_0\cM_{\iota_l}\right)^\vee$ with kernel $F^0\check{H}^1_{\iota_l}$. This gives a canonical splitting $$ \sigma_\bR: \left(W_0/W_0\cM_{\iota_l}\right)^\vee_\bC\to \left(p^{-1}\left(H^1_{\iota_l}\right)/W_0\cM_{\iota_l}\right)^\vee_\bC/F^0\check{H}^1_{\iota_l} $$ Note that because it really comes from an isomorphism of vector bundles, it is in fact holomorphic over the affine manifold. Note that both maps $\sigma_\bR$ and $\sigma_{\iota_l}$ are splittings. This means that composing either with the surjection onto the left term of the sequence \eqref{eqn:ses_iota_l_dual} gives the identity. So their difference has image in the kernel of the surjection in \eqref{eqn:ses_iota_l_dual}: $$ \sigma_{\iota_l}-\sigma_\bR: \left(W_0/W_0\cM_{\iota_l}\right)^\vee_\bC\to \check{H}^1_{\iota_l}/F^0\check{H}^1_{\iota_l} $$ Next, we assume that we are in some finite cover of the stratum where the local system $W_0$ is trivial; labeling the zeroes of the $1$-form suffices. Then we can choose an element (same as a global section) of $\left(W_0/W_0 \cM_{\iota_l}\right)^\vee_\bC$ denoted by $e$. By taking its image under the above map, we obtain a global over $\cM$ holomorphic section $$ \psi_e :=(\sigma_{\iota_l}-\sigma_{\bR})(e)\in \Gamma(\cM; \check{H}^1_{\iota_l}/F^0\check{H}^1_{\iota_l}) $$ \subsection{Proof of Step 2} Given the holomorphic section $\psi_e$ produced in Step 1, we now show it has constant Hodge norm. \paragraph{Notation.} Denote the Hodge decomposition of $\check{H}^1$ by $$ \check{H}^{1} = \check{H}^{0,-1}\oplus \check{H}^{-1,0} $$ We keep the same notation for indices involving the embeddings $\iota_l$. Note that $H^{0,-1}=F^0\check{H}^1$ gives a holomorphic subbundle, being identified with the annihilator of $H^{1,0}=F^{1}{H}^1$ (which is the bundle of holomorphic $1$-forms). Using the polarization of $H^1$, we see that $\check{H}^1$ is isomorphic to $H^1$, up to a shift of weight by $(1,1)$. The section $\psi_e$ produced above is a holomorphic section of $\check{H}^{-1,0}$, endowed with the holomorphic structure when viewed as a quotient $\check{H}^{-1,0} = \check{H}^1/F^0\check{H}^1$. This bundle has negative curvature (see \cite[Corollary 3.15]{ssimple}, with a weight shift). We want to apply Lemma 5.2 from \cite{ssimple} to conclude the function $\log \norm{\psi_e}$ is constant. For this, we need to check the boundedness and sublinear growth assumptions. This is done below. Note that this function is subharmonic by the calculation in \cite[Lemma 3.1]{ssimple} and the negative curvature of the bundle. If $\norm{\psi_e}\neq 0$ identically, then it can vanish only on a lower-dimensional analytic subset. Therefore, one can define the functions $f_N:=\max(-N,\log \norm{\psi_e})$ and apply \cite[Lemma 5.2]{ssimple} to them (and let $N\to +\infty$). As the maximum of two subharmonic functions, each $f_N$ is itself subharmonic. \paragraph{Checking assumptions.} First, we examine how $\psi_e$ was defined. We have the exact sequence \begin{align} \label{ses:all} 0\to W_0 \to E \to (H^1_{\iota_0}\oplus\cdots \oplus H^1_{\iota_{r-1}}) \to 0 \end{align} and its dual $$ 0\from W^\vee_0 \from E^\vee \from (\check{H}^1_{\iota_0}\oplus\cdots\oplus \check{H}^1_{\iota_{r-1}} )\from 0 $$ We have the $\bR$-splitting coming from Hodge theory (see \autoref{remark:splittings} (iii)) $$ W_0^\vee(\bR)\to E^\vee_{\bR} $$ This gives a direct sum decomposition \begin{align} \label{eqn:mhs_splitting} E^\vee(\bR) \cong W_0^\vee(\bR)\oplus \check{H}^1(\bR) \end{align} and an induced metric from each factor. At the level of the exact sequence \eqref{ses:all} this is the same as using the harmonic representatives of absolute cohomology classes to obtain relative cohomology classes. \paragraph{Terminology for norms.} Using the splitting from \eqref{eqn:mhs_splitting} and more generally the same for $H^1_{rel}$, we can put norms on the cocycle by putting a norm on each factor. Several norms will appear below, and we explain now the terminology. On $H^1$ and its dual $\check{H}^1$ one has the Hodge norms, coming from the Hodge structure (which are preserved by duality). These Hodge norms will be denoted $\norm{-}$ in the sequel. Eskin, Mirzakhani and Mohammadi \cite[Sec. 7]{EMM} introduced a modified Hodge norm on $H^1$, obtained by increasing the usual Hodge norm. Note that the dual of this modified Hodge norm, on $\check{H}^1$, is \emph{less than} the Hodge norm on $\check{H}^1$. The modified Hodge norms on $H^1$ and $\check{H}^1$ will be denoted $\norm{-}'$. Finally, we have the modified mixed Hodge norms on $H^1_{rel}$ and $\check{H}^1_{rel}$. These will be denoted $\vertiii{-}$ (note that they will not be dual to each other). On $H^1_{rel}$, the modified mixed Hodge norm is defined in \cite{EMM} by putting the constant norm on $W_0$ and the modified Hodge norm on $H^1$. On $\check{H}^1_{rel}$, the modified mixed Hodge norm is defined in \autoref{eqn:dual_modified_Hg} by changing the norm on the $W_0^\vee$ factor. Below, the adjective \emph{dual} means that the norm is on the dual space, and the adjective \emph{mixed} means that it is in relative (co)homology. We now proceed to the details. \paragraph{Modified Hodge norm.} The modified Hodge norm $\norm{-}'$ on $H^1$ is defined in \cite[eqn. (33) and below]{EMM}) and the cocycle on $H^1$ is integrable for it (\cite[Lemma 7.4]{EMM}). Next, recall the splitting $H^1_{rel}=W_0\oplus H^1$ coming from Hodge theory using holomorphic lifts (in their language -- harmonic representatives). Using it, the modified Hodge norm $\vertiii{-}$ on $H^1_{rel}$ is defined using the modified Hodge norm on $H^1$ and the constant norm on $W_0$ (see \cite[eq. (40) and above]{EMM}). For this modified Hodge norm on $H^1_{rel}$ the cocycle is bounded \cite[Lemma 7.5]{EMM}. The main consequence \cite[Lemma 7.6]{EMM} is the following. For the splitting $H^1_{rel}=W_0 \oplus H^1$, write the cocycle matrix as \begin{align} \label{eqn:explicit_KZ} \begin{bmatrix} 1 & U(x,s)\\ 0 & A(x,s) \end{bmatrix} \textrm{ where }U(x,s):H^1_x \to (W_0)_{g_s x} \end{align} Then $\norm{U(x,s)}_{op}\leq e^{m's}$ for some $m'$, where $H^1$ is viewed with the modified Hodge norm $\norm{-}'$ and $W_0$ with the constant norm. Explictly, this says \begin{align} \label{eqn:U_bdd_H1} \norm{U(x,s)v}\leq e^{m's}\norm{v}' \textrm{ where }v\in H^1_{x}\textrm{ and so }U(x,s)v\in (W_0)_{g_sx} \end{align} \paragraph{Dual modified Hodge norm.} For a point $x$ in the stratum, let $r(x)\in [1,\infty)$ be the ratio of the modified to the usual Hodge norms on $H^1$. In other words \begin{align} r(x) := \sup_{0\neq v\in H^1_x} \frac{\norm{v}' }{\norm{v} } \end{align} Note that $r(x)\geq 1$ since $\norm{-}'$ is always at least the usual Hodge norm (see before eq. (40) in \cite{EMM}). By definition $r(x)\cdot \norm{v} \geq \norm{v}'$ for any $v\in H^1_x$. On $\check{H}^1$ there are two norms -- the dual of the usual Hodge norm, denoted $\norm{-}$, and the dual of the modified Hodge norm, denoted $\norm{-}'$. The inequality involving $r(x)$ is now reversed, i.e. letting $\xi\in \check{H}^1_x$, we have \begin{align} \label{eqn:r_dual_bound} \norm{\xi} \leq r(x)\cdot \norm{\xi}' \end{align} To see this, using $\frac 1 {r(x)} \norm{v}'\leq \norm{v}$ for the second step, we have \begin{align} \norm{\xi} = \sup_{\norm{v}=1} |\xi(v)| \leq \sup_{\frac 1 r \norm{v}'=1} |\xi(v)| = r(x)\norm{\xi}' \end{align} \paragraph{Dual modified mixed Hodge norm.} We now explain how to modify the Hodge norm on the dual cocycle $\check{H}^1_{rel}$ which is relevant to our arguments. Recall the splitting $\check{H}^1_{rel} = W_0^\vee \oplus \check{H}^1$ coming from Hodge theory. Given a vector $w\oplus h\in H^1_{rel}$ its ordinary mixed Hodge norm is $\norm{w\oplus h}^2=\norm{w_0}^2+\norm{h}^2$ (using the usual Hodge norm for $h$ and constant norm for $w_0$). Define now its dual modified mixed Hodge norm via \begin{align} \label{eqn:dual_modified_Hg} \vertiii{w\oplus h}^2 := \norm{w_0}^2\cdot r(x)^2 + \norm{h}^2 \end{align} Note that the modification affects only the $W_0^\vee$ part of the norm, not the $\check{H}^1$. \begin{proposition} For the modified dual Hodge norm defined in \autoref{eqn:dual_modified_Hg}, the Kontsevich-Zorich cocycle is bounded. \end{proposition} \begin{proof} Given the explicit form of the KZ cocycle in \autoref{eqn:explicit_KZ}, its dual cocycle (for the dual splitting) will be given by the inverse transpose of that matrix. This reads \begin{align} \left( \begin{bmatrix} 1 & U(x,s)\\ 0 & A(x,s) \end{bmatrix}^t\right)^{-1} = \begin{bmatrix} 1 & 0\\ -(A(x,s)^t)^{-1} \cdot U(x,s)^t & (A(x,s)^t)^{-1} \end{bmatrix} \end{align} Next, recall that the cocycle $A(x,s)$ is bounded in both forward and backwards time, since it corresponds to $H^1$, and we have the usual Hodge norm on that part. To show boundedness of the full cocycle, it suffices to prove that $\vertiii{U(x,s)^t}_{op}\leq e^{m's}$ for some $m'$. Note that $U(x,s)^t : (W_0^\vee)_{g_sx}\to \check{H}^1_x$, since $U(x,s)$ goes the other way between dual spaces (see \eqref{eqn:explicit_KZ}). The bound on the operator norm of $U(x,s)^t$ needs to hold when $W_0^\vee$ is viewed with the norm $r(x)\norm{-}$, and $\check{H}^1$ is with the usual Hodge norm. The operator $U(x,s)$ is bounded for the modified Hodge norm on $H^1$ and usual norm on $W_0$ (see \eqref{eqn:U_bdd_H1}). Therefore $U(x,s)^t$ satisfies the same bound for the dual modified Hodge norm on $\check{H}^1$ and usual norm on $W_0^\vee$. This reads \begin{align} \norm{U(x,s)^t \check{w}}' \leq e^{m's} \norm{\check{w}} \textrm{ where }\check{w}\in (W_0^\vee)_x \end{align} However, using \autoref{eqn:r_dual_bound} which relates the usual and modified Hodge norms in the dual $\check{H}^1$ we find \begin{align} \norm{U(x,s)^t \check{w}}\cdot \frac 1 {r(x)} \leq \norm{U(x,s)^t \check{w}}' \leq e^{m's} \norm{\check{w}} \end{align} Moving $r(x)$ to the right side, this exactly says that $U(x,s)^t$ is bounded when $W_0^\vee$ carries the dual modified mixed Hodge norm. Finally, note that the identity map on $W_0^\vee$ no longer acts by isometries, since the norm on that factor depends now on the basepoint. However, from the boundedness of the cocycle on $H^1$ with the usual, as well as the modified Hodge norm, it follows that $r(g_s x)\leq e^{m''s}r(x)$ for all $s$. Thus, the Kontsevich-Zorich cocycle on $\check{H}^1_{rel}$ is bounded for the dual modified mixed Hodge norm. \end{proof} \paragraph{Properties of the dual modified mixed Hodge norm.} First, note that the boundedness properties of the cocycle descend to the various pieces of $\check{H}^1$ and $\check{H}^1_{rel}$. Returning to the notation in \autoref{ses:all}, \eqref{eqn:mhs_splitting}, we constructed a norm $\vertiii{-}$ satisfying \begin{enumerate} \item [(i)] The Kontsevich-Zorich cocycle on $E$ and $E^\vee$ is integrable for this norm. Moreover, it satisfies the absolute bound for some universal constant $C>0$ \begin{align} \label{eqn:g_t_bdd_mod_norm} -C\cdot T \leq \log \vertiii{g_T} \leq C\cdot T \end{align} \item [(ii)] The projection $E\to H^1$ is norm non-increasing and dually the embedding $\check{H}^1\into E^\vee$ is norm non-decreasing (where $H^1$ and $\check{H}^1$ have the usual Hodge norms). In other words, if $\psi\in \check{H}^1$ is a section, then \begin{align} \label{eqn:norm_non_dec} \norm{\psi}\leq \vertiii{\psi} \textrm{\hskip 0.5em (in fact, we have equality in this case)} \end{align} Moreover, if $\phi$ if a section of $H^1_{rel}$ and $\psi$ is its $\check{H}^1$-component, then $\vertiii{\phi}\geq \norm{\psi}$. In other words, the dual modified mixed Hodge norm of $\phi$ dominates the usual Hodge norm of $\psi$, since the decomposition $\check{H}^1_{rel}=W_0^\vee \oplus\check{H}^1$ is orthogonal for the dual modified mixed Hodge norm. \end{enumerate} Consider now $$ 0\to W_0 \to p^{-1}(H^1_{\iota_l})\to H^1_{\iota_l} \to 0 $$ In the dual picture, with the dual modified mixed Hodge norm, we have \begin{align} \label{eqn:ses_iota_l_dual_again} 0\from W_0^\vee \twoheadleftarrow p^{-1}(H^1_{\iota_l})^\vee \hookleftarrow \check{H}^1_{\iota_l} \from 0 \end{align} Corresponding to $W_0\cM_{\iota_l}$ in the first sequence is its annihilator $W_0\cM_{\iota_l}^\perp$ in the dual. It is naturally identified with $\left(W_0/W_0\cM_{\iota_l}\right)^\vee$. Now $p^{-1}(H^1_{\iota_l})^\vee$ also contains the annihilator of $T\cM_{\iota_l}$ denoted $T\cM_{\iota_l}^\perp$. The surjection in the dual sequence \eqref{eqn:ses_iota_l_dual_again} gives an isomorphism $$ W_0\cM_{\iota_l}^\perp \from T\cM_{\iota_l}^\perp $$ The section $\sigma_{\iota_l}$ from Step 1 is the inverse of this isomorphism. Given $e\in(W_0/W_0\cM_{\iota_l})^\vee$ (which is naturally $W_0\cM_{\iota_l}^\perp$) we have $\phi_e$ defined as $$ \phi_e:=\sigma_{\iota_l}(e) \in T\cM^\perp_{\iota_l}\subset p^{-1}(H^1_{\iota_l})^\vee $$ Note that $\phi_e$ is a flat global section of $p^{-1}(H^1_{\iota_l})^\vee$. This last bundle, equipped with the dual modified mixed Hodge norm, gives rise to an integrable cocycle; the Oseledets theorem thus holds. Just like in \cite[Lemma 5.1]{ssimple} $\phi_e$ must be in the zero Lyapunov subspace, otherwise its norm would be unbounded on any set of positive measure. In particular, its dual modified mixed Hodge norm grows subexponentially along a.e. \Teichmuller geodesic. Recall the splitting over $\bR$ for the sequence \eqref{eqn:ses_iota_l_dual_again}. This comes from the decomposition \begin{align} \label{eqn:dir_sum_R} p^{-1}(H^1_{\iota_l})^\vee (\bR)\cong \check{H}^1_{\iota_l} (\bR)\oplus \left(F^0\cap \conj{F^0}\right) \end{align} Here $F^0$ refers to the corresponding piece of the filtration of $p^{-1}(H^1_{\iota_l})$ and the construction was explained in \autoref{remark:splittings} (iii). \autoref{eqn:dir_sum_R} is a direct sum decomposition of $\bR$-vector bundles and $\sigma_\bR$ comes from inverting the $\bR$-isomorphism $$ W_0^\vee(\bR)\from F^0\cap \conj{F^0} $$ This means that $$\psi_e = (\sigma_{\iota_l} - \sigma_\bR)(e)\in\check{H}^1_{\iota_l}(\bR)$$ is just the $\check{H}^1_{\iota_l}(\bR)$-component of $\phi_e$ in the decomposition \eqref{eqn:dir_sum_R}. \begin{remark} For purposes of comparing metrics, we are using the isomorphism of $\bR$-vector bundles $$ \check{H}^1_{\iota_l}(\bR)\toisom \check{H}^1_{\iota_l}(\bC)/F^0\check{H}^1_{\iota_l} $$ The section we are considering can be viewed as living in either. In the second one, it is also holomorphic for the natural holomorphic structure. \end{remark} From the property of the dual modified mixed Hodge norm in \eqref{eqn:norm_non_dec}, $\vertiii{\phi_e}$ bounds from above the usual Hodge norm of its $\check{H}^1_{\iota_l}$-component (which is $\psi_e$). This gives the desired subexponential growth for the Hodge norm $\norm{\psi_e}$. We must also verify that upon acting by an element $g\in \SL_2\bR$, the function has not increased by more than $C\norm{g}$, for some absolute constant $C$. This follows from \autoref{eqn:g_t_bdd_mod_norm} and the fact that the norms are $\SO_2(\bR)$-invariant \cite[eq. (41)]{EMM}. To conclude, the conditions of \cite[Lemma 5.2]{ssimple} are satisfied; therefore the Hodge norm $\norm{\psi_e}$ is constant. \subsection{Proof of Step 3} Step 2 showed that $\norm{\psi_e}$ is constant. To conclude, we now show it must be zero. Because $\norm{\psi_e}$ is constant, using \cite[Remark 3.3]{ssimple} for the formula for $\delbar\del$ it follows that $$ 0=\delbar \del \norm{\psi_e}^2 = \ip{\Omega \psi_e}{\psi_e} - \ip{\nabla^{Hg}\psi_e}{\nabla^{Hg}\psi} $$ where $\Omega$ is the curvature of $\check{H}^{-1,0}_{\iota_l}$, which is negative-definite. We conclude \begin{align*} \sigma^\dag \psi_e = 0 \hskip 1cm \textrm{and} \hskip 1cm \nabla^{Hg}\psi_e=0 \end{align*} Recall that $\sigma:\Omega^{1,0}(\cM)\otimes \check{H}^{0,-1}_{\iota_l}\to\check{H}^{-1,0}_{\iota_l}$ is the second fundamental form of $\check{H}^1_{\iota_l}$ and the above identities hold in all direction on $\cM$, not just the $\SL_2\bR$ direction. The curvature satisfies $\Omega=\sigma\sigma^\dag$. Define now a section of $\check{H}^1_{\iota_l}$ by $\alpha := \conj{\psi_e}\oplus \psi_e$. Then $\alpha$ is flat for the Gauss-Manin connection (since $\nabla^{GM}=\nabla^{Hg}+\sigma+\sigma^\dag$). But the local system $H^1_{\iota_l}$ is irreducible (see \cite[Thm. 1.5]{Wright_field}) so it has no flat global sections over the affine manifold $\cM$. Therefore $\alpha=0$, so $\psi_e=0$ and thus $\sigma_{\iota_l}=\sigma_\bR$. This finishes the proof of \autoref{thm:iota_l_splitting}. \hfill \qed \section{Algebraicity and torsion} \label{sec:algebraicity_torsion} In this section, we prove the theorems stated in the introduction. First we combine the results of the previous section to find that a certain twisted version of the Abel-Jacobi map is torsion. Using this result, we then prove that affine invariant manifolds are algebraic. \subsection{Combining the splittings} \label{subsec:combinig_splittings} \textbf{Setup.} In the setting of the previous section, over an affine manifold $\cM$ we had an exact sequence of $\bZ$-mixed Hodge structures $$ 0\to W_0 \to E \overset{p}{\to} \bigoplus_\iota H^1_\iota \to 0 $$ We are assuming some fixed $\bZ$-structure on $\oplus_\iota H^1_\iota$. As was explained in \autoref{subsec:exts_fields} this corresponds to a map \begin{align} \label{eqn:def_xi} \xi:\check{W}_0(\bZ) \to \Jac_\bZ\left(\bigoplus_\iota H^1_\iota\right) \end{align} Thus for each element of $\check{W}_0(\bZ)$ we get a section of the bundle of Jacobians. We are working with a factor of the actual Jacobian, but omit this from the wording. By \autoref{thm:iota_l_splitting} for each $\iota$ the variation of $\iota(k)$-mixed Hodge structures $$ 0\to W_0/W_0 \cM_{\iota} \to p^{-1}(H^1_{\iota})/W_0 \cM_{\iota} \to H^1_{\iota}\to 0 $$ is split. In the language of \autoref{subsec:exts_fields} this means that pointwise on $\cM$ the induced map of abelian groups (and $\iota(k)$-vector spaces) $$ \xi_{\iota}:(W_0 \cM_{\iota})^\perp (\iota(k)) \to \Jac_{\iota(k)}H^1_{\iota} $$ is in fact the zero map. Recall $\xi_\iota$ is obtained from $\xi$ by composing with the quotient map $$ \Jac_\bZ\left(\bigoplus_\iota H^1_\iota\right) \onto \Jac_{\iota(k)} H^1_\iota $$ and restricting the domain to $(W_0\cM_\iota)^\perp$ (after extending scalars). Another description of $\xi_\iota$ is as follows. Given $c^j\in \iota(k)$ and $r_j\in \check{W}_0(\bZ)$ with $\sum_j c^j r_j\in (W_0\cM_\iota)^\perp$ we have $$ \xi_\iota\left(\sum_j c^j r_j\right) = \left.\sum_j c^j \xi(r_j)\right|_{H^1_\iota} $$ The next step will be to combine all the above statements, as $\iota$ ranges over all embeddings of $k$ into $\bR$. \paragraph{{The twisted cycle map.}} Since $\oplus_\iota H^1_\iota$ has real multiplication by $k$, we have an order $\cO\subseteq k$ which acts by endomorphisms on the bundle of abelian varieties $\Jac_\bZ(\oplus_\iota H^1_\iota)$. Inside $(W_0\cM_{\iota_0})^\perp$ which is a local system over $\iota_0(k)$, we can choose an $\cO$-lattice, i.e. the $\iota_0(\cO)$-submodule \begin{align} \label{eqn:Lambda_iota_def} \Lambda_{\iota_0}:= \left(\check{W}_0(\bZ)\otimes_{\bZ}\iota_0(\cO) \right)\cap (W_0 \cM_{\iota_0})^\perp \end{align} Note that $(W_0\cM_{\iota_0})^\perp\subset \check{W}_0$, since this is the annihilator of $W_0\cM\subset W_0$. After extending scalars to $\bQ$ we have an isomorphism $$ \Lambda_{\iota_0}\otimes_\bZ \bQ \toisom (W_0 \cM_{\iota_0})^\perp $$ \begin{definition} \label{def:twisted_cycle} Recall that for $a\in k$ we denote by $\rho(a)$ the corresponding endomorphism of the family of Jacobians and the map $\xi$ was defined in \autoref{eqn:def_xi}. For $c^j\in \cO$ and $r_j\in \check{W}_0(\bZ)$ we can then define a twisted cycle map \begin{align*} \nu:\Lambda_{\iota_0}&\to \Jac_\bZ\left(\bigoplus_\iota H^1_\iota\right)\\ \sum_j c^j r_j &\mapsto \sum_j \rho(c^j) \xi(r_j) \end{align*} \end{definition} The following is the main theorem of this section. It is a generalization of Theorem 3.3 of M\"{o}ller from \cite{Moller_torsion}. \begin{theorem} \label{thm:torsion} The image of $\nu$ is torsion in the Jacobian. \end{theorem} \begin{proof} The proof is in two steps. First, we show that the image of $\nu$ is zero after we apply the quotient map (where $K$ is a normal closure of $k$) $$ \Jac_\bZ\left(\bigoplus_\iota H^1_\iota\right) \onto \Jac_K\left(\bigoplus_\iota H^1_\iota\right) $$ To finish, we prove that it must have been torsion to begin with. We have the following spaces and the relation between them which will appear in the proof. \begin{eqnarray} \begin{tikzcd}[baseline=(current bounding box.center), column sep=tiny] & \check{W}_0 \arrow{rrr}{\xi} & & & \Jac(\oplus_\iota H^1_\iota) \arrow{dl} \arrow{d} \arrow{dr} & \\ (W_0 \cM_{\iota_0})^\perp \arrow[hook]{ur} & \cdots \arrow{u} & (W_0\cM_{\iota})^\perp \arrow[hook]{ul} &\Jac_{\iota_0(k)}(H^1_{\iota_0}) &\ldots & \Jac_{\iota(k)}(H^1_\iota) \label{eqn:diagram} \end{tikzcd} \end{eqnarray} To combine the splittings given by \autoref{thm:iota_l_splitting} we need to extend scalars to the field $K$ which contains all the $\iota(k)$. Consider the classifying map obtained from $\xi$ $$ \xi_K: \check{W}_0(K)\to \Jac_K\left(\bigoplus_\iota H^1_\iota\right) $$ Given $x=\sum_j c^j r_j \in \Lambda_{\iota_0}$ the splittings from \autoref{thm:iota_l_splitting} say that for any $g\in \Gal(K/\bQ)$ we have $$ \left.\xi_K(gx)\right|_{H^1_{g\iota_0}}=0 $$ Recall that to get the splitting we had to apply the Galois action both on $H^1$ and on $W_0$. Assuming $g\iota_0=\iota_l$ an explicit way to write the above vanishing is $$ \left.\xi_K\left(\sum_j \iota_l(c^j)r_j\right)\right|_{H^1_{\iota_l}}=0 $$ Using the $K$-linearity of $\xi_K$ we find $$ \left.\sum_j \iota_l(c^j) \xi_K(r_j)\right|_{H^1_{\iota_l}}=0 $$ Taking a direct sum over the different embeddings $\iota_l$ gives that $\nu(x)=0$ in $\Jac_K(\oplus_\iota H^1_\iota)$. To see this, recall that the real multiplication action of $k$ on the factor $H^1_\iota$ was by the embedding $\iota$. To finish, we want to conclude $\nu(x)$ is torsion in the $\bZ$-Jacobian. We know that $\nu(x)$ defines a holomorphic section of the corresponding family of abelian varieties. We just proved its image lies in the (group-theoretic) kernel of the map \begin{align} \label{eqn:ker_Jac_map} \Jac_\bZ\left(\bigoplus_\iota H^1_\iota\right)\onto \Jac_K \left(\bigoplus_\iota H^1_\iota\right) \end{align} Viewing the bundle of Jacobians as a flat family of real tori (flat in the sense of the Gauss-Manin connection) we conclude that $\nu(x)$ is itself flat. For this, the continuity of $\nu$ would be sufficient. However, the monodromy acts by parallel transport on a fixed fiber of this family of tori and $\nu(x)$ is invariant by this action. Since the underlying local system over $\bQ$ is irreducible (the monodromy acts irreducibly) we conclude that $\nu(x)$ must be among the rational points. This implies that $\nu(x)$ is a torsion section of the bundle of Jacobians. \end{proof} \begin{remark} \label{rmk:splittings} The splitting over $\iota(k)$ of the mixed Hodge structure obtained in \autoref{thm:iota_l_splitting} is equivalent to the composition (see diagram \eqref{eqn:diagram}) \begin{align} \label{eqn:long_composition} (W_0\cM_\iota)^\perp \to \check{W_0} \xrightarrow{\xi} \Jac(\oplus_\iota H^1_\iota) \to \Jac_{\iota(k)}(H^1_\iota) \end{align} being the zero map. The torsion condition obtained in \autoref{thm:torsion} implies that the twisted cycle map $\nu:\Lambda_{\iota_0}\to \Jac_\bZ(\oplus_\iota H^1_\iota)$ is torsion. In particular, $A\cdot \nu$ is the zero map for some integer $A\neq 0$. Note that the twisted cycle map $\nu$ involves the maps in \eqref{eqn:long_composition} ranging over all embeddings $\iota$. \noindent The condition that $A\cdot \nu$ is the zero map implies, in particular, that the composition defined in \eqref{eqn:long_composition} is also the zero map, as $\iota$ ranges over all embeddings. Indeed, these maps are the components of $\nu$ after an extension of scalars to the appropriate number field. \noindent Thus, the torsion condition in \autoref{thm:torsion} implies the splitting property in \autoref{thm:iota_l_splitting}. \end{remark} \subsection{Algebraicity in the general case} \paragraph{Setup.} We keep the notation from the previous section. $\cM$ denotes an affine invariant manifold, $T\cM_{\iota_0}$ its tangent bundle and $W_0\cM_{\iota_0}$ its intersection with the purely relative subbundle. We have the decomposition of the Hodge bundle $$ H^1=\left(\bigoplus_\iota H^1_\iota\right)\oplus V $$ Recall also that $\cM$ lives in a stratum $\cH$ and we have the tautological section $$ \omega:\cH\to T\cH $$ This section is algebraic and even defined over $\bQ$ (see \autoref{rmk:field_defn}). Combined with the Gauss-Manin connection it gives the local flat coordinates on the stratum $\cH$. \begin{theorem} \label{thm:algebraic} The affine invariant manifold $\cM$ is a quasi-projective algebraic subvariety of $\cH$, defined over $\conj{\bQ}$. \end{theorem} \begin{proof} The proof is similar to that of \autoref{prop:alg_special_case}. In particular, as per \autoref{rmk:self_inters} it is carried out in the open subset of $\cM$ where there are no self-intersections. We first define an algebraic variety $\cN'''$ which has the same properties as $\cM$ and contains it. Then we check that one of its irreducible components is contained in $\cM$ and thus has to coincide with it. First, let $\cN'\subseteq \cH$ be the locus where $(X,\omega)$ admits real multiplication of the same type as on $\cM$, with $\omega$ as an eigenform. On $\cN'$ we also have the map $\nu$ described in \autoref{def:twisted_cycle} $$ \nu:\Lambda_{\iota_0} \to \Jac_\bZ\left(\bigoplus_\iota H^1_{\iota}\right) $$ By \autoref{thm:torsion}, over $\cM$ the map lands in the torsion so there exists $A\in \bN$ such that $A\cdot\nu\equiv 0$ everywhere on $\cM$. Let $\cN''\subseteq \cN'$ be the sublocus where the equation $A\nu=0$ holds on $\cN'$. This is again algebraic, defined over $\conj{\bQ}$, and contains $\cM$. Recall now that we are working in a finite cover where the local system $W_0$ is trivial and so $W_0\cM_{\iota_0}$ is globally defined. Over $\cN''$ the condition $A\cdot \nu=0$ implies (see \autoref{rmk:splittings}) that the exact sequence of mixed Hodge structures splits (over $\iota_0(k)$) $$ 0\to W_0/W_0\cM_{\iota_0} \to p^{-1}(H^1_{\iota_0})/W_0\cM_{\iota_0}\to H^1_{\iota_0} \to 0 $$ The splitting subbundle in $p^{-1}(H^1_{\iota_0})/W_0 \cM_{\iota_0}$ is unique and itself algebraic. Denote this bundle by $T'$. It also gives a local system and it is the candidate for the tangent space. Recall we had the tautological section $\omega:\cH\to T\cH$ and let $\cN'''\subseteq\cN''$ be the locus where $\omega\in T'$. This is an algebraic variety over $\conj{\bQ}$ by construction, and we claim $\cM$ coincides with the irreducible component of $\cN'''$ containing it. To see this, first note that $T'=T\cM$ on $\cM$. However, locally near a point of $\cM$ the condition $\omega\in T'$ is the same as being on $\cM$. Indeed, requiring $\omega$ to lie in some local subsystem is the same as requiring the flat surface to be in a linear subspace in local period coordinates (see \autoref{rmk:loc_syst_per_coord}). Finally, note that $T'$ is a subquotient of $T\cH$, but the condition $\omega\in T'$ still makes sense. It is understood as $\omega$ belonging to the preimage of $T'$ in $T\cH$. This completes the proof of algebraicity. \end{proof} \begin{remark} \label{rmk:affine_mnfds_dimension} Let us explain how affine manifolds are distinguished among loci of real multiplication, since usually these are not $\GL_2^+\bR$-invariant. Suppose given a subvariety $\cN'\subset \cH$ parametrizing $(X,\omega)$ admitting real multiplication by $k$ with $\omega$ an eigenform. Given a $k$-local system $W_0T\subset W_0$ (thinking of it as $W_0\cM_{\iota_0}$), further require that the quotient mixed Hodge structure splits as in the above theorem. Let the splitting be given by a bundle (and also local system) denoted $T$. Finally, require that $\omega$ lies in $T$. This provides us with a locus $\cN\subseteq \cH$. Note that the requirement $\omega \in T$ implies that the Zariski tangent bundle of $\cN$, denoted $T\cN$, is contained in $T$. The variety $\cN$ will also be $\SL_2\bR$-invariant if and only if we have the equality $T\cN=T$. \end{remark} \begin{remark}[\textbf{On fields of algebraic definition}] \label{rmk:field_defn}\leavevmode \begin{enumerate} \item[(i)] For the purposes of this discussion, a variety is ``defined over a field $K$'' if it can be described in a projective space as the zero locus of polynomials with coefficients in $K$. It is quasi-projective if it can be described in a projective space as the locus where a given collection of polynomials vanish, and another collection doesn't vanish. A map between varieties (in particular a section of a bundle) is ``defined over $K$'' if it can be described using polynomials with coefficients in $K$. \item[(ii)] A stratum $\cH$ is a (quasi-projective) algebraic variety defined over $\bQ$. To see this, recall that the moduli space of genus $g$ Riemann surfaces $\cM_g$ is defined over $\bQ$, and so is the Hodge bundle $\cH_g\xrightarrow{\pi}\cM_g$. Moreover, the pullback of the Hodge bundle $\pi^*\cH_g$ to $\cH_g$ has the tautological section $\omega:\cH_g\to \pi^*\cH_g$, also defined over $\bQ$. Next, letting $\cC_g\to \cM_g$ be the universal bundle of Riemann surfaces, the Hodge bundle admits a map $Div:\cH_g\to \Sym^{2g-2}_{\cM_g}\cC_g$ which takes a $1$-form to its zero locus. The space $\Sym^{2g-2}_{\cM_g}\cC_g$ parametrizes $2g-2$-tuples of (not necessarily distinct) points on the same fiber of the universal family. This space is defined over $\bQ$ and has a stratification, also defined over $\bQ$, depending on multiplicities. A stratum of $\cH_g$ is then the preimage of one of the strata on $\Sym^{2g-2}_{\cM_g}\cC_g$. The variety thus obtained need not be connected. To distinguish connected components, one might apriori need to extend the base field $\bQ$. However, from the classification of connected components due to Kontsevich \& Zorich \cite{KZ_components}, each connected component can be described by an algebraic condition also defined over $\bQ$. Indeed, the spin invariant of a square root of the canonical bundle is invariant by Galois conjugation, and so is the property of being hyperelliptic. \item[(iii)] The tautological section $\omega$ of the Hodge bundle restricted to a stratum is then also defined over $\bQ$. Moreover, the stratum carries a universal family of Riemann surfaces, and also the canonical set of marked points corresponding to the zeros of $\omega$. The vector bundle $H^1_{rel}$ has a description as the algebraic de Rham cohomology of this family of Riemann surfaces, and is thus itself defined over $\bQ$. The natural map from the Hodge bundle $\pi^*\cH_g$ to $H^1_{rel}$ is also defined over $\bQ$, and so $\omega:\cH\to H^1_{rel}\cong T\cH$ is defined over $\bQ$. See also \cite[Section 2]{Moller_Linear_mnfds} for a detailed discussion of these constructions. \item[(iv)] The proof of \autoref{thm:algebraic} involved the locus $\cN'$ where a factor of the Jacobian had real multiplication by $\cO$. In $\cA_g$, this locus is defined over $\bQ$, but to select the components which have the $\cO$-module structure on $H^1$ as on $\cM$, we need to pass to some finite extension of $\bQ$. Hence, the locus $\cN'$ is defined over $\overline{\bQ}$. Next, to define the locus $\cN''$ by imposing the torsion condition $A\cdot \nu \equiv 0$ required a finite cover where the zeros are labeled. This requires another finite extension of the base field. \item[(v)] An intersection of two varieties (e.g. imposing the condition that $\omega$ lies in a subbundle) is defined over a field which contains the defining fields for both varieties. \end{enumerate} \end{remark} The above discussion explains why affine invariant manifolds are quasi-projective varieties defined over $\conj{\bQ}$. Acting by the Galois group of $\bQ$ on the defining equations inside $\cH$ produces new quasi-projective varieties. These will also be affine invariant manifolds. \begin{corollary} The group $\Gal(\bQ)$ acts on the set of affine invariant manifolds. \end{corollary} \begin{proof} From the proof of \autoref{thm:algebraic} (see also the discussion in \autoref{rmk:affine_mnfds_dimension}) an affine manifold $\cM$ is defined by the following set of algebraic conditions. \begin{itemize} \item The parametrized Riemann surfaces have a factor with real multiplication by $\cO$, and $\omega$ as an eigenform. \item There is a sublocal system $\Lambda_{\iota_0}\subset \check{W}_0$ such that its image under the twisted Abel-Jacobi map is torsion of some fixed degree, i.e. $A\cdot \nu\equiv 0$ for some integer $A\neq 0$. \item There is an upper bound for the dimension of this locus inside the stratum $\cH$ (computed from the ranks of the objects above) and the dimension of $\cM$ achieves this bound. \end{itemize} Each of the above conditions persists when the Galois group of $\bQ$ acts on the defining equations. Note that $\omega$ is defined with $\bQ$-coefficients, and so if it is an eigenform on one locus, it will also be an eigenform in the Galois-conjugate locus. \end{proof} \begin{remark} The results of Wright \cite{Wright_field} show that in local period coordinates on $H^1_{rel}$ an affine manifold is described by linear equations with coefficients in the number field $k$. These equations and field of definition are usually not related to the algebraic equations and respective fields. In particular, acting by the Galois group on the linear equations would typically not produce another affine manifold. For comparison, the field of affine definition of square-tiled surfaces is $\bQ$. However, algebraically these can be quite rich -- M\"{o}ller \cite[Thm. 5.4]{Moller_GT} showed the action of the Galois group is faithful on the corresponding \Teichmuller curves. \end{remark} \bibliographystyle{sfilip}
\section{Introduction} \label{sect:introduction} The metallicity of stars is the key to exploring of the chemical structure and evolution of the Galaxy. But which is the best way to determine this important astrophysical parameter? High-resolution spectroscopy is often claimed to be the most accurate method \mbox{\citep[e.g.][]{2009A&A...494...95M}}. It is certainly true that this is the only method available to measure abundances of individual chemical elements. However, the overall stellar metallicity can be reliably inferred by a variety of other means provided that a good calibration is available. Furthermore, the accuracy of high-resolution spectroscopy is not always assessed in a satisfactory way. Photometry and low-resolution spectroscopic studies have the advantage of providing results for large samples within a short time (both the observations and the analysis require less time). High-resolution spectra contain significantly more information and are therefore expected to give more accurate results. However, interpreting these data is based on theoretical stellar atmospheres and modelling of spectral lines. The complete procedure of an abundance analysis requires one to specify a considerable amount of input data, assumptions for the physics of relevant processes, and a significant number of free parameters. This leads to a wide variety of possibilities to analyse a given data set, which may lead to differences in the metallicity scales published by different research groups. We set out to investigate to which extent this is the case, thus testing the \emph{accuracy} of spectroscopic stellar metallicities. We note that individual metallicity studies might well be able to achieve a high \emph{precision} for a limited number of stars \mbox{\citep[e.g.][]{2009ApJ...704L..66M}}, allowing one to draw important conclusions on a small subsystem of Galactic stars. But to obtain a complete picture of the Galactic chemical evolution requires one to combine the results of different authors, which will each be subject to systematic uncertainties of unknown magnitude. For recent studies that compared spectroscopic analyses of well-known field stars using several different methods see for instance \mbox{\citet{lebzelter2012}} and Jofr\'e et al. (A\&A, submitted). We focus here on metallicity determinations of stars in open clusters (OCs). These objects have been used for a long time to investigate the radial metallicity gradient of the Galactic disk. Furthermore, the studied stars cover a wide range in stellar parameters ($T_{\rm eff}$\ and $\log g$). Stars in any one cluster all formed at the same time from the same material with a unique metallicity, which provides an additional test for the consistency of the determined metallicities. Since the 1990s, we have seen a surge of spectroscopic metallicity studies of OC stars. This probably reflects the progress in automated abundance analysis, but is certainly also an effect of the availability of efficient multi-object spectrographs, such as FLAMES at ESO's VLT or Hydra at the WIYN telescope on Kitt Peak. Open clusters are regarded as ideal probes for the chemical evolution of the Galactic disk, because they are numerous (more than 2000 presently known), are located throughout the disk, and span a wide range of ages \mbox{\citep{2002A&A...389..871D}}. An accurate knowledge of the cluster metallicity is of twofold importance. It obviously provides a measurement point for the disk chemistry at a certain location. Second, it is important for determining the cluster age and distance (using metallicity-dependent isochrones) and thus the point in time corresponding to the particular metallicity measurement. When interpreting these measurements, it is important, however, to consider possible motions of the clusters during their lifetime \citep[e.g.][]{2009MNRAS.399.2146W}. In spite of the astrophysical significance, metallicity estimations for OCs are still rare. In a previous paper, we compiled photometric metallicity determinations for 188 OCs \mbox{\citep[][hereafter Paper~I]{2010AaA...517A..32P}}. Recent compilations of spectroscopic metallicities include those of \mbox{\citet[][63 clusters]{2009A&A...494...95M,2010AaA...523A..11M}} and \mbox{\citet[][89 clusters]{2011AaA...535A..30C}}. These contain 29 clusters that are not included in our photometric sample, and thus the total current fraction of known clusters with a metallicity assessment is about 10\%. The two spectroscopic compilations take different approaches to arrive at a metallicity value for each cluster. Magrini et al. selected the result of one specific publication per cluster for their sample, while \mbox{\citet{2011AaA...535A..30C}} averaged all available determinations for each cluster. Both approaches suffer from the inhomogeneity inherent in data originating from many different research groups. There are several recent attempts of constructing homogeneous sets of spectroscopic OC metallicities. \mbox{\citet{2010AJ....139.1942F}} complemented their own sample of eleven clusters with results from only three other groups and presented careful evaluations of possible differences in measurements and methods among all groups. This approach resulted in a homogeneous but small sample of 26 clusters. For the BOCCE project\footnote{\url{http://www.bo.astro.it/~angela/bocce.html}} \mbox{\citep{2006AJ....131.1544B,2009IAUS..254..227B}} 45 clusters have been selected, for which age, distance, reddening, and metallicity are being determined in a homogeneous way. For most of the clusters they obtained their own photometry, and for a large portion high-resolution spectra were acquired using only a few instruments. Metallicities determined in a consistent way are published for eight clusters so far. The most ambitious current effort is the \emph{Gaia-ESO} Public Spectroscopic Survey \citep[][PIs S. Randich and G. Gilmore]{2012Msngr.147...25G}, which will obtain medium- and high-resolution spectra for about $10^5$ Galactic stars during five years with one instrument (FLAMES-GIRAFFE-UVES). The target list includes stars in about 100 OCs, and major efforts are put into the preparation of a homogeneous analysis of this unprecedented dataset in a unique collaboration across more than ten research groups. In the current paper, we aim to harvest the existing literature in the best possible way. The motivation is to obtain an up-to-date overview of the status and current limitations of OC metallicities. This is of crucial importance for the implementation of the on-going surveys (in particular the \emph{Gaia} space mission and the \emph{Gaia-ESO} survey) -- concerning both the best selection of target clusters and target stars, and the selection of clusters and stars for calibration purposes. Our approach is to compile atmospheric parameters and spectroscopic metallicities for individual stars in each cluster and compare the results of different authors for stars in common. The article is arranged as follows: in Sect.~\ref{sect:data} we describe the selection of the metallicity determinations. In Sect.~\ref{sect:assessment} we compare the results obtained by different authors who studied the same OCs, star-by-star as well as for the mean cluster metallicity. We also assess the importance of spectrum quality for mean cluster metallicity. In Sect.~\ref{sect:final} we present our final high-resolution sample. In Sect.~\ref{sect:discussion} we compare our sample with others and discuss possible applications of our sample to the study of Galactic structure, and Sect.~\ref{sect:conclusions} concludes the paper. \section{Data selection} \label{sect:data} \subsection{High-resolution sample} \label{sect:hires} To build a list of reference OCs with the most reliable metallicities, we first gathered individual stars -- highly probable members of OCs -- with atmospheric parameters determined from spectra of high resolution ($R=\lambda/\Delta\lambda$) and high signal-to-noise ratio (S/N). The lower limit in spectrum quality was set to $R$=25000, and S/N=50. We searched the PASTEL database \mbox{\citep{2010A&A...515A.111S}} and the recent literature for such stars in references posterior to 1990 and until June 2013. Only stars with $T_{\rm eff}$ $<$ 7000~K were included to avoid rapid rotators and chemical peculiarities. All determinations not in PASTEL at the time of writing will be included in the database in the next update. We eliminated confirmed non-members, spectroscopic binaries, and chemically peculiar stars and kept only stars with a high probability of membership. Membership information was mainly based on radial-velocity criteria presented in \mbox{\citet{2008AaA...485..303M,2009AaA...498..949M}}. Criteria presented in the articles from which we gathered the spectroscopic determinations and information extracted from the WEBDA\footnote{\url{http://webda.physics.muni.cz}} \citep{2003A&A...410..511M} and Simbad databases were also used for membership evaluation. We started with a list of 571 stars in 86 OCs, with 830 metallicity determinations from 94 papers. In Table~\ref{tab:appendix_members} we list the basic information for the full starting sample of cluster members, which should be sufficient to extract their parameters from the PASTEL catalogue (accessible via VizieR\footnote{\url{http://vizier.u-strasbg.fr/viz-bin/VizieR?-source=B\%2Fpastel}}). \onecolumn \begin{longtab} \centering \begin{longtable}{rlllcll} \caption{\label{tab:appendix_members} List of the full starting sample of cluster members described in Sect.~\ref{sect:hires}.} \\ \hline\hline \noalign{\smallskip} S & Cluster ID & RA (J2000) & DEC (J2000) & Vmag & Bibcode & Star ID \\ \noalign{\smallskip} \hline \noalign{\smallskip} \endfirsthead \caption{ continued.} \\ \hline\hline \noalign{\smallskip} S & Cluster ID & RA (J2000) & DEC (J2000) & Vmag & Bibcode & Star ID \\ \noalign{\smallskip} \hline \noalign{\smallskip} \endhead \noalign{\smallskip} \hline \endfoot \noalign{\smallskip} \hline\hline \noalign{\smallskip} \multicolumn{7}{l}{\parbox{1.00\textwidth}{Column descriptions: The first column identifies the measurements that are included in the final high-resolution sample (S=2, Sect.~\ref{sect:final}; S=1 for the measurements not included in the final sample). The \emph{Bibcode} column gives the ADS Bibcode of the article from which the metallicity determination was taken. The last column gives a star identification, which is resolvable by Simbad, except for five stars in NGC~6791 measured by \citet{2012ApJ...756L..40G}, for which the numbers from \citet{2003PASP..115..413S} are given.}}\\ \endlastfoot \input{table_appendix_members} \end{longtable} \end{longtab} \twocolumn For comparison purposes, we also extracted mean cluster metallicities from a number of studies (post-1990) at high resolution and low S/N ($<50$), and also at medium resolution ($R<25000$, high and low S/N), that is, below the quality criteria defined above. Some of these determinations are discussed in Sect.~\ref{sect:metalrich} at the individual star level, and several are included in Sect.~\ref{sect:mean} at the cluster level. The remaining determinations are not further discussed in this paper. For future reference, we list them in Table~\ref{tab:appendix_lowerquality}. \input{table_appendix_lowerquality} \subsection{Low-resolution sample} As low-resolution spectroscopic metallicity investigations we considered spectra taken at $R \approx 1000\,-\,2000$, from which spectroscopic indices, that is, the strength of absorption features, were measured (predominantly Fe~I and Fe peak blends). A calibration giving the index strength as a function of the atmospheric parameters allows one to determine the metallicity. Low-resolution studies of OC stars are relatively rare in the literature. In addition to a query within ADS, the bibliography of the OC database WEBDA was used, and references in high-resolution studies were checked in this respect. We found \mbox{\citet{1993AaA...267...75F}}, \mbox{\citet{1993PASP..105.1253T}}, \mbox{\citet{2002AJ....124.2693F}}, \mbox{\citet{2003PASP..115...96W}}, and \mbox{\citet{2005AJ....130.1916M}}, the first three of which claim to be on the same metallicity scale. Since \mbox{\citet{2002AJ....124.2693F}} revised the results of the former two studies in combination with new data, we included only their metallicity determinations of 39 OCs in total. Additionally, we adopted the result for the cluster NGC~6705, investigated by \mbox{\citet{1993PASP..105.1253T}}, which is not listed in \mbox{\citet{2002AJ....124.2693F}} due to their restriction to ages older than 0.7~Gyr. \mbox{\citet{2002AJ....124.2693F}} and \mbox{\citet{2005AJ....130.1916M}} determined [Fe/H]\ based on spectroscopic indices defined in \mbox{\citet{1987AJ.....93.1388F}}\footnote{Six and eleven indices were used in the two works, respectively.}. \mbox{\citet{2003PASP..115...96W}} used seven indices on the Lick/IDS system defined in \mbox{\citet{1994ApJS...94..687W}}. \mbox{\citet{2003PASP..115...96W}}, and \mbox{\citet{2005AJ....130.1916M}} seem to present a higher metallicity scale than \mbox{\citet{2002AJ....124.2693F}}. \mbox{\citet{2003PASP..115...96W}} studied two clusters (NGC~188 and NGC~6791) and obtained a metallicity about 0.2~dex higher than \mbox{\citet{2002AJ....124.2693F}} for both clusters. \mbox{\citet{2005AJ....130.1916M}} have three clusters (out of seven) in common with \mbox{\citet{2002AJ....124.2693F}}, indicating a similar tendency (an offset between 0.1 and 0.2~dex). The only exception is NGC~6705, for which the result of \mbox{\citet{2005AJ....130.1916M}} is slightly lower than that of \mbox{\citet[][they agree within the errors]{1993PASP..105.1253T}}. In total, we found 49 metallicity values for 43 individual clusters investigated with low-resolution spectroscopy, of which 34 are also included in the high-resolution sample. Four of the low-resolution clusters have neither photometric nor high-resolution spectroscopic metallicities. Some of these determinations are discussed in Sects.~\ref{sect:m67} to \ref{sect:metalpoor} and in Sect.~\ref{sect:mean}. The remaining determinations are not discussed in this paper. For future reference, we list them in Table~\ref{tab:appendix_lowres}. A detailed comparison of these results, and if possible, a recalibration by means of the high-resolution sample, is planned for the next paper in this series. \section{Assessment of spectroscopic metallicities} \label{sect:assessment} \begin{figure*}[t] \includegraphics[width=\textwidth,trim=0 180 0 280,clip]{mean_paper_before.pdf} \caption{Mean metallicity of OCs per publication, as listed in Table \ref{t:mean_paper}.} \label{f:mean_paper} \end{figure*} \subsection{Mean metallicities and quoted errors for the high-resolution sample} \label{sect:errors} A significant number of individual OCs in the high-resolution sample were studied by different authors with different instruments, methods, and line lists. Throughout this article, we use the term \emph{metallicity} synonymously with \emph{iron abundance}, [Fe/H]. Different authors used different approaches for determining the iron abundances -- using either only Fe~I lines or only Fe~II lines, or both. If abundances for both ions are given, we used the metallicity value as given by the authors, which can be either an unweighted mean of Fe~I and Fe~II abundances or a mean weighted by the number of lines used in each case. In the few cases where metallicity values for individual stars are not clearly stated, we formed an unweighted mean of Fe~I and Fe~II abundances. In the majority of the publications the metallicity is based on Fe~I only. In about 10\% of the cases a weighted mean is given. Another 10\% achieved ionization equilibrium in the analysis by adapting the atmospheric parameters to such an extent that the Fe~I and Fe~II abundances agree exactly. There is also a strong variation in the number of Fe lines used for the analysis, which can be anything between two and 130 Fe~I lines (and individual publications with 180 or 265 lines), and one to 15 Fe~II lines (40 lines in one publication). These variations are partly due to the different telescopes and instruments used to obtain the observed spectra, in particular their wavelength coverage. The observations were obtained at about 30 different telescopes, most of them in the 2--4~m class, and a few at 8--10~m. Echelle spectrographs were used for most observations, providing a large wavelength coverage, but settings focussing on different parts of the optical region were used. The minimum wavelength of the spectra varies between publications from 360 to 700~nm, and the maximum wavelength from 600 to 1060~nm. We show the individual metallicities averaged by OC and by reference in Fig.~\ref{f:mean_paper} and list them in Table~\ref{t:mean_paper}. The figure shows a certain lack of homogeneity for the clusters studied by several authors. The errors on individual [Fe/H] determination quoted by the authors are typically around 0.1~dex, and for most determinations they are less 0.2~dex (see Fig.~\ref{f:errorhistogram}). Only 14 determinations for seven clusters in six papers have quoted metallicity errors between 0.25 and 0.35~dex (Berkeley~22, \mbox{\citealt{2013AJ....145..107J}}; Berkeley~29 and NGC~2141, \mbox{\citealt{2005AJ....130..597Y}}, see also Sect.~\ref{sect:other}; Collinder~261, \mbox{\citealt{2003AJ....126.2372F}}, see also Sect.~\ref{sect:peculiar}; NGC~2112, \mbox{\citealt{1996AJ....112.1551B}}, see also Sect.~\ref{sect:mean}; NGC~3680, \mbox{\citealt{2001AaA...374.1017P}}, see also Sect.~\ref{sect:metalpoor}; NGC~6705, \mbox{\citealt{2012AaA...538A.151S}}). For 81 determinations, the authors do not quote any error for the metallicity. There is a weak dependence of errors on $T_{\rm eff}$\ and [Fe/H], such that the largest errors are quoted for stars with $T_{\rm eff}$$<$5000~K and [Fe/H]$<-0.1$~dex. We note that the errors quoted by most authors are in fact the standard deviations of the abundances determined from the selected Fe lines, with the exception of \mbox{\citet{2010AaA...511A..56P}} and \mbox{\citet{2011AaA...535A..30C}}. These authors quote the standard error of the line abundances, and we multiplied their errors by the square root of the number of lines used. The uncertainty in stellar parameters causes an additional uncertainty in the Fe abundances. This type of external error is typically estimated to be 0.1~dex in the publications included in this work. An additional source of systematic differences between different studies might arise from the choice of the solar reference metallicity. We did not assess the extension of this effect for each individual paper. However, for the studies that are based on a differential analysis with respect to a solar spectrum or that use astrophysical oscillator strengths, the derived or adopted value of the solar Fe abundance is not important. We expect this to be the case for the majority of the publications. The remaining studies may be affected by external errors of up to 0.1~dex. Standard deviations of the mean cluster metallicity from each individual paper are mostly lower than 0.1~dex, with a peak below 0.05~dex, which shows that the internal uncertainties of metallicity determinations are probably lower than the quoted errors (see Fig.~\ref{f:errorhistogram}). When we combine all determinations per cluster from all papers, the peak of the standard deviations is slightly higher than 0.05~dex, which suggests that external errors are inherent in the datasets. In the following sections, we investigate some cases with a large number of metallicity determinations by different authors, including low-resolution studies, and some cases with large standard deviations around the mean metallicity. In the high-resolution sample, there are 26 OCs for which metallicity determinations are available for fewer than three stars. The reliability of the metallicities for these clusters is difficult to assess. We will apply the conclusions drawn from the comparisons for the well-studied clusters\footnote{The starting sample contains 47 clusters with three to ten stars, ten clusters with twelve to 20 stars, two clusters with 29 stars, and one with 76.} to these poor-studied ones for the selection of the final spectroscopic OC metallicity sample. \onecolumn \begin{longtab} \begin{longtable}{ r l l l l r l } \caption{\label{t:mean_paper} Weighted mean metallicity (column 5) of OCs, computed for each reference (column 7) giving atmospheric parameters based on high-resolution, high-S/N spectra. Only highly probable members and non-SBs were considered. See also Fig.~\ref{f:mean_paper}.} \\ \hline\hline \noalign{\smallskip} (1) & (2) & (3) & (4) & (5) & (6)& (7) \cr OC no. & Name & $T_{\rm eff}$\ range & $\log g$\ range & Mean [Fe/H] & \# & Reference \cr \noalign{\smallskip} \hline \noalign{\smallskip} \endfirsthead \caption{ continued.} \\ \hline\hline \noalign{\smallskip} OC no. & Name & $T_{\rm eff}$\ range & $\log g$\ range & Mean [Fe/H] & \# & Reference \cr \noalign{\smallskip} \hline \noalign{\smallskip} \endhead \noalign{\smallskip} \hline \endfoot \noalign{\smallskip} \hline\hline \noalign{\smallskip} \multicolumn{7}{l}{\parbox{1.00\textwidth}{The $T_{\rm eff}$\ and $\log g$\ ranges and the number of determinations (\#) are given in columns 3, 4, and 6. In most cases, the latter corresponds to the number of stars analysed. In the case of \citet{2009AaA...493..309S} giant stars are counted twice because two different line lists were used. In the cases of \citet{2000AaA...360..499T} and \citet{2004AaA...424..951P}, for three of seven and three of 21 stars, respectively, two or three different spectra were used. The resulting values were treated as independent determinations.}}\\ \endlastfoot \input{table_mean_paper} \end{longtable} \end{longtab} \twocolumn \begin{figure} \includegraphics[width=\columnwidth,trim=40 40 60 60,clip]{errorhistogram.pdf} \caption{Distribution of metallicity errors quoted by authors (red line) compared with the distributions of standard deviations for mean metallicities per cluster for each paper (blue line), and for all papers (black line), combining determinations for at least three stars.} \label{f:errorhistogram} \end{figure} \subsection{Solar-metallicity cluster: M67} \label{sect:m67} \begin{figure*} \centering \includegraphics[width=\textwidth,trim=10 50 10 50,clip]{M67-compare-giants.pdf} \caption{ Metallicity versus reference $T_{\rm eff}$\ for individual \emph{giant} stars in NGC~2682 (M67) that were analysed by more than one author using high- and low-resolution spectra. The metallicities are taken from the following publications: \citet{1980ApJ...241..981C,1981AaA....99..221F,2000AaA...360..499T,2002AJ....124.2693F,2005AJ....130..597Y,2009AaA...493..309S,2010AaA...511A..56P,2010AJ....139.1942F,2013AJ....145..107J}; and \citet{2013MNRAS.431.3338R}. Data for the following stars are shown (reference $T_{\rm eff}$, number from \citealt{1906PhDT.........1F}): (4250, 108); (4280, 170); (4450, 105); (4700, 164); (4710, 224); (4730, 141); (4730, 266); (4750, 84); (4760, 151); and (4800, 286). The bars at the upper right indicate minimum, mean, and maximum uncertainties for the individual [Fe/H] values quoted in the publications. } \label{fig:M67giants} \end{figure*} Metallicities for \emph{giant} stars ($\log g$ $\le$ 3.0) in M67 (NGC~2682) have been determined in a considerable number of publications since 1990, including the low-resolution studies by \mbox{\citet{2002AJ....124.2693F}} and \mbox{\citet{2005AJ....130.1916M}}. Seven studies published after 1990 are based on high-resolution, high-S/N spectra. We plot the metallicities determined for ten individual giant stars by more than one author at high and low resolution in Fig.~\ref{fig:M67giants} as a function of reference $T_{\rm eff}$\ (see figure caption for references). The reference $T_{\rm eff}$\ values were taken from \mbox{\citet{2000AaA...360..499T}} except for Fagerholm 286 (average of \mbox{\citealt{2009AaA...493..309S}} and \mbox{\citealt{2010AaA...511A..56P}}). Each $T_{\rm eff}$\ value corresponds to one star, except for $T_{\rm eff}$=4730~K, which corresponds to two stars. The reference $T_{\rm eff}$\ values agree to within 100~K with those given in the individual publications. The internal uncertainties for the individual [Fe/H]\ values quoted by the authors range from 0.04 \mbox{\citep{2013MNRAS.431.3338R}} to 0.17~dex \mbox{\citep{2013AJ....145..107J}}, with a mean of 0.12~dex. In addition, we show metallicity determinations for four of these stars from two studies made before CCD detectors became available \mbox{\citep{1980ApJ...241..981C,1981AaA....99..221F}}, which are based on somewhat lower-resolution spectra, and atmospheric models and atomic data available at that time. It is obvious that the metallicity determined from spectroscopic indices \mbox{\citep{2002AJ....124.2693F}} is on a more metal-poor scale than the high-resolution metallicities ($\sim$0.1~dex difference). One of the older studies \mbox{\citep[][$R\approx$17000]{1980ApJ...241..981C}} shows a large systematic offset from the high-resolution studies, while the other one \mbox{\citep{1981AaA....99..221F}} agrees very well. All but five of the high-resolution metallicities are confined within $\pm 0.06$~dex, with no obvious dependence on resolution or $T_{\rm eff}$. \mbox{\citet[][Sect. 6.1.1]{2010AJ....139.1942F}} compared their results for three stars (Fagerholm 105, 141, and 170) with three other studies (for Fe as well as other elements), and arrived at similar conclusions. The extended comparison presented here indicates that older abundance-analysis techniques applied to solar metallicity giant stars have larger systematic uncertainties than modern ones (post-1990). \begin{figure*} \centering \includegraphics[width=\textwidth,trim=10 50 10 50,clip]{M67-compare-dwarfs.pdf} \caption{ Same as Fig.~\ref{fig:M67giants}, for \emph{dwarf} stars in M67 (NGC~2682). The metallicities are taken from the following publications: \citet{1991AJ....102.1070H,1992ApJ...387..170F,2000AJ....120.1913S,2006AaA...450..557R,2008AaA...489..403P}; and \citet{2009AaA...493..309S}. Data for the following stars are shown (reference $T_{\rm eff}$, number from \citealt{1977A&AS...27...89S}): (5750, 746); (5850, 1255); (5900, 1048); (5950, 1256); (6050, 1283); (6100, 1092); (6100, 1287); (6150, 994); (6200, 998); and (6600, 997). The S/N of the spectra used in the different publications is higher than 50, except when stated otherwise. The bars to the right of the legend indicate minimum and maximum uncertainties for the individual [Fe/H] values quoted in the publications. } \label{fig:M67dwarfs} \end{figure*} For \emph{dwarf} stars in M67 we found metallicities in six publications, all based on high-resolution and high S/N spectra. Ten dwarf stars were analysed by at least two authors. We plot the metallicities determined for these stars in Fig.~\ref{fig:M67dwarfs} as a function of reference $T_{\rm eff}$\ on the same vertical scale as in Fig.~\ref{fig:M67giants} for the giant stars (see figure caption for references). The reference $T_{\rm eff}$\ values are averages of the determinations from different works. Each $T_{\rm eff}$\ value corresponds to one star, except for $T_{\rm eff}$=6100~K, which corresponds to two stars (see figure caption). The reference $T_{\rm eff}$\ values agree to within 100~K with those given in the individual publications. The quoted uncertainties for the individual [Fe/H]\ values range from 0.03 \mbox{\citep{2006AaA...450..557R}} to 0.13~dex \mbox{\citep{1991AJ....102.1070H}}. Among the dwarf analyses, the oldest study by \mbox{\citet{1991AJ....102.1070H}} stands out among the others because it shows the largest dispersion. Otherwise, the spread in abundances is the same as for the giant stars. The lower-resolution metallicities agree with the others, but the small number of these points does not allow a general conclusion. The star with the highest temperature (Sanders 997) is an M67 blue straggler and a probable spectroscopic binary star \mbox{\citep{2000AJ....120.1913S}}. A comparison of dwarf and giant metallicities for this extraction from the starting sample of cluster members results in equal mean values within the standard deviations (cf. Sect.~\ref{sect:final} for a comparison based on the full final sample). The mean of the 28 post-1990 high-resolution values for giants (full and open black symbols in Fig.~\ref{fig:M67giants}) is $-0.01\pm0.05$~dex, while the mean of the 15 most recent high-resolution values for dwarfs (full black symbols in Fig.~\ref{fig:M67dwarfs}) is 0.02$\pm$0.04~dex. \subsection{Metal-rich cluster: NGC~6791} \label{sect:metalrich} \begin{figure*} \centering \includegraphics[width=\textwidth,trim=10 50 10 50,clip]{NGC6791-compare-giants.pdf} \caption{Same as Fig.~\ref{fig:M67giants}, for giant stars in NGC~6791. The metallicities are taken from the following publications: \citet{2003PASP..115...96W,2002AJ....124.2693F,2012ApJ...756L..40G,2006ApJ...643.1151C}; and \citet{2006ApJ...642..462G}. Data for the following stars are shown (reference $T_{\rm eff}$\ in K, star number from \citealt{2003PASP..115..413S}): (3860, 8904); (3930, 7972); (3940, 8266); (3970, 5342); (4000, 11814); (4090, 8563); (4120, 10898); (4130, 8988); (4190, 4952); (4200, 6288); (4400, 3369); (4410, 4715); (4420, 7922); (4430, 2723); (4440, 9462); (4450, 10806); (4460, 8082); and (4470, 9316). The bars to the right indicate minimum and maximum uncertainties for the individual [Fe/H] values quoted in the publications. } \label{fig:NGC6791} \end{figure*} For NGC~6791, we found five publications since 1990 with metallicity determinations for 18 giant stars appearing in at least two of them. We plot the metallicities determined for these stars in Fig.~\ref{fig:NGC6791} as a function of reference $T_{\rm eff}$\ on the same vertical and horizontal scale as in Fig.~\ref{fig:M67giants} for the M67 giant stars (see figure caption for references). The reference $T_{\rm eff}$\ values are averages of the determinations from different works. Each $T_{\rm eff}$\ value corresponds to one star (see figure caption). The reference $T_{\rm eff}$\ values agree to within 100~K with those given in the individual publications. The minimum and maximum uncertainties for the individual [Fe/H]\ values were found in two low-resolution studies and represent the standard deviations of metallicities determined from several indices (0.01~dex in \mbox{\citealt{2003PASP..115...96W}}, and in 0.2~dex \mbox{\citealt{2002AJ....124.2693F}}). There are no multiple metallicity determinations available for dwarf stars in this cluster. Again, the values from spectroscopic indices by \mbox{\citet{2002AJ....124.2693F}} are lower than the others, by a larger amount than at solar metallicity ($\sim$0.3~dex). However, the indices-based metallicities by \mbox{\citet{2003PASP..115...96W}} are close to the values derived from higher-resolution spectra. There is one star in common between the medium- and high-resolution studies of \mbox{\citet{2012ApJ...756L..40G}}, \mbox{\citet{2006ApJ...643.1151C}}, and \mbox{\citet{2006ApJ...642..462G}}, where the highest-resolution study gives the highest metallicity. The error quoted by \mbox{\citet{2006ApJ...642..462G}} is twice as large as the difference, however, and thus the difference is not significant. Nine additional stars are in common between \mbox{\citet{2012ApJ...756L..40G}} and \mbox{\citet{2006ApJ...643.1151C}}, and for all except two stars, the metallicities agree very well. The mean metallicity values derived from the whole samples of these publications show differences similar to the standard deviations. \mbox{\citet{2012ApJ...756L..40G}} determined [Fe/H]=0.42$\pm$0.05 for 16 stars, \mbox{\citet{2006ApJ...643.1151C}} determined [Fe/H]=0.38$\pm$0.02 for ten stars, and \mbox{\citet{2006ApJ...642..462G}} determined [Fe/H]=0.47$\pm$0.07 for four stars\footnote{Note that Stetson 8082 is not included in Table~\ref{t:mean_paper} because its S/N is lower than 50.}. These differences are probably not caused by the different resolutions, because the analysis of five subgiants at $R=45000$ by \mbox{\citet{2012ApJ...756L..40G}} results in a mean metallicity that agrees with all three of the medium- and high-resolution studies of giants within the standard deviations (see Table~\ref{t:mean_paper}). Reasons for the differences could be the different instruments used (Hydra at WIYN and SARG at TNG), different stellar samples, different spectrum synthesis codes, and different line lists (\mbox{\citealt{2006ApJ...642..462G}} used the broadest wavelength range). \subsection{Metal-poor clusters} \label{sect:metalpoor} \begin{figure*} \centering \includegraphics[width=\textwidth]{metal-poor-compare-giants.pdf} \caption{ Same as Fig.~\ref{fig:M67giants}, for giant stars in ten different metal-poor clusters (five each in the upper and lower panels). The metallicities are taken from the following publications: \citet{1993AaA...267...75F,1994AaA...283..911G,2001AaA...374.1017P,2002AJ....124.2693F,2004AJ....128.1676C,2005AJ....130..597Y,2006AaA...458..121S,2008AaA...488..943S}; \citet[][Table~3]{2009AaA...493..309S}; \citet{2010AJ....139.1942F}; \citet{2011AaA...535A..30C}; \citet{2012AJ....144...95Y}; and \citet{2013AJ....145..107J}. For star identifications and corresponding reference $T_{\rm eff}$\ see Table~\ref{tab:metalpoor}. The bar below the legend in the upper panel indicates the mean of the uncertainties for the individual [Fe/H] values quoted in the high-resolution publications (ranging from 0.04 to 0.3~dex). } \label{fig:metalpoor} \end{figure*} \begin{table} \caption{Reference $T_{\rm eff}$\ values and identifications of giant stars in ten different metal-poor clusters (see Fig.~\ref{fig:metalpoor}).} \label{tab:metalpoor} \centering \begin{tabular}{lllrlr} \hline\hline \noalign{\smallskip} $T_{\rm eff}$ & Cluster & \multicolumn{2}{c}{ID 1} & \multicolumn{2}{c}{ID 2}\\ \noalign{\smallskip} \hline \noalign{\smallskip} 3930 & Berkeley 22 & K & 643 & & \\ 3980 & Pfleiderer 4 & RGB & 1 & \multicolumn{2}{l}{J23505744+6220031\tablefootmark{$\dagger$}} \\ 3990 & Melotte 66 & KJF & 2236 & & 4151 \\ 4100 & Berkeley 32 & KM & 2 & DBT & 2689 \\ 4100 & Berkeley 32 & KM & 4 & DBT & 1556 \\ 4110 & NGC 2243 & & 4209 & MMU & 3633 \\ 4380 & Berkeley 22 & K & 414 & & \\ 4400 & Berkeley 18 & K & 1383 & & \\ 4400 & Berkeley 20 & MPJF & 8 & SBR & 1240 \\ 4500 & NGC 2243 & & 4110 & MMU & 1313 \\ 4510 & Berkeley 21 & TPM & 51 & & 415a\tablefootmark{$\ddagger$} \\ 4510 & NGC 3680 & EGG & 44 & AHTC & 1031 \\ 4580 & Berkeley 18 & K & 1163 & & \\ 4640 & Berkeley 21 & TPM & 50 & & 50\tablefootmark{$\ddagger$} \\ 4650 & NGC 3680 & EGG & 26 & AHTC & 3017 \\ 4660 & NGC 3680 & EGG & 13 & AHTC & 3003 \\ 4660 & NGC 3680 & EGG & 53 & KGP & 1873 \\ 4680 & NGC 3680 & EGG & 41 & AHTC & 1050 \\ 4700 & Berkeley 32 & KM & 12 & DBT & 1393 \\ 4750 & Melotte 66 & KJF & 1953 & SBR & 1346 \\ 4760 & Berkeley 32 & KM & 19 & DBT & 787 \\ 4780 & Berkeley 32 & KM & 27 & DBT & 605 \\ 4830 & Berkeley 32 & KM & 17 & DBT & 533 \\ 4890 & Berkeley 32 & KM & 16 & DBT & 737 \\ 4980 & Berkeley 32 & KM & 18 & DBT & 997 \\ 5020 & Berkeley 29 & BHT & 398 & FMP & 948 \\ \noalign{\smallskip} \hline\hline \noalign{\smallskip} \end{tabular} \tablefoot{ Columns \emph{ID~1} and \emph{ID~2} give two alternative identifications, with acronyms used by the Simbad database, with some exceptions. \emph{ID~1} corresponds to the numbering system adopted by WEBDA, except for Berkeley~32 and Pfleiderer~4. \tablefoottext{$\dagger$}{2MASS} \tablefoottext{$\ddagger$}{Numbering by \citet{1979AJ.....84..204C}} } \end{table} To evaluate the influence of spectrum quality and analysis methods on the metallicities of metal-poor clusters, we compared results from 13 publications since 1990 for giant stars appearing in ten different clusters. For these clusters, at least one determination gives a value of [Fe/H]=$-0.3$~dex or lower. We plot the metallicities determined for the individual stars in Fig.~\ref{fig:metalpoor} as a function of $T_{\rm eff}$\ (see figure caption for references). The reference $T_{\rm eff}$\ values are taken from the works with highest resolution (or are averages from works with the same resolution). Each $T_{\rm eff}$\ value corresponds to one star, except for $T_{\rm eff}$=4100~K in the lower panel and 4660~K in the upper panel, which correspond to two stars each. The reference $T_{\rm eff}$\ values and identifications of the stars are listed in Table~\ref{tab:metalpoor}. The reference $T_{\rm eff}$\ values agree to within 200~K with those given in the individual publications. The quoted uncertainties for the individual [Fe/H]\ values range from 0.08 \mbox{\citep{2009AaA...493..309S}} to 0.3~dex \mbox{\citep{2001AaA...374.1017P,2013AJ....145..107J}}, and the mean of the uncertainties for high-resolution studies is 0.17~dex. For these clusters, we can mainly compare the metallicities by \citet[][Melotte~66, Berkely~21]{1993AaA...267...75F} and \mbox{\citet{2002AJ....124.2693F}} based on spectroscopic indices to the high-resolution metallicities. The mean star-by-star difference for 14 stars (excluding two stars in \mbox{\citealt{2001AaA...374.1017P}}) is $-0.16\pm0.07$~dex, intermediate between the differences for M67 and NGC~6791 (both at higher metallicity). Five of the stars appear in two or more high-resolution studies that are based on different observations: Berkeley~29 BHT~398, Berkeley~20 MPJF~8, and Berkeley~32 KM~17, as well as two stars in NGC~3680 (discussed below). For the two stars in Berkeley~29 and Berkeley~20, the study by \mbox{\citet[][S/N=25--50 and 40--80, respectively]{2008AaA...488..943S}} resulted in higher metallicities (by 0.14~dex) than the studies by \mbox{\citet{2004AJ....128.1676C}} and \mbox{\citet[][S/N=70 and 56, respectively]{2005AJ....130..597Y}}, while the determinations for the star in Berkeley~32 by \mbox{\citet{2006AaA...458..121S}} and \mbox{\citet{2011AaA...535A..30C}} are in excellent agreement. There are two other stars in Berkeley~32 (at $T_{\rm eff}$=4100~K), for which \mbox{\citet{2010AJ....139.1942F}} and \mbox{\citet{2013AJ....145..107J}} analysed the same observed spectra at $R=30000$ and obtained consistent metallicities. On the other hand, the lower panel of Fig.~\ref{fig:metalpoor} shows nine stars in five clusters, for which \mbox{\citet{2012AJ....144...95Y}} and \mbox{\citet{2013AJ....145..107J}} analysed the same spectra at $R=47000$. The derived metallicities are systematically different by 0.14~dex. Note that the largest differences of about 0.2~dex are seen for the two stars with $T_{\rm eff}$$<$4000~K. At this point, we refer to the extensive comparison work presented by \mbox{\citet[][Appendix]{2010AJ....139.1942F}} for Berkeley~32. They used the line lists and equivalent widths measured for nine stars by \mbox{\citet{2006AaA...458..121S}} and determined abundances and atmospheric parameters with their own methods. A detailed discussion of the possible effects of variations in individual analysis ingredients is given. For one star (KM~18), they independently measured equivalent widths in the spectrum used by \mbox{\citet{2006AaA...458..121S}} and found excellent agreement. On the other hand, using different selections of lines for the abundance determination (for a fixed set of atmospheric parameters) resulted in [Fe/H] differences of up to 0.15~dex. For NGC~3680, the metallicities from high-resolution studies cluster around two significantly different values (no. 63 in Table~\ref{t:mean_paper}): \mbox{\citet{2009AaA...493..309S}} obtained a solar value (based on three giant and two dwarf stars), close to the results by \mbox{\citet{2009AaA...502..267S}} for one of the three giants and the results of \mbox{\citet{2008AaA...489..403P}} for the same two dwarfs at $R$=100000. On the other hand, \mbox{\citet{2001AaA...374.1017P}} determined low metallicities ([Fe/H]$\approx -0.3$~dex) for six giant stars, which include the three stars studied by \mbox{\citet{2009AaA...493..309S}} and \mbox{\citet[][see Fig.~\ref{fig:metalpoor}]{2009AaA...502..267S}}. First, we note that \mbox{\citet{2001AaA...374.1017P}} quoted metallicity errors of 0.25 to 0.35~dex, which are among the largest quoted in any publication. These errors were computed in an unconventional way -- standard deviations of line abundances (interpreted as random errors due to uncertainties in equivalent widths and atomic parameters) were added linearly and not in quadrature to metallicity errors due to uncertainties in atmospheric parameters (estimated to be 0.18~dex). The errors are thus probably overestimated compared with those of other authors. Second, for their final cluster metallicity, \mbox{\citet{2001AaA...374.1017P}} excluded the star EGG~41 with the lowest metallicity, because they did not trust the $T_{\rm eff}$\ value, and the star EGG~34 ([Fe/H]=$-$0.07~dex, not shown in Fig.~\ref{fig:metalpoor}), because they suspected it to be a binary star. This does not change the low cluster metallicity ([Fe/H]=$-$0.27$\pm$0.03~dex). In addition, they added a +0.1~dex systematic error estimated from two Hyades stars, and quoted a final cluster metallicity of $-$0.17$\pm$0.12~dex -- closer to the other studies, but still significantly lower. Because of these discrepancies, which indicate systematic errors in the determinations by \mbox{\citet{2001AaA...374.1017P}}, we decided to disregard these determinations for the computation of the average metallicity of NGC~3680 (Sect.~\ref{sect:final}). \subsection{Peculiar case: Collinder 261} \label{sect:peculiar} \begin{table} \caption{Atmospheric parameters and identifications of giant stars in Collinder~261 (see Fig.~\ref{fig:Collinder261}).} \label{tab:Collinder261} \centering \begin{tabular}{rrrllr} \hline\hline \noalign{\smallskip} $T_{\rm eff}$ & $\log g$ & [Fe/H] & Reference & \multicolumn{2}{c}{ID} \\ \noalign{\smallskip} \hline \noalign{\smallskip} 3950 & 0.5 & $-$0.01 & \citet{2007AJ....133.1161D} & PJM & 1871 \\ 3980 & 0.4 & $-$0.32 & \citet{2005AaA...441..131C} & PJM & 1871 \\ 4000 & 0.7 & $-$0.31 & \citet{2003AJ....126.2372F} & PJM & 1871 \\ 4180 & 1.6 & $-$0.08 & \citet{2005AaA...441..131C} & PJM & 2105 \\ 4300 & 1.8 & $-$0.02 & \citet{2007AJ....133.1161D} & PJM & 1485 \\ 4300 & 1.5 & $-$0.32 & \citet{2003AJ....126.2372F} & PJM & 2105 \\ 4340 & 1.8 & $-$0.06 & \citet{2005AaA...441..131C} & PJM & 1485 \\ 4350 & 1.7 & +0.12 & \citet{2008AaA...488..943S} & SBR & 2 \\ 4400 & 2.1 & $-$0.03 & \citet{2007AJ....133.1161D} & PJM & 1481 \\ 4400 & 1.5 & $-$0.16 & \citet{2003AJ....126.2372F} & PJM & 1045 \\ 4450 & 1.8 & $-$0.01 & \citet{2007AJ....133.1161D} & PJM & 1045 \\ 4470 & 2.1 & +0.01 & \citet{2005AaA...441..131C} & PJM & 1045 \\ 4490 & 2.2 & $-$0.11 & \citet{2003AJ....126.2372F} & PJM & 1080 \\ 4500 & 2.1 & +0.00 & \citet{2005AaA...441..131C} & PJM & 1080 \\ 4500 & 2.1 & +0.02 & \citet{2007AJ....133.1161D} & PJM & 1080 \\ 4500 & 2.0 & +0.01 & \citet{2007AJ....133.1161D} & PJM & 27 \\ 4500 & 2.3 & +0.16 & \citet{2008AaA...488..943S} & SBR & 6 \\ 4500 & 1.9 & $-$0.03 & \citet{2007AJ....133.1161D} & PJM & 29 \\ 4546 & 2.2 & +0.18 & \citet{2008AaA...488..943S} & SBR & 7 \\ 4550 & 2.0 & +0.00 & \citet{2007AJ....133.1161D} & PJM & 2001 \\ 4580 & 1.8 & $-$0.02 & \citet{2005AaA...441..131C} & PJM & 2001 \\ 4600 & 2.0 & +0.00 & \citet{2007AJ....133.1161D} & PJM & 1526 \\ 4600 & 2.0 & +0.14 & \citet{2008AaA...488..943S} & SBR & 5 \\ 4600 & 2.0 & $-$0.01 & \citet{2007AJ....133.1161D} & PJM & 1801 \\ 4650 & 2.3 & $-$0.01 & \citet{2007AJ....133.1161D} & PJM & 1472 \\ 4670 & 2.2 & +0.09 & \citet{2008AaA...488..943S} & SBR & 11 \\ 4700 & 2.4 & +0.20 & \citet{2008AaA...488..943S} & SBR & 10 \\ 4720 & 2.1 & +0.04 & \citet{2008AaA...488..943S} & SBR & 9 \\ \noalign{\smallskip} \hline\hline \noalign{\smallskip} \end{tabular} \tablefoot{Column \emph{ID} gives the star identification, with acronyms used by the Simbad database.} \end{table} The cluster Collinder~261 has been studied in four different publications, all at high resolution and high S/N. The analyses resulted in four different cluster metallicities, from $-0.2$ to +0.1~dex in steps of 0.1~dex (no. 12 in Table~\ref{t:mean_paper}). Figure~\ref{fig:Collinder261} shows the metallicity determinations for individual giant stars from all four works and their [Fe/H]\ uncertainties. Six stars were studied by more than one author, and the data for these are connected with lines in the figure. Star identifications, data, and references are given in Table~\ref{tab:Collinder261}. All four works are based on similar-quality data from three different instruments (S/N$\approx$100, $R\approx 45000$, except for \mbox{\citealt{2003AJ....126.2372F}} with $R\approx 25000$). All four works are based on an equivalent-width analysis, and three of them used the same code for computing the model equivalent widths \mbox{\citep[MOOG,][]{1973PhDT.......180S}}. They derived the stellar atmospheric parameters in the same way, forcing excitation and ionization equilibrium on the line abundances. Thus, the diverging results obtained by these authors could be due to differences in the model atmospheres, spectral-line selection, and atomic data. Additional probable sources for discrepancies are continuum tracing and equivalent-width measurement \mbox{\citep{2008AaA...488..943S}}. In three of the publications, the model atmospheres were those of R.~Kurucz, although different versions were used: \mbox{\citet{2005AaA...441..131C}} and \mbox{\citet{2008AaA...488..943S}} used models from the grid on CDROM \mbox{\citep{1993KurCD..13.....K}}, while \mbox{\citet{2007AJ....133.1161D}} interpolated in the grid published by \mbox{\citep{1997AaA...318..841C}}. \mbox{\citet{2003AJ....126.2372F}} interpolated in the grid of \mbox{\citet{1976AaAS...23...37B}}. In Table~\ref{tab:Col261lines}, we list the number of iron lines used per star and the corresponding wavelength ranges. These properties of the line lists are very similar, except for that of \mbox{\citet{2007AJ....133.1161D}}, which includes bluer wavelengths. All four works derived metallicities with respect to a reference object. \mbox{\citet{2003AJ....126.2372F}} used oscillator-strength values derived from an Arcturus spectrum. \mbox{\citet{2005AaA...441..131C}} and \mbox{\citet{2008AaA...488..943S}} quoted their Fe abundances relative to abundances derived from a solar analysis using equivalent widths measured from high-resolution solar atlases. The homogeneity of the abundances by \mbox{\citet{2007AJ....133.1161D}} within the cluster was achieved by a differential line-by-line abundance analysis relative to one of the cluster stars, whereas the abundance zero-point was given by the adopted solar abundance from \mbox{\citet{1992AJ....104.2121S}}. These analysis approaches should minimize the uncertainties caused by atomic line data, although to different degrees. The determination of gravity could be a major source for the differences, because it relies on ionization balance between abundances from Fe~I and Fe~II lines. Since Fe~II lines are scarce in spectra of cool stars, the derived gravity critically depends on the set of Fe~II lines used and the accuracy of their equivalent widths. However, for only two of the six stars in common, there are significant differences in gravity (PJM~1871 -- the coolest star, and PJM~1045 at 4400--4500~K, see Table~\ref{tab:Collinder261}), and the differences in gravity do not seem to be correlated with differences in abundances. \begin{table} \caption{Abundance analyses of Collinder~261 stars -- approximate number of Fe~I and Fe~II lines used per star, and corresponding wavelength ranges.} \label{tab:Col261lines} \centering \begin{tabular}{lrrrr} \hline\hline \noalign{\smallskip} & \multicolumn{2}{c}{Fe~I} & \multicolumn{2}{c}{Fe~II} \\ Reference & n & $\lambda$ [nm] & n & $\lambda$ [nm] \\ \noalign{\smallskip} \hline \noalign{\smallskip} \citet{2003AJ....126.2372F} & 40 & 538 -- 785 & 8 & 541 -- 652 \\ \citet{2005AaA...441..131C} & 100 & 550 -- 700 & 10 & 550 -- 700 \\ \citet{2007AJ....133.1161D} & 60 & 422 -- 620 & 12 & 449 -- 615 \\ \citet{2008AaA...488..943S} & 100 & 550 -- 681 & 13 & 553 -- 652 \\ \noalign{\smallskip} \hline\hline \end{tabular} \end{table} \mbox{\citet{2005AaA...441..131C}} provided a detailed comparison of their work with \mbox{\citet{2003AJ....126.2372F}}. They showed that the equivalent widths of \mbox{\citet{2003AJ....126.2372F}} are systematically smaller for stronger lines ($\gtrsim$80~m\AA) for the lines in common (among which there are 24 Fe~I lines). This could in part explain the lower metallicities. On the other hand, \mbox{\citet{2005AaA...441..131C}} used the \mbox{\citet{2003AJ....126.2372F}} equivalent widths and their own methods for the star with the largest abundance difference (PJM~2105), and derived an [Fe/H]\ value close to that obtained from their own data. They ascribed this to their different approach in estimating the microturbulent velocity (based on computed instead of observed equivalent widths). For the test star, the microturbulence derived by \mbox{\citet{2005AaA...441..131C}} is indeed 0.25~kms$^{-1}$ lower. However, for two other stars with (smaller) abundance differences, the microturbulence is about 0.1~kms$^{-1}$ higher. \mbox{\citet{2007AJ....133.1161D}} compared their results with those of \mbox{\citet{2005AaA...441..131C}}, and noted that they are similar except for the coolest star with lowest gravity (PJM~1871). They also performed a test on three stars, using two different types of model atmospheres (Kurucz and MARCS, \mbox{\citealt{1997AaA...318..521A}}), and found that the differences in Fe abundance were lower than 0.03~dex. \begin{figure} \centering \includegraphics[width=\columnwidth,trim=10 50 10 50,clip]{Collinder261-compare.pdf} \caption{ Metallicity versus $T_{\rm eff}$\ for individual giant stars in Collinder~261. For star identifications, data, and references see Table~\ref{tab:Collinder261}. } \label{fig:Collinder261} \end{figure} Collinder~261 is a rather old OC, located in the inner disk, which makes it an interesting object for studying the evolution of metallicity distributions in the Galactic disk. It is included in the list of standard clusters proposed by \mbox{\citet{2006MNRAS.371.1641P}}, with an age of 8.8~Gyr and a distance of 2.2~kpc. There is no photometric metallicity determination available, that is, it is not included in Paper~I. Therefore, we used the method of \mbox{\citet{2010AaA...514A..81P}} (see also Sect.~\ref{sect:final}) to estimate its metallicity using $B, V, I$ photometry, published by \mbox{\citet{1996MNRAS.283...66G}} and \mbox{\citet[][$V$ and $I$ only]{1994AJ....107.1079P}}. The \mbox{\citet{1994AJ....107.1079P}} photometry was brought onto the scale by \mbox{\citet{1996MNRAS.283...66G}} by applying corrections of +0.031 and +0.033~mag for $V$ and $V-I$, respectively. An average and standard deviation of the photometry was calculated. We selected only stars available in both studies with errors $<$~0.05~mag, covering the whole cluster. The colour-colour and colour-magnitude diagrams were examined to select most probable main-sequence cluster members. The method is based on grids of evolutionary models, originally for [Fe/H] from $-0.7$ to $+0.4$ and ages up to 4~Gyr, which were extended to older ages using Geneva isochrones \mbox{\citep{2001A&A...366..538L}}. To transform $V-I$ into effective temperatures (in addition to $B-V$), the empirical relation for A to K dwarfs by \mbox{\citet{1998A&A...333..231B}} as well as their colour-excess ratios (including colour terms) were used.To transform the photometry into the $T_\mathrm{eff}-\mathrm{log}(L/L_{\odot})$ plane, the cluster parameters by \mbox{\citet{2006MNRAS.371.1641P}} were adopted. Since only $E(B-V)$ is tabulated, the temperatures deduced from $V-I$ were scaled to the $B-V$ results by applying an offset of $-$0.04~mag to the transformed reddening value. This offset can either be due to a small error in the calibration or to an abnormal reddening law in this direction. The final temperatures from both indices agree within $<$\,2\%, and were averaged. All other steps of the method were applied as given in \mbox{\citet{2010AaA...514A..81P}}. \begin{figure} \includegraphics[width=\columnwidth,trim=10 10 5 5,clip]{col261.pdf} \caption{Theoretical HR-diagram illustrating the photometric metallicity determination for Collinder~261 using the method by \citet{2010AaA...514A..81P}. Logarithmic luminosity is plotted versus normalized logarithmic effective temperature, $T_{\rm N}=\log$$T_{\rm eff}$$-\log$$T_{\rm eff}$(ZAMS), where $T_{\rm eff}$(ZAMS) is the $T_{\rm eff}$\ of a zero-age main-sequence with solar composition for a given luminosity (see Table~4 of \citealt{2010AaA...514A..81P}).} Using the parameters by \citet{2006MNRAS.371.1641P}, the metallicity was determined as $Z$=0.021 ([Fe/H]$\approx$0.05~dex, isochrone shown as thick line). In addition, the isochrone for $Z$=0.012 ([Fe/H]$\approx-$0.2, thin line) is shown, representing the high-resolution metallicity value of \citet{2003AJ....126.2372F}. \label{f:cr261hrd} \end{figure} In Fig.~\ref{f:cr261hrd}, the best-fit isochrone for the given parameters can be found, which results in a metallicity estimate of $Z$=0.021 (heavy-element-mass fraction, corresponding to [Fe/H]$\approx$0.05~dex for $Z_\odot$=0.019 and a helium-mass fraction $Y=0.23+2.25Z$). This supports the spectroscopic results around solar metallicity or higher. In addition, the low-resolution result by \mbox{\citet{2002AJ....124.2693F}} with [Fe/H]=$-$0.16~dex would support solar metallicity, when applying an ad-hoc correction of +0.2~dex, as suggested by comparison with high-resolution results (see e.g. Sect.~\ref{sect:mean}). In conclusion, we decided to disregard the determinations by \mbox{\citet{2003AJ....126.2372F}} for the computation of the average metallicity of Collinder~261 (Sect.~\ref{sect:final}) because of the lower resolution and the inconsistencies with the other three studies. \subsection{Other cases with high dispersion or large discrepancies} \label{sect:other} \paragraph{Melotte 20:} The highest dispersion in the high-resolution sample occurs for Melotte~20 ($\alpha$~Per cluster), studied by \mbox{\citet[][see Table~\ref{t:mean_paper}]{1996AJ....111..424G}} -- 0.26~dex for three stars, one supergiant and two dwarfs. For one of the dwarf stars (HE~490), their result is close to the only other high-resolution study for four dwarf stars in this cluster \mbox{\citep{1990ApJ...351..467B}}, and to metallicity estimates from lower-resolution spectroscopy and photometry. The metallicity of the second dwarf star (HE~767) is about 0.2~dex higher. \mbox{\citet{1996AJ....111..424G}} discussed several possible explanations for this discrepancy, but did not find supporting evidence for any of them and concluded that the problem remains unresolved. The metallicity of the supergiant ($\alpha$~Per, $\log g$=1.2) is about 0.3~dex lower than that of HE~490. The authors ascribed this discrepancy to non-LTE effects, based on a contemporary estimation of these effects for metal-poor stars with similar $T_{\rm eff}$, but slightly larger $\log g$\ \mbox{\citep{1996ApJS..103..183L}}. We note that the \mbox{\citet{1996AJ....111..424G}} analysis used stellar-atmosphere models obtained from R.~Kurucz in 1992, which assume plane-parallel geometry. However, the extended atmospheres of supergiants are more accurately described by spherical geometry. According to \mbox{\citet{Heit:06}}, the inappropriate geometry leads to an overestimation of Fe abundances by $\approx$0.05~dex for the stellar parameters and Fe lines used by \mbox{\citet{1996AJ....111..424G}}, that is, in the opposite direction of the non-LTE effect. \paragraph{IC~4651:} This cluster appears in four high-resolution studies (no. 16 in Table~\ref{t:mean_paper}). In the study by \mbox{\citet{2004AaA...422..951C}}, the large dispersion of 0.19~dex obtained for the average metallicity of five stars agrees with the individual errors quoted by the authors. However, the large dispersion is due to one star, a cool giant near the RGB tip (IC~4651 56, $\log g$=0.3). The metallicity of this star is about 0.4~dex lower than that of the other four stars, which are red clump stars. Again, the authors ascribed the discrepancy to non-LTE effects, supported by the large difference between spectroscopic and photometric (evolutionary) gravities determined for the RGB tip star. For this type of star, geometry only affects abundances derived from Fe~II lines, which may be underestimated by $\approx$0.05~dex \mbox{\citep{Heit:06}}. The mean cluster metallicity quoted by the authors (0.11$\pm$0.01~dex) is based on the four red clump stars. It has the smallest dispersion of the four studies and agrees perfectly with the others. \paragraph{Collinder~121:} Only one star in this cluster, the supergiant HD~50877, has been studied by two authors. \mbox{\citet{1998AaA...338..623M}} quoted [Fe/H]=+0.25, and \mbox{\citet{2007AaA...475.1003H}} obtained [Fe/H]=$-$0.32. They also quoted substantially different $T_{\rm eff}$\ and $\log g$\ values (3200~K/0.0 and 3900~K/0.65, respectively). As mentioned above, non-LTE effects are suspected to occur in the atmospheres of supergiants, and therefore these stars are poorly suited to estimate the metallicity of an OC. \paragraph{NGC~2141:} Two different stars were studied in three publications (one each by \mbox{\citealt{2005AJ....130..597Y}} and \mbox{\citealt{2009AJ....137.4753J}}, and both by \mbox{\citealt{2013AJ....145..107J}}). Both stars are bright giants ($\log g$=1.2). The first two works resulted in rather different metallicities ($-0.18\pm0.15$~dex and +0.00$\pm$0.16~dex, respectively). Although these values agree within the quoted errors, the large errors and the large discrepancy point to considerable uncertainties inherent in the atmospheric modelling of these stars. \mbox{\citet[][Sect. 5.1.3]{2009AJ....137.4753J}} investigated this problem in more detail. They obtained the spectrum used by \mbox{\citet{2005AJ....130..597Y}} taken with the same instrument and setup, and two more spectra of the same stars with higher resolution, but lower S/N used by \mbox{\citet{2004PhDT........36B}}, who had derived an even lower metallicity for the two stars. Analysing these spectra from scratch in the same way as their own observations, they arrived at a consistent metallicity close to solar. They demonstrated that different values for the microturbulence parameter are a major source for the abundance differences. Systematic differences in measured equivalent widths and different sets of spectral lines may contribute as well. \mbox{\citet{2013AJ....145..107J}} reanalysed the same spectra as were used in \mbox{\citet{2005AJ....130..597Y}} and \mbox{\citet{2009AJ....137.4753J}} with a new version of the radiative transfer code and more recent, spherical stellar atmosphere models than \mbox{\citet{2009AJ....137.4753J}}. They obtained consistent metallicities of $-0.09\pm0.18$~dex for the two stars, and for the Fe~I and the Fe~II line lists, that is, in between the previous discrepant determinations. \begin{table*} \caption{Abundance analyses of NGC~2632 stars (Praesepe) -- mean [Fe/H] values and standard deviation, spectrum quality, approximate number of Fe~I and Fe~II lines used, and corresponding wavelength ranges.} \label{t:praesepe} \centering \begin{tabular}{llrrrrrr} \hline\hline \noalign{\smallskip} & & & & \multicolumn{2}{c}{Fe~I} & \multicolumn{2}{c}{Fe~II} \\ [Fe/H] & Reference & R & S/N & n & $\lambda$ [\AA] & n & $\lambda$ [\AA] \\ \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{7}{l}{dwarfs} \\ $+$0.04 $\pm$ 0.03 & \citet{1992ApJ...387..170F} & 28k & 100--200 & 7/14 & 7015--7205 & & \\ $+$0.11 $\pm$ 0.00 & \citet{2007ApJ...655..233A} & 55k & $>$100 & 15 & 5300--6200 & 9 & 4490--6150 \\ $+$0.27 $\pm$ 0.04 & \citet{2008AaA...489..403P} & 100k & $\approx$80 & 60 & 4800--6800 & 10 & 4800--6800 \\ \multicolumn{7}{l}{giants} \\ $+$0.15 $\pm$ 0.05 & \citet{2011AaA...535A..30C} & 30k & 150--215 & 177 & 5055--8945 & 9 & 5991--7711 \\ \noalign{\smallskip} \hline\hline \end{tabular} \end{table*} \paragraph{NGC~2632:} For Praesepe, metallicity determinations are available for 15 different stars (three giants and twelve dwarfs) in four publications (no. 58 in Table~\ref{t:mean_paper}). In each paper, a different set of stars was analysed, which makes a direct comparison impossible. The results in each of the three publications using dwarf stars show small dispersions for their respective stellar samples (at most 0.04~dex for two to six stars), but the mean metallicities are significantly different from each other (by 0.07 to 0.23~dex). The discrepancies might be due, apart from the different objects studied, to the different spectroscopic material, and the different selection of Fe lines used. In particular, the wavelength regions and spectral resolutions vary between the three works, while the S/Ns are similar (see Table~\ref{t:praesepe}). It is curious that the study based on the spectra with the highest resolution quotes the most discrepant value. The metallicity derived by \mbox{\citet{2007ApJ...655..233A}} for dwarf stars agrees best with the value obtained for giants by \mbox{\citet{2011AaA...535A..30C}}. \subsection{Importance of spectrum quality for mean cluster metallicity} \label{sect:mean} \begin{figure*} \centering \includegraphics[width=\textwidth,trim=10 40 10 50,clip]{clusters-compare.pdf} \caption{ Mean cluster metallicity determined by different authors from spectra with different resolution versus reference [Fe/H]. High resolution: $R\ge25000$, medium resolution: $R\ge13000$, low resolution: $R\approx 1000$, spectroscopic indices. All spectra have high S/N. The reference [Fe/H] values are given in Table~\ref{tab:mean1}. The solid line is the one-to-one relation. The bars below the legend represent the average standard deviations of the cluster means for high, medium, and low resolution from left to right, respectively. Metallicities and references for medium- and low-resolution studies are given in Table~\ref{tab:mean1}. Metallicities and references for high-resolution studies are given in Table~\ref{t:mean_paper}.} \label{fig:mean1} \end{figure*} \begin{table*} \caption{Metallicities and references for medium- and low-resolution studies for clusters shown in Fig.~\ref{fig:mean1}.} \label{tab:mean1} \centering \begin{tabular}{lrrrrrll} \hline\hline \noalign{\smallskip} Cluster ID & ref. [Fe/H] & mean [Fe/H] & std. dev. & resolution & \# & star type & reference \\ \noalign{\smallskip} \hline \noalign{\smallskip} \input{table_mean1} \noalign{\smallskip} \hline\hline \noalign{\smallskip} \end{tabular} \tablefoot{ The second column lists the reference metallicity. The reference [Fe/H] values are the weighted means of high-resolution determinations for all stars in each cluster (cf. Table~\ref{tab:appendix_members}; IC~4756, NGC~1245, and NGC~6633, as well as NGC~2158, NGC~2355, and NGC~2420 were shifted for better visibility) or the medium-resolution determination for NGC~2204. } \end{table*} We now turn from comparing determinations for individual stars to a comparison of mean cluster metallicities based on low-, medium- and high-resolution spectra, \emph{all with high S/N}. We started from a list of 33 clusters studied at medium resolution ($13000\le R<25000$) and searched for low- or high-resolution determinations for the same clusters. In total, we found multiple-resolution determinations for 23 clusters. Nineteen of these have determinations from low-resolution spectra (spectroscopic indices) in four publications. The medium-resolution determinations are taken from ten publications. The low- and medium-resolution references are listed in Table~\ref{tab:mean1}. The cluster Tombaugh~2 is not included in Table~\ref{tab:mean1}. It is part of the \mbox{\citet{2002AJ....124.2693F}} sample, and has been studied at medium resolution by \mbox{\citet{2008MNRAS.391...39F}} and \mbox{\citet{2010AaA...509A.102V}}. These two authors analysed 14 and 13 radial velocity members, respectively, using the same instrument, but different wavelength regions. \mbox{\citet{2008MNRAS.391...39F}} obtained two different mean metallicities (about solar and $-0.3$~dex) for the two halves of their sample, while \mbox{\citet{2010AaA...509A.102V}} arrived at a consistent metallicity of about $-0.3$~dex for all of their stars. The possible reasons for the different results are discussed at length in \mbox{\citet{2010AaA...509A.102V}}. We add that of the five stars in common between the two studies, one has the same atmospheric parameters and metallicity, while three have different parameters and metallicities, and one has different parameters but the same metallicity. We note in particular that the metallicity difference increases with the difference in microturbulence parameter between the two works. Two additional radial-velocity members of Tombaugh~2 have been analysed by \mbox{\citet{1996AJ....112.1551B}} at high resolution. However, no atmospheric parameters are given, and for the metallicity only the absolute value is quoted. Adopting the reference solar abundance for Fe~I given in \mbox{\citet{1992AJ....104.1818B}}, who used the same analysis methods, the metallicities of the two stars would be $-0.6$ and $-0.8$~dex. We regard the status of this cluster as inconclusive. All but one cluster are included in the high-resolution sample (31 publications, see Table~\ref{t:mean_paper}). Most of them have several different determinations that agree very well with each other. In Fig.~\ref{fig:mean1}, we plot the mean cluster metallicities from each publication as a function of reference [Fe/H]\ (between $-0.25$ and +0.15~dex), as given in Table~\ref{tab:mean1}. Two clusters at the two metallicity extremes lie outside of this range (NGC~2243 and NGC~6791). The metallicities based on indices by \mbox{\citet{2002AJ....124.2693F}} are lower by up to 0.5~dex than the medium- and high-resolution metallicities (the differences increase with metallicity). This discrepancy has previously been noted for individual stars in representative clusters in Sects.~\ref{sect:m67} to \ref{sect:metalpoor}, and by \mbox{\citet{2010AaA...511A..56P}} for 28 OCs in the \mbox{\citet{2002AJ....124.2693F}} sample, which also have $R>$15000 determinations \mbox{\citep[see also][Sect.~6.1.3]{2010AJ....139.1942F}}. A systematic difference of about 0.2~dex is also seen between the \mbox{\citet{2002AJ....124.2693F}} OC metallicities and metallicities compiled for Cepheids at similar Galactocentric distances \mbox{\citep[see Fig.~4 in][]{2009AaA...504...81P}}. On the other hand, for NGC~6705 at [Fe/H]=0.06~dex, both the \mbox{\citet{1993PASP..105.1253T}} and \mbox{\citet{2005AJ....130.1916M}} indices result in higher metallicities than at high resolution, while for NGC~1245 (0.02~dex), NGC~2682 (0.03~dex), and NGC~6791 (0.42~dex), the determinations by \mbox{\citet{2005AJ....130.1916M}} or \mbox{\citet{2003PASP..115...96W}} are about 0.1~dex lower than at high resolution. The medium-resolution results agree in general very well with the high-resolution ones, except for IC~4756, NGC~188, NGC~2158, NGC~2420, and NGC~6705. For IC~4756 at 0.018~dex the work of \mbox{\citet{2007AJ....134.1216J}} gives a 0.15~dex lower metallicity value than the other medium-resolution publication and the three high-resolution publications. For NGC~188, NGC~2158, and NGC~2420 the medium-resolution values by \mbox{\citet{2011AJ....142...59J}} are 0.1 to 0.2~dex lower than the high-resolution determinations. For NGC~6705 at 0.06~dex, the medium-resolution value by \mbox{\citet{2000PASP..112.1081G}} is 0.15~dex higher than two of the high-resolution values, while the high-resolution value of \mbox{\citet{2012AaA...538A.151S}} is lower than the other two by a similar amount, and all mean [Fe/H]\ values agree within the standard deviations. NGC~2112 (reference [Fe/H]=0.14~dex) has two discrepant high-resolution determinations. This old anticentre OC is useful for determining the radial metallicity gradient of the Galactic disk. \mbox{\citet{1996AJ....112.1551B}} found it to be mildly deficient, [Fe/H]=$-0.16\pm0.25$~dex (only one star was observed at high resolution), while \mbox{\citet{2008MNRAS.386.1625C}} found it to be slightly supersolar, [Fe/H]=+0.14~dex, close to the Hyades value. \mbox{\citet{1996AJ....112.1551B}} observed a cool red giant, while \mbox{\citet{2008MNRAS.386.1625C}} observed one dwarf, one clump giant, and one F~giant with a small dispersion of 0.03~dex. NGC~6633 (reference [Fe/H]=0.026~dex) has four discrepant high-resolution determinations. The closest [Fe/H]\ values, +0.03~dex for six measurements of three stars, and +0.11~dex for one star, are obtained by \mbox{\citet{2009AaA...493..309S}} and \mbox{\citet{2009AaA...502..267S}}, respectively. \mbox{\citet{2005ApJS..159..141V}} observed the same star as \mbox{\citet{2009AaA...502..267S}}, NGC~6633 100 = HD~170174, which is also in common with \mbox{\citet{2009AaA...493..309S}}, but derived the significantly higher value of +0.35~dex. The temperature adopted by \mbox{\citet{2005ApJS..159..141V}} is substantially hotter ($T_{\rm eff}$=5245~K, while \mbox{\citealt{2009AaA...502..267S}} adopted $T_{\rm eff}$=5015~K, and \mbox{\citealt{2009AaA...493..309S}} $T_{\rm eff}$=4980~K). We conclude that the different temperature scale employed by \mbox{\mbox{\citet{2005ApJS..159..141V}}} is the cause for the discrepant metallicity determination. Therefore, we discarded this determination when we calculated the final mean cluster metallicity (Sect.~\ref{sect:final}). The fourth metallicity determination by \mbox{\citet{2005MNRAS.363L..81A}} is significantly lower than the others ($-0.21$~dex) and is based on two F~dwarfs. The most metal-poor one (JEF~1), with [Fe/H]=$-$0.31~dex is a moderate rotator with $v$sin$i$=19~km~s$^{-1}$ and a high $T_{\rm eff}$ = 6870~K, while the cooler F~dwarf (HJT~1251 = JEF~16) has a value of [Fe/H]=$-$0.15~dex, which agrees better with the other determinations. In summary, at resolutions higher than $\sim$10000, cluster metallicities may have small or large dispersions regardless of the resolution value. Other factors such as analysis method, temperature scale, or properties of the sample stars seem to play a larger role for the reliability of the metallicity than spectral resolution. Finally, we assessed the impact of S/N by comparing mean cluster metallicities determined by different authors from spectra with high ($>50$) or low S/N, \emph{all of them with high resolution}. For six clusters we found metallicity determinations from both high-S/N spectra (21 publications) and low-S/N spectra (8 publications). The clusters, metallicities, and references for low-S/N determinations are given in Table~\ref{tab:mean2}. Metallicities and references for high-S/N determinations can be found in Table~\ref{t:mean_paper}. For NGC~6475, the low-S/N determination results in a higher metallicity than the high-S/N determination, although the values agree within two standard deviations. For NGC~188, NGC~2243, and NGC~6791, the one or two low-S/N metallicities are lower than the one or two high-S/N ones for each cluster, but they agree within one standard deviation. For Berkeley~29 and NGC~2682, the one or two low-S/N determinations available for each cluster lie within the range of the high-S/N mean metallicities. Thus, spectra with a S/N as low as 20 might be sufficient to determine reliable cluster metallicities. However, the limited number of cases for comparison does not enable us to draw a firm conclusion. \begin{table*} \caption{Metallicities and references for low-S/N studies for clusters discussed in Sect.~\ref{sect:mean}, last paragraph.} \label{tab:mean2} \centering \begin{tabular}{lrrrrrll} \hline\hline \noalign{\smallskip} Cluster ID & mean [Fe/H] & std. dev. & resolution & S/N & \# & star type & reference \\ \noalign{\smallskip} \hline \noalign{\smallskip} Berkeley~29&$-$0.32&0.03&45000&25--50&5&giant& \citet{2008AaA...488..943S} \\ NGC~188&$-$0.12&0.16&$\approx$30000&$\approx$45&7&dwarf+subgiant&\citet{1990AJ....100..710H} \\ NGC~188&0.01&0.08&35000--57000&20--35&5&dwarf& \citet{2003AaA...399..133R} \\ NGC 2243 & $-$0.54 & 0.10 & 19300/28800\tablefootmark{$\dagger$} & 30--40 & 76 & dwarf+giant & \citet{2013AaA...552A.136F} \\ NGC~2682 (M67)&$-$0.04&0.01&$\approx$30000&30--45&2&dwarf& \citet{1991AJ....102.1070H} \\ NGC~2682 (M67)&0.04&0.01&28400&30--45&2&dwarf& \citet{1992ApJ...387..170F} \\ NGC~6475&0.11&0.03&40000&20--30&10&dwarf& \citet{1997MNRAS.292..252J} \\ NGC~6791&0.30&0.08&45000&40&2&turn-off& \citet{2009AJ....137.4949B} \\ \noalign{\smallskip} \hline\hline \end{tabular} \tablefoottext{$\dagger$}{Higher $R$ between 638 and 663~nm, lower $R$ between 660 and 696~nm.} \end{table*} \section{Final high-resolution sample} \label{sect:final} To construct the final list of reference spectroscopic cluster metallicities, we retained the restrictions in spectral resolution and S/N, even thoug lower-quality spectra may also provide reliable results. For the high-quality sample we compiled the metallicity values for individual stars from the complete literature, while the lower-quality sample compiled by us is incomplete and comprises only average cluster metallicities. Following the considerations in Sect.~\ref{sect:assessment}, we decided to restrict the temperature and gravity range of the determinations used to compute the average metallicity of each OC. We adopted the $T_{\rm eff}$\ range 4400--6500~K with $\log g$\ $\geq$ 2.0 to eliminate any rapidly rotating hot dwarfs or stars with chemical peculiarities, and bright giants possibly affected by non-LTE effects. When two or more determinations were available for the same star, with parameters on both sides of the limiting value, the star was included, but only determinations that met the constraints were included in the average metallicity (four dwarfs and ten giants in 12 clusters). For dwarfs, the lower $T_{\rm eff}$\ limit eliminates three stars, without any critical consequence on the corresponding OCs (the metallicity of these cool dwarfs is very uncertain anyway). The $T_{\rm eff}$\ restriction is important to keep in mind for future metallicity calibrations, that is, the colour interval where they are valid. The $T_{\rm eff}$\ restriction removed several clusters from the sample, namely Collinder~121 and NGC~2141 (see Sect.~\ref{sect:other}), as well as Pfleiderer~4, Trumpler~2, Trumpler~20, NGC~1883, NGC~2158, and the old OC NGC~2243. The latter is supposedly one of the most metal-deficient OCs according to the value of [Fe/H]=$-$0.48~dex determined by \mbox{\citet{1994AaA...283..911G}} for two bright giants. Our final list includes 458 stars in 78 OCs, with 641 metallicity determinations corresponding to 86 papers. The measurements that are included in the final high-resolution sample are identified in Table~\ref{tab:appendix_members}. We computed the average metallicity from all determinations for each cluster, weighted by the inverse square of the individual errors quoted by the authors. For a significant number of determinations the authors did not quote any uncertainties. In these cases, we assumed an error of 0.1~dex. This approach might seem problematic because authors did not compute the errors in the same way -- some quoted only internal errors while others took some external sources of uncertainty into account. Therefore we tested several additional approaches to compute the average metallicity: a) setting the lower limit of the errors used for weighting to 0.1~dex (that is, the typical uncertainty due to the uncertainties in stellar parameters, see Sect.~\ref{sect:errors}), and b) using equal weights for all metallicity values. The average metallicity values do not change by more than 0.03~dex for both approaches (a and b) for most clusters. They change by less than 0.1~dex for five and eight clusters (approaches a and b, respectively), with no systematic trend with metallicity. The standard deviations become larger (by more than 0.02~dex) for three clusters, and smaller for one cluster. We tested another approach for computing the average cluster metallicity -- a two-step mean, where we first calculated the mean metallicity for each star (we recall that a significant number of stars have been analysed by several authors), and then the mean for each cluster. Using this approach, only six clusters deviate from the nominal approach (by 0.02~dex), which shows that a straight one-step average is justified. The resulting mean metallicities (one-step average with weights using the original errors) are plotted in Fig.~\ref{f:mean_feh} and listed in Table~\ref{t:mean_feh}. The figure can be directly compared with Fig.~\ref{f:mean_paper}. For several clusters the standard deviations shown in Fig.~\ref{f:mean_feh} are smaller than the dispersion between different determinations for one and the same cluster seen in Fig.~\ref{f:mean_paper}. This is most evident for no. 22 -- Melotte~25 (the Hyades), and no. 74 -- NGC~6633. For the Hyades, the largest number of metallicity determinations are available (129 in Table~\ref{t:mean_paper}). After discarding 30\% probably unreliable determinations, we are still left with a significant number (92 in Table~\ref{t:mean_feh}), resulting in good statistics. The case of NGC~6633 is discussed in Sect.~\ref{sect:mean}. Other clusters with improved dispersions are no. 6 -- Berkeley~29, no. 12 -- Collinder~261, no. 16 -- IC~4651, no. 58 -- NGC~2632, no. 60 -- NGC~2682 (M67), no. 63 -- NGC~3680, and no. 80 -- NGC~752. In Fig.~\ref{f:mean_before_after} we compare the average metallicities for the final sample with the weighted average metallicities per cluster for the starting sample, which includes all determinations for all highly probable members without restrictions on atmospheric parameters. The mean difference (final sample minus all members) is $0.01\pm0.04$~dex, with minimum and maximum differences of $-0.10$ and +0.18~dex, respectively. This shows that the metallicities do not change significantly by restricting the $T_{\rm eff}$\ and $\log g$\ ranges, except for a few clusters with large standard deviations. The cluster changing by the largest amount (+0.18~dex) is Melotte~20 ($\alpha$ Per cluster), discussed in Sect.~\ref{sect:other}. \begin{figure*}[t] \includegraphics[width=\textwidth,trim=0 180 0 280,clip]{mean_feh.pdf} \caption{Weighted average metallicity of OCs as listed in Table~\ref{t:mean_feh}. The error bars represent the standard deviation of the mean.} \label{f:mean_feh} \end{figure*} \begin{table*} \caption{Weighted average metallicity, and standard deviation, of OCs after restricting the $T_{\rm eff}$\ and $\log g$\ ranges, with same running number for OCs as in Table~\ref{t:mean_paper}. See also Fig.~\ref{f:mean_feh}.} \label{t:mean_feh} \begin{tabular}{rllrr rllrr} \hline\hline \noalign{\smallskip} OC no. & Name & Mean [Fe/H] & \# & $n$ & OC no. & Name & Mean [Fe/H] & \# & $n$ \cr \hline \noalign{\smallskip} \input{table_mean_feh} \noalign{\smallskip} \hline\hline \noalign{\smallskip} \end{tabular} \tablefoot{ The column headed ``\#'' gives the number of metallicity determinations, and the column headed ``$n$'' the number of individual member stars. } \end{table*} At the metal-rich end of the metallicity distribution of the final sample, we find three OCs with extreme values: NGC~6253 ([Fe/H]$\approx+0.3$~dex; \mbox{\citealt{2007AaA...473..129C}}; \mbox{\citealt{2007AaA...465..185S}}; \mbox{\citealt{2012MNRAS.423.3039M}}), NGC~6583 \mbox{\citep[[Fe/H]$\approx+0.4$~dex;][]{2010AaA...523A..11M}}, and NGC~6791 \citep[[Fe/H]$\approx+0.4$~dex;][]{2006ApJ...642..462G,2012ApJ...756L..40G}. At the metal-poor end, there are eight OCs with metallicities of about $-$0.3~dex and below: Berkeley~18, Berkeley~20, Berkeley~29, Berkeley~32, Melotte~66, Melotte~71, NGC~2266, and Saurer~1. The metallicities of three of these clusters rely on a substantial number of stars: 11 stars in Berkeley~32, leading to [Fe/H]=$-$0.30 $\pm$0.06 (\mbox{\citealt{2006AaA...458..121S}}; \mbox{\citealt{2011AaA...535A..30C}}; \mbox{\citealt{2012AJ....144...95Y}}; \mbox{\citealt{2013AJ....145..107J}}), six stars in Melotte~66, leading to [Fe/H]=$-$0.32 $\pm$0.03 \mbox{\citep{2008AaA...488..943S}}, and four stars in Berkeley~29, leading to [Fe/H]=$-$0.36 $\pm$0.07 \mbox{\citep{2004AJ....128.1676C,2008AaA...488..943S}}. On the contrary, the metallicity of Melotte~71 ([Fe/H]=$-$0.27) is based on only one very faint star, PJ~127, $V\simeq17$, which has $T_{\rm eff}$=4610~K, $\log g$=2.16, and an uncertainty on the metallicity of 0.24 \mbox{\citep{1996AJ....112.1551B}}. NGC~2266 is the cluster with the lowest metallicity, [Fe/H]=$-$0.44, which is also based on only one star, although a brighter one ($V\simeq11$) and with a lower uncertainty on the metallicity of 0.05 \mbox{\citep{2013MNRAS.431.3338R}}. It is worth mentioning that the authors argued that this cluster may be part of the thick disk, based on its space motions. The low metallicity of Berkeley~18 is also based on one faint star, but two measurements (star K 1163 with $V\simeq16$, see Sect.~\ref{sect:metalpoor}). The remaining two of the most-metal-poor clusters (Berkeley~20 and Saurer~1) have metallicities relying on two stars. We have eleven more OCs in the final sample with metallicities relying on only one star, most of them with subsolar metallicities. \begin{figure} \includegraphics[width=\columnwidth,trim=135 40 150 50,clip]{mean_feh_before_after.pdf} \caption{Weighted average metallicity of OCs as listed in Table~\ref{t:mean_feh} versus weighted average metallicity of OCs, including all member stars from all publications in Table~\ref{t:mean_paper}. The error bars represent the standard deviation of the mean. Clusters to the left of the green vertical line do not appear in the final sample.} \label{f:mean_before_after} \end{figure} For the three clusters with the largest number of determinations ($>30$), we can separate the stars into giants ($\log g$ $\le$ 3.0) and dwarfs (cf. Sect.~\ref{sect:m67} for M67). The weighted mean metallicities for the separated samples and the number of determinations are listed in Table~\ref{tab:dwarfsgiants}. The determinations for the Hyades (Melotte~25) are mainly for dwarfs. For M67 (NGC~2682) and IC~4651, the determination numbers are more similar for giants and dwarfs. We found no indication for a metallicity difference between the dwarfs and giants for any of these clusters. \begin{figure} \includegraphics[width=\columnwidth,trim=135 40 150 50,clip]{mean_feh_final_phot.pdf} \caption{Comparison of mean metallicities listed in Table~\ref{t:mean_feh} and those obtained from photometry in Paper~I for 62 OCs in common (two clusters lie outside the axis ranges -- no. 35, NGC~2112, and no. 4, Berkeley~21).} \label{f:HR-phot} \end{figure} We compare in Fig.~\ref{f:HR-phot} the mean metallicities determined from high-resolution spectroscopy and that obtained from photometry (Paper~I). There are 62 clusters in common, and the mean difference between photometric and spectroscopic values is $-0.10$~dex with a median of $-0.08$ and a standard deviation of 0.23. The two most extreme cases are NGC~2112 and Berkeley~21, with differences of $-1.44$ and $-0.78$~dex, respectively. These are also two of the most metal-poor clusters in the photometric sample ([Fe/H]$_{\rm phot}= -1.30$ and $-0.96$~dex, respectively). There are no other clusters with photometric metallicities below $-0.5$~dex in the common sample. We conclude that the photometric metallicity scale is more metal-poor than the spectroscopic one. Combining our two samples of clusters with photometric and spectroscopic metallicities would result in a total number of 204 clusters with known metallicity. However, the differences in metallicity found for the two samples and the large intrinsic dispersion of the photometric determinations requires a calibration of the photometric sample to the spectroscopic metallicity scale. This process and the discussion of the full sample will be presented in a forthcoming article. \begin{table} \caption{Weighted mean metallicities and standard deviations for giant ($\log g$ $\le$ 3.0) and dwarf samples in three clusters.} \label{tab:dwarfsgiants} \centering \begin{tabular}{lrr} \hline\hline \noalign{\smallskip} Name & Mean [Fe/H] & \# \\ \noalign{\smallskip} \hline \noalign{\smallskip} Melotte~25 giants & 0.12 $\pm$ 0.04 & 16 \\ Melotte~25 dwarfs & 0.13 $\pm$ 0.06 & 76 \\ \noalign{\smallskip} \hline \noalign{\smallskip} NGC~2682 giants & $-$0.05 $\pm$ 0.05 & 23 \\ NGC~2682 dwarfs & 0.01 $\pm$ 0.06 & 29 \\ \noalign{\smallskip} \hline \noalign{\smallskip} IC~4651 giants & 0.12 $\pm$ 0.05 & 15 \\ IC~4651 dwarfs & 0.12 $\pm$ 0.04 & 20 \\ \noalign{\smallskip} \hline\hline \end{tabular} \end{table} \section{Discussion} \label{sect:discussion} \subsection{Comparison with other metallicity studies} \label{sect:discusscompare} \begin{figure} \includegraphics[width=\columnwidth,trim=135 40 150 50,clip]{mean_feh_final_NP13.pdf} \caption{Weighted average metallicity of OCs as listed in Table~\ref{t:mean_feh} versus metallicities determined by \citet{2013arXiv1307.2094N}.} \label{f:final_NP13} \end{figure} \mbox{\citet{2013arXiv1307.2094N}} determined metallicities of 58 OCs from photometry, using an isochrone-fitting method developed by \mbox{\citet{2010AaA...514A..81P}}. In Fig.~\ref{f:final_NP13} we compare the mean metallicities of the final high-resolution sample (Table~\ref{t:mean_feh}) with the metallicities of \mbox{\citet{2013arXiv1307.2094N}}. We have 23 clusters in common with their sample, all of which have metallicities between $-0.1$ and +0.2~dex. The mean difference between the photometric and spectroscopic values is $-0.02\pm~0.05$~dex, with a maximum absolute difference of 0.09~dex. We confirm the conclusion of \mbox{\citet{2013arXiv1307.2094N}} that their tool is useful for estimating OC metallicities from photometry. \begin{figure} \includegraphics[width=\columnwidth,trim=50 40 60 60,clip]{met_hist.pdf} \caption{Distribution of metallicities for the final spectroscopic sample (red line), the photometric sample (Paper I, blue line), and the sample of \citet[grey full histogram]{2009IAUS..254P..15C}.} \label{f:methistogram} \end{figure} A recent compilation of data, including metallicities, for a large number of OCs is the catalogue of \mbox{\citet{2003AJ....125.1397C}}. A preliminary update is presented in \mbox{\citet{2009IAUS..254P..15C}}, which contains 144 clusters with metallicity, distance, and age values. However, the metallicities are a mixture of photometric and spectroscopic determinations. Fig.~\ref{f:methistogram} shows the histogram of cluster metallicities from \mbox{\citet{2009IAUS..254P..15C}} together with those of our spectroscopic and photometric samples. The figure indicates that the peak at solar metallicity in the histogram of \mbox{\citet{2009IAUS..254P..15C}} may be dominated by spectroscopic metallicities, while the metal-poor tail is mainly due to photometric metallicities. \begin{figure*} \includegraphics[width=\textwidth,trim=0 20 0 0,clip]{MagriniHeiterCommon.pdf} \caption{Comparison of mean metallicities listed in Table~\ref{t:mean_feh} and those compiled by \citet{2010AaA...523A..11M}, for 49 OCs in common, as a function of Galactocentric distance. The points represent [Fe/H](Magrini) minus [Fe/H](this work). The error bars represent the standard deviations of the cluster means from both lists, added in quadrature. This can be directly compared with Fig.~9 of \citet{2010AaA...523A..11M}.} \label{f:MagriniHeiterCommon} \end{figure*} In Fig.~\ref{f:MagriniHeiterCommon} we compare the mean metallicities of our final high-resolution sample with those compiled by \mbox{\citet[][see Sect.~\ref{sect:introduction}]{2010AaA...523A..11M}}. The mean difference for all clusters in common is 0.02~dex (median 0.00~dex), with a standard deviation of 0.05~dex. However, the dispersion varies with $R_{\rm GC}$, and the lowest and highest values of the deviations are $-0.10$ and +0.16~dex, respectively. The scatter seen in this figure should be added to Fig.~9 of \mbox{\citet{2010AaA...523A..11M}}, representing the uncertainty originating from the way of combining cluster abundances from different authors. For example, the scatter is 0.06~dex for $7.5 \lesssim R_{\rm GC}\lesssim 9.5$~kpc, adding to the uncertainty of the inner disk gradient determined by \mbox{\citet{2010AaA...523A..11M}}. We have 64 clusters in common with the sample of \mbox{\citet[][see Sect.~\ref{sect:introduction}]{2011AaA...535A..30C}}. We do not show a detailed comparison of metallicities, since their sample is very similar to ours, and we included all of their references that meet our constraints. The differences are that their metallicities are based on published cluster means, not on individual stars, and that they set a lower limit on $R$ (15000) than we do, and no limit on S/N (spectra with S/N as low as 5 are included). Ultimately, the OC metallicities should be combined with other tracers of metallicity in the Galaxy, such as HII regions, B-type stars, planetary nebulae, and Cepheids. However, such combined samples will have to deal with differences in metallicity scale, not only for the same type of objects due to different methods, but also between different types of objects. A large sample of Cepheids has been used by \mbox{\citet{2009AaA...504...81P}} to study the metallicity gradient of the Galactic disk. They compared the metallicities of their sample as a function of Galactocentric distance with several samples of OCs. They found significant differences in derived metallicity gradients, which were partly attributed to different Galactocentric distributions of these tracers. It is worth noting that \mbox{\citet{2009AaA...504...81P}} combined spectroscopic metallicities for over 200 Cepheids with photometric ones for about 60. The mean difference between photometric and spectroscopic metallicities ($-0.03$~dex) and the intrinsic dispersion (0.15~dex) are smaller than for our two samples. This could partly be due to the more homogeneous photometric metallicities determined by \mbox{\citet{2009AaA...504...81P}} from one set of photometric bands (Walraven and K-band) and a metallicity calibration based on recent stellar evolution models. However, a systematic difference between two of their sources for spectroscopic metallicities is apparent in their Fig. 3. \subsection{Application to Galactic structure} \label{sect:galactic} \begin{figure} \includegraphics[width=\columnwidth,trim=135 35 180 50,clip]{xymet_OC12.pdf} \caption{Distributions of the OCs from Paper~I (diamonds), the OCs in the spectroscopic sample (coloured circles), and the OCs in the \citet[][version 3.2]{2002A&A...389..871D} catalogue (grey circles) projected onto the Galactic plane, in a linear coordinate system where $X$ increases from the Sun towards the Galactic centre, and $Y$ increases in the direction of Galactic rotation at the location of the Sun. The Galactic centre is located at ($X,Y$) = (0,0). Metallicity is represented by the colour scale shown in the bar to the right. Solid lines indicate distances from the Galactic centre of 6, 8, and 10~kpc. Three clusters in the spectroscopic sample and one in the photometric sample lie outside the shown $X$ range.} \label{f:xydistribution} \end{figure} In this section we use our final sample of OCs with spectroscopic metallicities for a preliminary investigation of the distribution of metals in the Galactic disk. In Fig.~\ref{f:xydistribution} we show the spatial distribution of OCs projected onto the Galactic plane. Together with our final spectroscopic sample, we also include the photometric sample from Paper~I and all clusters in the \mbox{\citet[][Version 3.2]{2002A&A...389..871D}} catalogue. The metallicity of each cluster in the first two samples is indicated by the colour of the symbol. We recall that the photometric and spectroscopic metallicities are not yet on the same scale. Even though a combined sample would trace the metallicity throughout the disk better than the spectroscopic one alone, gaps are evident, which should be filled by targeted observations of known clusters. We did not attempt to derive a more detailed metallicity distribution for the spectroscopic sample in the immediate solar neighbourhood ($\pm2$~kpc), as we did in Fig.~4 of Paper~I. The number of clusters in that area is simply too small (55 compared with 128 in Paper~I). However, we speculate that the distribution based on a combined calibrated sample will be more smooth than that based on the photometric sample alone. In particular, the most distinct feature of Fig.~4 in Paper~I, a dip in metallicity at $X=-8.4$ and $Y=-0.4$~kpc, will disappear. It is caused by the cluster NGC~2112, which has a spectroscopic metallicity of $+0.14\pm0.08$~dex, based on three stars in two publications, while its photometric metallicity of $-1.3\pm0.2$~dex is based on Str\"omgren colours of four stars in one publication. \begin{figure*} \includegraphics[width=\textwidth]{gradient.pdf} \caption{Cluster metallicity as a function of Galactocentric distance computed from the coordinates of the clusters and their distance from the Sun taken from the \citet[][version 3.2]{2002A&A...389..871D} catalogue. The distance of the Sun from the Galactic centre (at the right of the graph) is assumed to be 8~kpc. Open symbols are clusters with only one metallicity measurement. Also shown are predictions of several models for the Galactic chemical evolution. See Sect.~\ref{sect:galactic} for details.} \label{f:fehrgc} \end{figure*} Figure~\ref{f:fehrgc} shows the cluster metallicity from Table~\ref{t:mean_feh} as a function of Galactocentric distance $R_{\rm GC}$ compared with the predictions of several models for Galactic chemical evolution. The Galactocentric distances were computed from the coordinates of the clusters and their distance from the Sun as given in the catalogue of \mbox{\citet[][version 3.2]{2002A&A...389..871D}}. The distance of the Sun from the Galactic centre is assumed to be $R_{\rm GC,\odot}$=8~kpc. The error bars for $R_{\rm GC}$ correspond to a variation of the cluster distances by $\pm$20\%. Most clusters are located in the range $6 \lesssim R_{\rm GC,\odot} \lesssim 10$~kpc. The clusters within this range are also mostly located close to the Sun in the Galactic plane, as can be seen in Fig.~\ref{f:xydistribution}, guided by the solid lines. The metallicity distributions predicted by the following models are shown in Fig.~\ref{f:fehrgc}: \begin{itemize} \item The \emph{simple model} with instantaneous recycling described, for example, in \mbox{\citet[][eq.~8.7]{2009nceg.book.....P}}, green long-dashed line. This model allows one to derive an analytic expression for the average abundance of Fe in a stellar population: \[ \langle z(Fe) \rangle = 1+\frac{\mu \ln(\mu)}{1-\mu}, \] where the gas fraction $\mu = g/(s+g)$, where $g$ is the density of gas in the system, and $s$ is the density of matter in the form of stars. The dependence of the gas fraction and, in turn, of the metallicity on $R_{\rm GC}$ is obtained by assuming the following relations for the gas and stellar surface densities in $M_\odot$ pc$^{-2}$ as a function of Galactocentric distance in kpc: \[ g = 15.0 e^{\frac{-R_{\rm GC}}{9.9}} \] (fit to data shown in Fig.~1 of \mbox{\citealt{1993AIPC..278..267D}}, for $R_{\rm GC} > 4.5$~kpc), and \[ s = 198 e^{\frac{-R_{\rm GC}}{4.0}}. \] The latter was normalized to give a gas fraction of 0.2 at the distance of the Sun (see \mbox{\citealt{2009nceg.book.....P}}, Table~7.9). \item The \emph{extreme inflow model}, with a metallicity dependence derived by \mbox{\citet{1972NPhS..236....7L}} as: \[ \langle z(Fe) \rangle = 1+\frac{1}{s/g (e^{-s/g}-1)} \] (see \citealt{2009nceg.book.....P}, eq.~8.28), blue short-dashed line. \item The model by \mbox{\citet{1997ApJ...477..765C}} -- gradient of iron at 12~Gyr from their Table~4, but using $R_{\rm GC,\odot}$=8~kpc, grey dash-dotted line. \item The model by \mbox{\citet{2006MNRAS.366..899N}} -- present-day metallicities of stars from their Fig.~11, red solid line. \item The model by \mbox{\citet{2009MNRAS.396..203S}} -- mean metallicities of stars at the present time from their Fig.~11, black solid line. \end{itemize} Even though the first two analytical models can be regarded as two opposite extreme cases, their predicted radial dependence of metallicity is quite similar. It does not agree with the observed metallicity dependence, in particular at distances between about 10 and 15~kpc. The main feature of the two-infall model by \mbox{\citet{1997ApJ...477..765C}} is that the thick and thin disks form by accretion of extragalactic material on very different timescales (1~Gyr and 8~Gyr, respectively). The model predictions agree well with several observed properties of the Galaxy, such as the stellar metallicity distribution and the fraction of metal-poor stars in the solar neighbourhood. However, the metallicity gradient predicted by the model is much shallower than that suggested by the OC metallicities (Fig.~\ref{f:fehrgc}). In the model of \mbox{\citet{2006MNRAS.366..899N}}, the evolution of the gas infall rate is prescribed based on spherical infall theory and the current observed distribution of the total disk surface mass. For other ingredients (star formation, IMF, chemical evolution), the model uses standard prescriptions. Among the models discussed here, the radial dependence of metallicity predicted by this model agrees best with the OC metallicities, except that its metallicity is somewhat low at the solar radius. However, \mbox{\citet{2006MNRAS.366..899N}} mentioned that their model predicts a significantly steeper metallicity gradient for past times. Including information on the cluster ages might improve the agreement. Also, changing the IMF in the model from \mbox{\citet{1955ApJ...121..161S}} to \mbox{\citet{2003PASP..115..763C}} leads to a steeper gradient. The model of \mbox{\citet{2009MNRAS.396..203S}} introduces both radial gas flows and radial migration of stars, beyond the standard ingredients of chemical evolution models. This model predicts star counts as a function of various stellar parameters in agreement with observations, and it reproduces correlations between tangential velocity and abundance patterns. It also produces a thick disk alongside the thin disk within the Galaxy. Regarding the predicted metallicity as a function of $R_{\rm GC}$, the model curve does not coincide with any of the observed OC metallicities. To bring the model and observations into agreement would require a shift in metallicity of at least 0.3~dex. However, the model gradient is very similar to the observed one around the solar radius ($7\lesssim R_{\rm GC} \lesssim 12$~kpc). The steep gradient is mainly caused by the radial gas flow, as \mbox{\citet{2009MNRAS.396..203S}} showed by varying the model parameters. Their model also predicts distributions over metallicity of stars at fixed $R_{\rm GC}$ with a full-width at half maximum (FWHM) around 0.35~dex. This value agrees exactly with the FWHM of the metallicity distribution of the 26 clusters in our sample with $7\lesssim R_{\rm GC} \lesssim 9$~kpc. Although the list of discussed models is not exhaustive, we can conclude that none of the current models for Galactic chemical evolution succeeds in predicting the metallicity gradient observed for the Galactic disk based on OCs. However, the significance of the comparison is limited by the inhomogeneous distribution of clusters over distance and the fact that the sample contains clusters of different ages. Moreover, we did not take into account possible radial migrations of the clusters. We postpone a more detailed comparison to a forthcoming paper, where we will combine the photometric and spectroscopic metallicity determinations in a proper way. \section{Summary and conclusions} \label{sect:conclusions} For this article, metallicities of individual stars in OCs resulting from spectroscopy at high resolution (R$>$25000) and high signal-to-noise ratio (S/N$>$50) were exhaustively gathered from the recent literature (publication year 1990 and later). Only the most probable members in 86 OCs were considered for further analysis. Some discrepancies in metallicities of individual stars and cluster means were found for a fraction of OCs studied by several authors. The sources of these differences were analysed in relation to the quoted errors, and various aspects of the spectroscopic analyses with an impact on metallicity were discussed. Comparisons were also made with metallicities based on lower resolution spectra and spectral indices, and on lower S/N observations. We found that low-resolution estimates are in general more metal poor than higher-resolution determinations. At medium and high resolution ($R$$\ge$13000) and S/N$>$20, the temperature scale, the line list, the methodology, and the choice of the microturbulence parameter, seem to play a larger role for the reliability of the metallicities than the resolution and S/N. However, the largest contribution to the observed dispersion comes from the properties of the stars. Chemically peculiar stars and binaries must obviously be removed when averaging the metallicity determinations in an OC, but also bright giants that are possibly affected by non-LTE, and hot dwarfs that are possibly affected by rapid rotation. These considerations led us to build a clean sample of metallicity determinations of individual stars in a restricted temperature and gravity range, after rejecting some studies that appeared to be affected by systematic uncertainties. The numbers of different stars and metallicity determinations were reduced by 25\% and nearly 30\%, respectively, between the starting and the final sample. The final sample includes 458 stars with 641 metallicity determinations in 86 papers, which were used to compute the weighted average metallicity of 78 OCs. We found no difference in mean metallicities deduced from dwarfs and giants, based on three OCs with a significant number of determinations for both groups. Photometric metallicities compiled in Paper~I were found to be systematically more metal-poor by 0.11~dex than the spectroscopic ones presented here, with a standard deviation of 0.23~dex. However, recent photometric determinations by \mbox{\citet{2013arXiv1307.2094N}} agree much better. The compilation of spectroscopic metallicities by \mbox{\citet{2010AaA...523A..11M}}, which consists of the cluster metallicity from one selected publication per cluster, agrees on average with our spectroscopic metallicities for the clusters in common. However, metallicities for individual clusters deviate by up to 0.16~dex. We used our final sample to test four models that predict the radial metallicity gradient in the Galaxy. None of them was found to fully agree with the OC metallicities versus Galactocentric distance. The model by \mbox{\citet{2009MNRAS.396..203S}} has a similar slope, but shows a metallicity shift of 0.3~dex. The metallicity dispersion at the solar radius predicted by this model is similar to the one measured for our sample. This comparison shows that existing models of Galactic chemical evolution cannot reproduce the currently available sample of cluster metallicities even within the rather large observational uncertainties, and calls for further developments on the theoretical side. This work demonstrates, however, that it is crucial to enlarge the number of OCs with accurate metallicities from spectroscopic studies. Metallicity determinations are needed in particular at small and large Galactocentric distances, and in several regions of the Galactic disk that are poorly sampled. To obtain reliable cluster metallicities will require an increase in the number of individual stars studied in each cluster, a careful selection of the stars, and a homogeneous analysis of the spectra. This is one of the main aims of the Gaia-ESO Public Spectroscopic Survey, which will dramatically improve the situation of spectroscopic metallicities of OCs, with 100 clusters to be observed in the next four years. Studies of Galactic structure should take advantage of the large number of existing photometric metallicities for OCs (Paper~I). In combination with the current spectroscopic sample, a sample of more than 200 clusters with known metallicities can be constructed. A calibration of the photometric sample using the spectroscopic sample will be required to define a common metallicity scale with a small intrinsic dispersion. The details of this calibration and the impact of the combined sample will be the subject of the next paper of this series. \begin{acknowledgements} UH acknowledges support from the Swedish National Space Board (Rymdstyrelsen). We thank the Bordeaux 1 University for providing UH a one-month position in 2009 and 2010. This work was supported by the SoMoPro II Programme (3SGA5916), co-financed by the European Union and the South Moravian Region, grant GA \v{C}R 7AMB12AT003, and the financial contributions of the Austrian Agency for International Cooperation in Education and Research (CZ-10/2012). We thank the referee, Elena Pancino, for valuable comments. This research has made use of the WEBDA database, operated at the Department of Theoretical Physics and Astrophysics of the Masaryk University, and of the SIMBAD database, operated at CDS, Strasbourg, France. \end{acknowledgements} \bibliographystyle{aa}
\section{Introduction} It is well known that orthogonal polynomials on the real line are characterized by their three-term recurrences \cite{Chihara, Gautschi, Szego}. Specifically, if $P_n(x)$ is a family of orthogonal polynomials, with $P_{-1}(x)=0$ and $P_0(x)=1$, then we can find two sequences $\alpha_n$ and $\beta_n$ such that $$P_n(x)=(x-\alpha_n) P_{n-1}(x)-\beta_n P_{n-2}(x).$$ \noindent In the case that $\alpha_n=\alpha$ and $\beta_n=\beta$ are constant sequences, then the coefficient array $$\{d_{n,k}\}$$ of the polynomials is a Riordan array \cite{Barry_Moment, Barry_Meixner, SGWW}, where $$P_n(x)=\sum_{k=0}^n d_{n,k}x^k.$$ This Riordan array will then be of the form $$\left(\frac{1-\delta x-\epsilon x^2}{1+\alpha x+\beta x^2}, \frac{1}{1+\alpha x+\beta x^2}\right).$$ In this note, we shall look at the situation of Laurent biorthogonal polynomials whose defining recurrences have constant coefficients. Based on the above results for orthogonal polynomials, it is natural to ask if there is a similar connection between Riordan arrays and Laurent biorthogonal polynomials whose defining recurrences have constant coefficients. We shall show that the answer is in the affirmative. We recall that a family of polynomials $P_n(x)$ is said to be a family of Laurent biorthogonal polynomials \cite{Hendriksen, Kamioka_14, Kamioka_15, Zhedanov} if there exist two sequences $\alpha_n$ and $\beta_n$ such that $$P_n(x)=(x-\alpha_n) P_{n-1}(x)- \beta_n x P_{n-2}(x).$$ \noindent The results in this note are typified by the following two propositions. \begin{proposition} Let $P_n(x)=P_n(x;\alpha, \beta)$ be the family of Laurent biorthogonal polynomials defined by the recurrence $$P_n(x)=(x-\alpha) P_{n-1}(x)-\beta x P_{n-2}(x),$$ with $P_0(x)=1$ and $P_1(x)=x-\alpha$. Then the coefficient array of the polynomials $P_n(x)$ is given by the Riordan array $$\left(\frac{1}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right).$$ \end{proposition} \noindent Similarly, we have \begin{proposition} Let $P_n(x)=P_n(x;\alpha, \beta)$ be the family of Laurent biorthogonal polynomials defined by the recurrence $$P_n(x)=(x-\alpha) P_{n-1}(x)-\beta x P_{n-2}(x),$$ with $P_0(x)=1$ and $P_1(x)=x-(\alpha+\beta)$. Then the coefficient array of the polynomials $P_n(x)$ is given by the Riordan array $$\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right).$$ \end{proposition} We will concentrate on this latter version, as we find it easier to work with the element $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$ of the Bell subgroup of the Riordan group. Before we prove the above result, we recall some notation and elements of Riordan group theory. Readers familiar with Riordan groups may skip this section. For an integer sequence $a_n$, that is, an element of $\mathbb{Z}^\mathbb{N}$, the power series $f(x)=\sum_{n=0}^{\infty}a_n x^n$ is called the \emph{ordinary generating function} or g.f. of the sequence. $a_n$ is thus the coefficient of $x^n$ in this series. We denote this by $a_n=[x^n]f(x)$. For instance, $F_n=[x^n]\frac{x}{1-x-x^2}$ is the $n$-th Fibonacci number \seqnum{A000045}, while $S_n=[x^n]\frac{1-\sqrt{1-6x+x^2}}{2x}$ is the $n$-th (large) Schr\"oder number \seqnum{A006318}. For a power series $f(x)=\sum_{n=0}^{\infty}a_n x^n$ with $f(0)=0$ we define the reversion or compositional inverse of $f$ to be the power series $\bar{f}(x)$ such that $f(\bar{f}(x))=x$. We shall sometimes write this as $\bar{f}= \text{Rev}f$. \newline\newline For a lower triangular matrix $(a_{n,k})_{n,k \ge 0}$ the row sums give the sequence with general term $\sum_{k=0}^n a_{n,k}$. \noindent The \emph{Riordan group} \cite{SGWW, Spru}, is a set of infinite lower-triangular integer matrices, where each matrix is defined by a pair of generating functions $g(x)=1+g_1x+g_2x^2+\cdots$ and $f(x)=f_1x+f_2x^2+\cdots$ where $f_1\ne 0$ \cite{Spru}. We assume in addition that $f_1=1$ in what follows. The associated matrix is the matrix whose $i$-th column is generated by $g(x)f(x)^i$ (the first column being indexed by 0). The matrix corresponding to the pair $g, f$ is denoted by $(g, f)$ or $\cal{R}$$(g,f)$. The group law is then given by \begin{displaymath} (g, f)\cdot(h, l)=(g, f)(h, l)=(g(h\circ f), l\circ f).\end{displaymath} The identity for this law is $I=(1,x)$ and the inverse of $(g, f)$ is $(g, f)^{-1}=(1/(g\circ \bar{f}), \bar{f})$ where $\bar{f}$ is the compositional inverse of $f$. If $\mathbf{M}$ is the matrix $(g,f)$, and $\mathbf{a}=(a_0,a_1,\ldots)^T$ is an integer sequence with ordinary generating function $\cal{A}$ $(x)$, then the sequence $\mathbf{M}\mathbf{a}$ has ordinary generating function $g(x)$$\cal{A}$$(f(x))$. The (infinite) matrix $(g,f)$ can thus be considered to act on the ring of integer sequences $\mathbb{Z}^\mathbb{N}$ by multiplication, where a sequence is regarded as a (infinite) column vector. We can extend this action to the ring of power series $\mathbb{Z}[[x]]$ by $$(g,f):\cal{A}(\mathnormal{x}) \mapsto \mathnormal{(g,f)}\cdot \cal{A}\mathnormal{(x)=g(x)}\cal{A}\mathnormal{(f(x))}.$$ \begin{example} The so-called \emph{binomial matrix} $\mathbf{B}$ is the element $(\frac{1}{1-x},\frac{x}{1-x})$ of the Riordan group. It has general element $\binom{n}{k}$, and hence as an array coincides with Pascal's triangle. More generally, $\mathbf{B}^m$ is the element $(\frac{1}{1-m x},\frac{x}{1-mx})$ of the Riordan group, with general term $\binom{n}{k}m^{n-k}$. It is easy to show that the inverse $\mathbf{B}^{-m}$ of $\mathbf{B}^m$ is given by $(\frac{1}{1+mx},\frac{x}{1+mx})$. \end{example} The proof of the propositions above is dependent on the sequence characterization of Riordan arrays \cite{He, Alter}, and in particular on the A-sequence. One version of the sequence characterization of Riordan arrays is given below. \begin{proposition} \cite{He} Let $D=[d_{n,k}]$ be an infinite triangular matrix. Then $D$ is a Riordan array if and only if there exist two sequences $A=[a_0,a_1,a_2,\ldots]$ and $Z=[z_0,z_1,z_2,\ldots]$ with $a_0 \neq 0$ such that \begin{itemize} \item $d_{n+1,k+1}=\sum_{j=0}^{\infty} a_j d_{n,k+j}, \quad (k,n=0,1,\ldots)$ \item $d_{n+1,0}=\sum_{j=0}^{\infty} z_j d_{n,j}, \quad (n=0,1,\ldots)$. \end{itemize} \end{proposition} The coefficients $a_0,a_1,a_2,\ldots$ and $z_0,z_1,z_2,\ldots$ are called the $A$-sequence and the $Z$-sequence of the Riordan array $D=(g(x),f(x))$, respectively. Letting $A(x)$ and $Z(x)$ denote the generating functions of these sequences, respectively, we have \cite{Alter} that $$\frac{f(x)}{x}=A(f(x)), \quad g(x)=\frac{d_{0,0}}{1-xZ(f(x))}.$$ We therefore deduce that $$A(x)=\frac{x}{\bar{f}(x)},$$ and $$Z(x)=\frac{1}{\bar{f}(x)}\left[1-\frac{d_{0,0}}{g(\bar{f}(x))}\right].$$ When $(g, f)=(g, xg)$, we then obtain that $$Z(x)=\frac{A(x)-1}{x}.$$ \noindent We now turn to the proof of the proposition. \begin{proof} We let $\left(d_{n,k}\right)$ be the coefficient array of the polynomials $P_n(x)$ that are defined by the recurrence $$P_n(x)=(x-\alpha) P_{n-1}(x)-\beta x P_{n-2}(x).$$ \noindent The recurrence implies that $$\sum_{k=0}^n d_{n,k}x^k=(x-\alpha) \sum_{k=0}^{n-1}d_{n-1,k}x^k-\beta x \sum_{k=0}^{n-2}d_{n-2,k}x^k,$$ or $$\sum_{k=0}^n d_{n,k}x^k=\sum_{k=0}^{n-1}d_{n-1,k}x^{k+1}-\alpha \sum_{k=0}^{n-1}d_{n-1,k}x^k - \beta \sum_{k=0}^{n-2}d_{n-2,k}x^{k+1},$$ or $$\sum_{k=0}^n d_{n,k}x^k=\sum_{k=1}^n d_{n-1,k-1}x^k-\alpha \sum_{k=0}^{n-1}d_{n-1,k}x^k - \beta \sum_{k=1}^{n-1}d_{n-2,k-1}x^k.$$ Thus for $0<k<n$, we require that $$d_{n,k}=d_{n-1,k-1}-\alpha d_{n-1,k}- \beta d_{n-2,k-1}.$$ Now the A-sequence of the Riordan array $\left(\frac{1}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$ (and that of $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$) is generated by the power series $$A(x)=\frac{x}{\bar{f}(x)},$$ where $f(x)=\frac{x(1-\beta x)}{1+\alpha x}$. Thus $$A(x)=\frac{2 \beta x}{1-\alpha x - \sqrt{1-(\alpha+2\beta)x+\alpha^2 x^2}},$$ and the A-sequence $[a_0, a_1, a_2,\ldots]$ begins $$1, -(\alpha+\beta), \beta(\alpha+\beta), -\beta(\alpha+\beta)(\alpha+2\beta),-\beta (\alpha+\beta)((\alpha^2+5 \alpha \beta+5 \beta^2), \ldots.$$ By the sequence characterization of Riordan arrays, we have $$d_{n,k+1}=a_0d_{n-1,k}+a_1 d_{n-1,k+1}+a_2 d_{n-1,k+2}+\ldots .$$ We can reverse this relation \cite{Alter} using the coefficients $b_n$ obtained by $$b_n=-\frac{1}{a_0} \sum_{j=1}^n a_j b_{n-j}, \quad b_0=\frac{1}{a_0},$$ to obtain $$d_{n-1,k}=a_0d_{n,k+1}-a_1 d_{n,k+1}-(a_2-a_1^2) d_{n,k+2}-(a_3-2a_1a_2+a_1^3)d_{n,k+3}+\ldots$$ which in this case gives $$d_{n-1,k}=d_{n,k+1}+(\alpha+\beta)d_{n,k+1}+(\alpha+\beta)(\alpha+2\beta)d_{n,k+2}+(\alpha+\beta)(\alpha^2+5\alpha \beta+5\beta^2)d_{n,k+3}+\ldots.$$ Thus we get \begin{eqnarray*} d_{n+1,k+1}&=&d_{n,k}-(\alpha+\beta)d_{n,k+1}-\beta(\alpha+\beta)d_{n,k+2}-\beta(\alpha+\beta)(\alpha+2\beta)d_{n,k+3}+\ldots \\ &=&d_{n,k}-\alpha d_{n,k+1}-\beta \{d_{n,k+1}+(\alpha+\beta)d_{n,k+2}+\beta(\alpha+\beta)(\alpha+2\beta)d_{n,k+3}+\ldots\},\\ &=&d_{n,k}-\alpha d_{n,k+1}-\beta d_{n-1,k}.\end{eqnarray*} \end{proof} \noindent We can derive an expression for $P_n(x)$ using the definition of Riordan arrays. The $(n,k)$-th element $d_{n,k}$ of the coefficient array $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$ is given by \begin{eqnarray*}d_{n,k}&=&[x^n] \frac{1-\beta x}{1+\alpha x}\left(\frac{x(1-\beta x)}{1+\alpha x}\right)^k\\ &=& [x^{n-k}]\frac{(1-\beta x)^{k+1}} {(1+\alpha x)^{k+1}}\\ &=& (-1)^{n-k} \sum_{j=0}^{k+1}\binom{k+1}{j}\binom{n-j}{n-k-j}\alpha^{n-k-j}\beta^j. \end{eqnarray*} Thus we get $$P_n(x)=\sum_{k=0}^n (-1)^{n-k} \sum_{j=0}^{k+1} \binom{k+1}{j}\binom{n-j}{n-k-j}\alpha^{n-k-j}\beta^j x^k.$$ \noindent We note that an ``$\alpha_{i,j}$" matrix \cite{Alter} for the Riordan array $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$ is given by $$\begin{bmatrix} -\beta & 0\\1 &-\alpha \end{bmatrix}.$$ \begin{example} We take the simple case of $\alpha=\beta=1$. Thus we look at the Riordan array $$\left(\frac{1-x}{1+x}, \frac{x(1-x)}{1+x}\right).$$ \noindent This matrix begins \begin{displaymath}\left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 & 0 & 0 & \cdots \\-2 & 1 & 0 & 0 & 0 & 0 & \cdots \\ 2 & -4 & 1 & 0 & 0 & 0 & \cdots \\ -2 & 8 & -6 & 1 & 0 & 0 & \cdots \\ 2 & -12 & 18 & -8 & 1 & 0 & \cdots \\-2 & 16 & -38 & 32 & -10 & 1 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right),\end{displaymath} which is the coefficient array of the polynomials $P_n(x)=P_n(x;1,1)$ which satisfy $$P_n(x)=(x-1)P_{n-1}(x)-xP_{n-2}.$$ \noindent The inverse array begins \begin{displaymath}\left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 & 0 & 0 & \cdots \\2 & 1 & 0 & 0 & 0 & 0 & \cdots \\ 6 & 4 & 1 & 0 & 0 & 0 & \cdots \\ 22 & 16 & 6 & 1 & 0 & 0 & \cdots \\ 90 & 68 & 30 & 8 & 1 & 0 & \cdots \\394 & 304 & 146 & 48 & 10 & 1 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right),\end{displaymath} where we see that the moments are given by the large Schr\"oder numbers $1,2,6,22,90,\ldots$, \seqnum{A06318}. \end{example} \noindent Other examples of arrays of this type may be found for instance in the On-Line Encyclopedia of Integer Sequences \cite{SL1, SL2}, where the last two number triangles are \seqnum{A080246} and \seqnum{A080247}, respectively. \section{Associated orthogonal polynomials} We can associate a family of orthogonal polynomials to the family $P_n(x;\alpha, \beta)$ by taking a suitable $\beta$-fold inverse binomial transform. \begin{proposition} The matrix $$\left(1, \frac{x}{1-\beta x}\right)^{-1} \cdot \left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$$ is the coefficient array of the family of orthogonal polynomials $\tilde{P}_n(x)=\tilde{P}_n(x;\alpha, \beta)$ that satisfies the three-term recurrence $$\tilde{P}_n(x)=(x-(\alpha+2\beta))\tilde{P}_{n-1}(x)-\beta(\alpha+\beta) \tilde{P}_{n-1}(x),$$ with $\tilde{P}_0(x)=1$ and $\tilde{P}_1(x)=x-(\alpha+\beta)$. \end{proposition} \begin{proof} We have $$\left(1,\frac{x}{1-\beta x}\right)^{-1}\cdot \left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)= \left(\frac{1}{1+(\alpha+\beta)x},\frac{x}{(1+\beta x)(1+(\alpha+\beta )x}\right).$$ The inverse of this Riordan array has tri-diagonal production matrix which begins \begin{displaymath}\left(\begin{array}{ccccccc} \alpha+\beta & 1 & 0 & 0 & 0 & 0 & \ldots \\ \beta(\alpha+\beta) & \alpha+2\beta & 1 & 0 & 0 & 0 & \ldots \\ 0 & \beta(\alpha+\beta) & \alpha+2\beta & 1 & 0 & 0 & \ldots \\ 0 & 0 & \beta(\alpha+\beta) & \alpha+2\beta & 1 & 0 & \ldots \\ 0 & 0 & 0 & \beta(\alpha+\beta) & \alpha+2\beta & 1 & \ldots \\0 & 0 & 0 & 0 & \beta(\alpha+\beta) & \alpha+2\beta &\ldots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right),\end{displaymath} from which we deduce the three-term recurrence. \end{proof} \begin{corollary} $$\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)= \left(1,\frac{x}{1-\beta x}\right)\cdot \left(\frac{1}{1+(\alpha+\beta)x},\frac{x}{(1+\beta x)(1+(\alpha+\beta )x}\right).$$ \end{corollary} \begin{corollary} We have the relations $$P_n(x)=\sum_{k=0}^n \binom{n-1}{n-k}\beta^{n-k}\tilde{P}_k(x), \quad \quad \tilde{P}_n(x)=\sum_{k=0}^n \binom{n-1}{n-k}(-\beta)^{n-k}P_k(x).$$ \end{corollary} \noindent We finish this section by noting that $$\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)\cdot \frac{1}{1-tx}=\frac{1-\beta x}{1+(\alpha-t)x+\beta t x^2},$$ and hence we have the generating function $$\frac{1-\beta x}{1+(\alpha-t)x+\beta t x^2}=\sum_{n=0}^{\infty} P_n(t)x^n.$$ \section{Moments and T-fractions} We identify the moments of $P_n(x)$ and $\tilde{P}_n(x)$ with the first column elements of the inverses of their coefficient arrays. The moments coincide, since they both have the generating function $$\frac{1-\alpha x- \sqrt{1-(\alpha+2\beta)x+\alpha^2 x^2}}{2 \beta x}.$$ In order to find a closed expression for these moments, we calculate the inverse of $\left(\frac{1-\beta x}{1+\alpha x},\frac{x(1-\beta x)}{1+\alpha x}\right)$. We use Lagrange inversion \cite{Merlini_When} for this. Given that $\left(\frac{1-\beta x}{1+\alpha x},\frac{x(1-\beta x)}{1+\alpha x}\right)$ is an element of the Bell subgroup of the Riordan group, its inverse will be of the form $\left(\frac{v(x)}{x}, v(x)\right)$, where $$v(x)=\text{Rev}\left(\frac{x(1-\beta x)}{1+\alpha x}\right).$$ \noindent Then we have \begin{eqnarray*} [x^n]\frac{v(x)}{x} v(x)^k &=& [x^{n+1}] v(x)^{k+1}\\ &=& [x^{n+1}] \left(\text{Rev}\left(\frac{x(1-\beta x)}{1+\alpha x}\right)\right)^{k+1}\\ &=&\frac{1}{n+1} [x^n] (k+1) x^k\left(\frac{1+\alpha x}{1-\beta x}\right)^{n+1}\\ &=& \frac{k+1}{n+1}[x^{n-k}] \left(\frac{1+\alpha x}{1-\beta x}\right)^{n+1}\\ &=& \frac{k+1}{n+1} [x^{n-k}] \sum_{j=0}^{n+1} \alpha^j x^j \sum_{i=0}^{\infty} \binom{-(n+1)}{i}(-1)^i \beta^i x^i\\ &=& \frac{k+1}{n+1} \sum_{j=0}^{n+1} \binom{n+1}{j}\binom{2n-k-j}{n-k-j}\alpha^j \beta^{n-k-j}.\end{eqnarray*} \noindent Thus the moments $\mu_n$ of the family of Laurent biorthogonal polynomials $P_n(x)$ are given by $$\mu_n=\frac{1}{n+1} \sum_{j=0}^{n+1} \binom{n+1}{j}\binom{2n-j}{n-j}\alpha^j \beta^{n-j}.$$ A more compact form is $$\mu_n=\sum_{k=0}^n \binom{n+k}{2k}C_k \alpha^{n-k} \beta^k,$$ where $C_n=\frac{1}{n+1}\binom{2n}{n}$ is the $n$-the Catalan number \seqnum{A000108}. The theory of $T$-fractions and Laurent biorthogonal polynomials now tells us that the moments $\mu_n$ are generated by the following continued fraction \cite{Wall}. $$g_{\alpha,\beta}(x)= \cfrac{1}{1-\alpha x- \cfrac{\beta x}{1-\alpha x- \cfrac{\beta x}{1-\alpha x- \cdots}}}.$$ Now the sequence $\mu_n$ is also the moment sequence for the orthogonal polynomials $\tilde{P}_n(x)$. Hence we also have the following Stieltjes continued fraction for $g_{\alpha, \beta}(x)$. $$g_{\alpha, \beta}(x)= \cfrac{1}{1-(\alpha+\beta)x- \cfrac{\beta(\alpha+\beta)x^2}{1-(\alpha+2\beta)x- \cfrac{\beta(\alpha+\beta)x^2}{1-(\alpha+2\beta)x-\cdots}}}.$$ \noindent From this we deduce the following result concerning the Hankel transform of the moments $\mu_n$ \cite{Layman}. \begin{proposition} The Hankel transform of the moments $\mu_n$ of $P_n(x)$ is given by $$h_n=(\beta(\alpha+\beta))^{\binom{n+1}{2}}.$$ \end{proposition} \begin{example} The large Schr\"oder numbers \seqnum{A006318} $$S_n=\sum_{k=0}^n \binom{n+k}{2k}C_k$$ are the case $\alpha=\beta=1$. Thus the generating function for the large Schr\"oder numbers may be written $$ \cfrac{1}{1- x- \cfrac{x}{1- x- \cfrac{ x}{1- x- \cdots}}}.$$ \noindent It is a classical result that the Hankel transform of the large Schr\"oder numbers is $2^{\binom{n+1}{2}}$. \end{example} \noindent It is interesting to note that the row sums of the inverse matrix $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)^{-1}$ have a generating function that can be expressed as the continued fraction $$g_{\alpha, \beta}(x)= \cfrac{1}{1-(\alpha+\beta+1)x- \cfrac{\beta(\alpha+\beta)x^2}{1-(\alpha+2\beta)x- \cfrac{\beta(\alpha+\beta)x^2}{1-(\alpha+2\beta)x-\cdots}}}.$$ Hence they too have a Hankel transform given by $$h_n=(\beta(\alpha+\beta))^{\binom{n+1}{2}}.$$ The form of the continued fraction shows that the row sums are also moments for a family of orthogonal polynomials whose parameters can be read from the continued fraction. Finally, the bi-variate generating function for the moment matrix $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)^{-1}$ is given by the generating function $$ \cfrac{1}{1-(\alpha+\beta)x-xy- \cfrac{\beta(\alpha+\beta)x^2}{1-(\alpha+2\beta)x- \cfrac{\beta(\alpha+\beta)x^2}{1-(\alpha+2\beta)x-\cdots}}}.$$ \begin{example} The orthogonal polynomials $\tilde{P}_n(x;1,1)$ associated with the Riordan array $\left(\frac{1-\beta x}{1+\alpha x},\frac{x(1-\beta x)}{1+\alpha x}\right)$ have coefficient array given by $$\left(\frac{1}{1+2x}, \frac{x}{1+3x+2x^2}\right).$$ \noindent The inverse of this matrix begins \begin{displaymath}\left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 & 0 & 0 & \cdots \\2 & 1 & 0 & 0 & 0 & 0 & \cdots \\ 6 & 5 & 1 & 0 & 0 & 0 & \cdots \\ 22 & 23 & 8 & 1 & 0 & 0 & \cdots \\ 90 & 107 & 49 & 11 & 1 & 0 & \cdots \\394 & 509 & 276 & 84 & 14 & 1 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right),\end{displaymath} indicating that the large Schr\"oder numbers are the moments for this family of orthogonal polynomials. This is borne out by the fact that the inverse is given by $$\left(\frac{1-x-\sqrt{1-6x+x^2}}{2x}, \frac{1-3x-\sqrt{1-6x+x^2}}{4x}\right).$$ \noindent This is \seqnum{A133367}. \end{example} \section{Derivatives} We consider the derivatives $$\frac{d}{dx} P_n(x)=R_{n-1}(x), \quad n>0.$$ Since $P_n(x)$ is precisely of degree $n$, $R_n(x)$ is also of degree $n$, with $$R_0(x)=\frac{d}{dx}(x-(\alpha+\beta))=1.$$ We let $e_{n,k}$ be the $(n,k)$-th term of the coefficient array of $R_n(x)$. Thus we have $$R_n(x)=\frac{d}{dx}P_{n+1}(x)=\sum_{k=0}^n e_{n,k}x^k.$$ \begin{proposition} The coefficient array whose $(n,k)$-th term is $\frac{1}{n+1} e_{n,k}$ is given by the Riordan array $$\left(\left(\frac{1-\beta x}{1+\alpha x}\right)^2, \frac{x(1-\beta x)}{1+\alpha x}\right).$$ \end{proposition} \noindent This results from a general result concerning Riordan arrays. We first note that we have \begin{displaymath}\left(\begin{array}{ccccccc} 0 & 0 & 0 & 0 & 0 & 0 & \cdots \\1 & 0 & 0 & 0 & 0 & 0 & \cdots \\ 0 & 2 & 0 & 0 & 0 & 0 & \cdots \\ 0 & 0 & 3 & 0 & 0 & 0 & \cdots \\ 0 & 0 & 0 & 4 & 0 & 0 & \cdots \\0 & 0 & 0 & 0 & 5 & 0 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right)\cdot \left(\begin{array}{c}1\\x\\x^2\\x^3\\x^4\\x^5\\\vdots\end{array}\right) = \left(\begin{array}{c}0\\1\\2x\\3x^2\\4x^3\\5x^4\\\vdots\end{array}\right).\end{displaymath} Thus we have \begin{displaymath}\left(\begin{array}{ccccccc} d_{0,0} & 0 & 0 & 0 & 0 & 0 & \cdots \\d_{1,0} & d_{1,1} & 0 & 0 & 0 & 0 & \cdots \\ d_{2,0} & d_{2,1} & d_{2,2} & 0 & 0 & 0 & \cdots \\ d_{3,0} & d_{3,1} & d_{3,2} & d_{3,3} & 0 & 0 & \cdots \\ d_{4,0} & d_{4,1} & d_{4,2} & d_{4,3} & d_{4,4} & 0 & \cdots \\d_{5,0} & d_{5,1} & d_{5,2} & d_{5,3} & d_{5,4} & d_{5,5} &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right)\cdot \left(\begin{array}{ccccccc} 0 & 0 & 0 & 0 & 0 & 0 & \cdots \\1 & 0 & 0 & 0 & 0 & 0 & \cdots \\ 0 & 2 & 0 & 0 & 0 & 0 & \cdots \\ 0 & 0 & 3 & 0 & 0 & 0 & \cdots \\ 0 & 0 & 0 & 4 & 0 & 0 & \cdots \\0 & 0 & 0 & 0 & 5 & 0 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right)\cdot \left(\begin{array}{c}1\\x\\x^2\\x^3\\x^4\\x^5\\\vdots\end{array}\right) = \left(\begin{array}{c}0\\R_0(x)\\R_1(x)\\R_2(x)\\R_3(x)\\R_4(x)\\\vdots\end{array}\right),\end{displaymath} where in this case $\left(d_{n,k}\right)$ represents the Riordan array $(g, f)=\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x)}{1+\alpha x}\right)$. We then have \begin{displaymath}\left(\begin{array}{ccccccc} d_{1,1} & 0 & 0 & 0 & 0 & 0 & \cdots \\d_{2,1} & d_{2,2} & 0 & 0 & 0 & 0 & \cdots \\ d_{3,1} & d_{3,2} & d_{3,3} & 0 & 0 & 0 & \cdots \\ d_{4,1} & d_{4,2} & d_{4,3} & d_{4,4} & 0 & 0 & \cdots \\ d_{5,1} & d_{5,2} & d_{5,3} & d_{5,4} & d_{5,5} & 0 & \cdots \\d_{6,1} & d_{6,2} & d_{6,3} & d_{6,4} & d_{6,5} & d_{6,6} &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right)\cdot \left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 & 0 & 0 & \cdots \\0 & 2 & 0 & 0 & 0 & 0 & \cdots \\ 0 & 0 & 3 & 0 & 0 & 0 & \cdots \\ 0 & 0 & 0 & 4 & 0 & 0 & \cdots \\ 0 & 0 & 0 & 0 & 5 & 0 & \cdots \\0 & 0 & 0 & 0 & 0 & 6 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right)\cdot \left(\begin{array}{c}1\\x\\x^2\\x^3\\x^4\\x^5\\\vdots\end{array}\right) = \left(\begin{array}{c}R_0(x)\\R_1(x)\\R_2(x)\\R_3(x)\\R_4(x)\\R_5(x)\\\vdots\end{array}\right).\end{displaymath} \noindent By the Riordan array structure of $(g, f)$, the leftmost matrix is given by $(gf, f)$. In other words, we have $$\left(\left(\frac{1-\beta x}{1+\alpha x}\right)^2, \frac{x(1-\beta x)}{1+\alpha x}\right) \cdot \text{Diag}(1,2,3,\ldots)=\left(\begin{array}{c}R_0(x)\\R_1(x)\\R_2(x)\\R_3(x)\\R_4(x)\\R_5(x)\\\vdots\end{array}\right).$$ \section{Generalized orthogonality} Ismail and Masson \cite{I_M} have studied a notion of generalized orthogonality, associated with polynomials that satisfy recurrences of the type $$P_n(x)=(x-c_n)P_{n-1}(x)-\lambda_n(x-a_n)P_{n-2}.$$ \noindent We specialize this to the constant coefficient case. Thus in this section we let $P_n(x)=P_n(x;\alpha, \beta, \gamma)$ be a family of polynomials that obeys the recurrence $$P_n(x)=(x-\alpha)P_{n-1}(x)-\beta(x-\gamma)P_{n-2}(x),$$ with $P_0(x)=1$ and $P_1(x)=x-\alpha-\beta$. Following a similar development to that in the first section, we arrive at the following results. \begin{proposition} The coefficient array of the polynomials $P_n(x)$ is given by the Riordan array $$\left(\frac{1-\beta x}{1+\alpha x - \beta \gamma x^2}, \frac{x(1-\beta x)}{1+\alpha x - \beta \gamma x^2}\right).$$ \end{proposition} \noindent Note that when $\gamma=0$, we retrieve our original case. When $\alpha+\beta \ne \gamma$, we have \begin{proposition} The moments $\mu_n$ of the family of polynomials $P_n(x)$ are also the moments of the family of associated orthogonal polynomials $\tilde{P}_n(x)$ whose coefficient array is given by the Riordan array $$\left(\frac{1+\beta x}{1+(\alpha+2\beta)x+\beta(\alpha+\beta-\gamma)x^2},\frac{x}{1+(\alpha+2\beta)x+\beta(\alpha+\beta-\gamma)x^2}\right).$$ \end{proposition} \begin{corollary} The Hankel transform of the moments $\mu_n$ is given by $$h_n=(\beta(\alpha+\beta-\gamma))^{\binom{n+1}{2}}.$$ \end{corollary} \noindent The relationship between the generalized orthogonal polynomials $P_n(x)$ of this section and their associated orthogonal polynomials $\tilde{P}_n(x)$ is given by the following proposition, which can be verified by applying the multiplication rule for Riordan arrays. \begin{proposition} \footnotesize $$ \left(\frac{1-\beta x}{1+\alpha x - \beta \gamma x^2}, \frac{x(1-\beta x)}{1+\alpha x - \beta \gamma x^2}\right)= \tilde{B}\cdot \left(\frac{1+\beta x}{1+(\alpha+2\beta)x+\beta(\alpha+\beta-\gamma)x^2},\frac{x}{1+(\alpha+2\beta)x+\beta(\alpha+\beta-\gamma)x^2}\right),$$ \normalsize where $$\tilde{B}=\left(1, \frac{x}{1-\beta x}\right).$$ \end{proposition} \begin{example} We consider the polynomials $P_n(x)=P_n(x;2,1,1)$ with coefficient array $$\left(\frac{1-x}{1+2x-x^2}, \frac{x(1-x)}{1+2x-x^2}\right).$$ \noindent This begins \begin{displaymath}\left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 &0 & 0 & \cdots \\-3 & 1 & 0 & 0 & 0 & 0 & \cdots \\ 7 & -6 & 1 & 0 & 0 & 0 & \cdots \\ -17 & 23 & -9 & 1 & 0 & 0 & \cdots \\ 41 & -76 & 48 & -12 & 1 & 0 & \cdots \\-99 & 233 & -204 & 82 & -15 & 1 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right).\end{displaymath} $P_n(x)$ satisfies the recurrence $$P_n(x)=(x-2)P_{n-1}-(x-1)P_{n-2}$$ with $P_0(x)=1$ and $P_1(x)=x-3$. The moment array is given by $\left(\frac{1-x}{1+2x-x^2}, \frac{x(1-x)}{1+2x-x^2}\right)^{-1}$ which begins \begin{displaymath}\left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 &0 & 0 & \cdots \\3 & 1 & 0 & 0 & 0 & 0 & \cdots \\11 & 6 & 1 & 0 & 0 & 0 & \cdots \\ 47 & 31 & 9 & 1 & 0 & 0 & \cdots \\ 223 & 160 & 60 & 12 & 1 & 0 & \cdots \\ 1135 & 849 & 366 & 98 & 15 & 1 &\cdots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right).\end{displaymath} Thus the moment sequence $\mu_n$ starts $$1,3,11,47,223,1135,\ldots$$ \noindent This is the binomial transform \seqnum{A174347} of the large Schr\"oder numbers. $$\mu_n=\sum_{k=0}^n \binom{n}{k} S_k.$$ \noindent This is a consequence of the fact that \begin{equation}\label{Eq}\left(\frac{1-x}{1+2x-x^2}, \frac{x(1-x)}{1+2x-x^2}\right) \cdot B = \left(\frac{1-x}{1+x}, \frac{x(1-x)}{1+x}\right).\end{equation} \noindent We note that $\left(\frac{1-x}{1+x}, \frac{x(1-x)}{1+x}\right)$ is the coefficient array of the Laurent biorthogonal polynomials $$Q_n(x)=(x-1)Q_{n-1}-xQ_{n-2}.$$ Equation (\ref{Eq}) is equivalent to $$P_n(x+1)=Q_n(x).$$ \end{example} \noindent We can also transform the coefficient array $ \left(\frac{1-\beta x}{1+\alpha x - \beta \gamma x^2}, \frac{x(1-\beta x)}{1+\alpha x - \beta \gamma x^2}\right)$ of $P_n(x;\alpha, \beta, \gamma)$ by multiplication on the right to obtain the coefficient array of an associated family of Laurent biorthogonal polynomials. This is the content of the next result, which is verifiable by straight-forward Riordan array multiplication. \begin{proposition} We have $$\left(\frac{1-\beta x}{1+\alpha x - \beta \gamma x^2}, \frac{x(1-\beta x)}{1+\alpha x - \beta \gamma x^2}\right) \cdot B_{\gamma} = \left(\frac{1-\beta x}{1+(\alpha-\gamma)x}, \frac{x(1-\beta x)}{1+(\alpha-\gamma)x}\right),$$ where $$B_{\gamma}=\left(\frac{1}{1-\gamma x},\frac{x}{1-\gamma x}\right).$$ \end{proposition} \begin{corollary} Using an obvious notation, we have $$P_n(x+\gamma; \alpha, \beta, \gamma)=P_n(x; \alpha-\gamma, \beta).$$ \end{corollary} \begin{proof} This follows since $$B_{\gamma}\cdot (1,x,x^2,\ldots)^T=(1,x+\gamma, (x+\gamma)^2, \ldots)^T.$$ \end{proof} \section{Determinant representations} It is possible to give determinant representations for the polynomial sequences \cite{Yang} in this note. Taking the polynomials defined by the Riordan array $M=\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x}{1+\alpha x}\right)$, we calculate $A(x)$ and $Z(x)$ for the inverse $\left(\frac{1-\beta x}{1+\alpha x}, \frac{x(1-\beta x}{1+\alpha x}\right)^{-1}$ to get $$A(x)=\frac{1+\alpha x}{1-\beta x}, \quad \quad Z(x)=\frac{\alpha + \beta }{1-\beta x}.$$ \noindent Thus the production matrix of the inverse $M^{-1}$ begins \begin{displaymath}\left(\begin{array}{ccccccc} \alpha+\beta & 1 & 0 & 0 & 0 & 0 & \ldots \\ \beta(\alpha+\beta) & \alpha+\beta & 1 & 0 & 0 & 0 & \ldots \\ \beta^2(\alpha+\beta) &\beta(\alpha+\beta) & \alpha+\beta & 1 & 0 & 0 & \ldots \\ \beta^3(\alpha+\beta) & \beta^2(\alpha+\beta) & \beta(\alpha+\beta) & \alpha+\beta & 1 & 0 & \ldots \\ \beta^4(\alpha+\beta) & \beta^3(\alpha+\beta) & \beta^2(\alpha+\beta) & \beta(\alpha+\beta) & \alpha+\beta & 1 & \ldots \\ \beta^5(\alpha+\beta) & \beta^4(\alpha+\beta) & \beta^3(\alpha+\beta) & \beta^2(\alpha+\beta) & \beta(\alpha+\beta) & \alpha+\beta &\ldots\\ \vdots & \vdots & \vdots & \vdots & \vdots & \vdots & \ddots\end{array}\right),\end{displaymath} \noindent Letting $\mathbf{P}_n$ denote the principal submatrix of order $n$ of this production matrix, we have $$P_n(x)=P_n(x;\alpha, \beta)=\det(x I_n-\mathbf{P}_n),$$ where $I_n$ is the identity matrix of order $n$. Similar remarks hold for the other polynomials defined in this article.
\section{The BinaMIcS project} BinaMIcS (Binarity and Magnetic Interactions in various classes of Stars) is an international project led by France (PI E. Alecian), which includes over 90 observers, modellers and theoreticians from 13 countries (see http://lesia.obspm.fr/BinaMIcS). The goal of the BinaMIcS project is to exploit binarity to yield new constraints on the physical processes at work in hot and cool magnetic stars. In particular, BinaMIcS aims at studying (1) the role of magnetism during stellar formation, (2) magnetospheric star-star (and star-planet) interactions, (3) the impact of tidal flows on fossil and dynamo magnetic fields, and (4) the impact of magnetism on mass and angular momentum transfer. Studying binaries rather than single magnetic stars allows to better constrain the fundamental parameters of the targets. Moreover, binaries are an ideal laboratory to study physical processes related to star-star interactions, such as tidal deformation, wind-wind collisions, magnetospheric coupling, tidally excited pulsations,... Finally, they are important to understand the origin of magnetic fields. For example, in cool stars, they allow to test the synchronisation of the binary versus the dynamo interaction; in intermediate-mass stars, magnetic fields are known to be anomalously rare in binaries and this needs to be understood; and in massive stars, some theories propose that a field could be generated in stellar mergers. However, spectropolarimetric studies of binary stars are particularly challenging. First, one needs to obtain spectropolarimetric measurements over the rotation periods of both stars and over the orbital period. Second, one needs to disentangle the intensity spectra of both components as well as the Zeeman signatures in the polarised light to be able to characterize the magnetic field of both stars. \section{The BinaMIcS observing program} The BinaMIcS project has been allocated two Large Programs (LP) of observations: one LP of 604 hours for 4 years (February 2013 to January 2017) with ESPaDOnS at CFHT (PIs E. Alecian and G. Wade), and one LP of 128 hours for 2 years (March 2013 to February 2015) with Narval at TBL (PI C. Neiner). This Narval LP is renewable for another 2 years. BinaMIcS observes 3 samples of targets: (1) The cool Targeted Component (cTC): this sample consists of selected cool magnetic SB2 dwarfs, RS CVn, BY Dra, W UMa and pre-main sequence stars, with various fundamental and orbital parameters. These targets will be observed with a great coverage to perform a detailed characterisation of their magnetic field as well as detailed modelling. The results will be compared with those obtained for single stars. We will also study the variations and correlation with eccentricity. (2) The hot Targeted Component (hTC): this sample consists of all 13 known magnetic SB2 systems of O, B or A spectral type visible from CFHT or TBL. This includes 1, 6 and 6 systems for which the hottest star is an O, B and A star, respectively. These targets will be observed with a great coverage to perform a detailed characterisation of their magnetic field as well as detailed modelling. The results will be compared with those obtained for single stars. We will also study the wind and magnetospheric interactions between the 2 components. (3) The hot Survey Component (SC): this sample consists of over 200 close SB2 binaries, including eclipsing ones, with at least one component of O, B, A or early F spectral type and a secondary component not later than F. The goal is to search for a magnetic field in these targets. Those systems that will be discovered to be magnetic will be transfered to the hTC sample. We will derive the statistical occurence of magnetic fields in this sample and compare it with the results obtained by the MiMeS collaboration for single stars (\cite{wade2013}). \section{First targeted observations} \begin{figure}[!ht] \begin{center} \resizebox{6.9cm}{!}{\includegraphics[clip]{sigcrb.eps}} \end{center} \caption{Examples of 3 spectropolarimetric observations of the cool system $\sigma^2$ CrB. Top: LSD Stokes V profiles showing the magnetic signatures in both stars. Bottom: Corresponding LSD intensity profiles.} \label{sig2crb} \end{figure} The BinaMIcS observations started in February 2013. A complete dataset has already been obtained for several TC stars. For the cTC, we observed BH CVn, a RS CVn (F2IV+K2IV) system for which only one of the 2 components is magnetic, and BY Dra (K6Ve+...) for which both components are found to be magnetic. Snapshots have also been obtained for a number of other cool systems (UZ Tau E, $\sigma^2$ CrB, V1379 Aql, ER Vul, HD\,216489, HD\,28, HD\,34029,...) to allow the selection of the best cTC targets. $\sigma^2$ CrB has been selected for follow-up as a cTC target since the first spectropolarimetric observations show clear magnetic detections in both components (see Fig.~\ref{sig2crb}). For the hTC, we already observed the Plaskett star, a system discovered to be magnetic by the MiMeS collaboration (Grunhut et al., these proceedings), as well as HD\,5550, an Ap+A system. These TC datasets are currently beeing analysed and a full characterisation of the magnetic fields of each system will be published in 2014. \section{Preliminary survey results} \begin{figure}[!ht] \begin{center} \resizebox{8.8cm}{!}{\includegraphics[clip]{hd160922_2col.eps}} \end{center} \caption{Examples of spectropolarimetric observations of the F4+F5 system HD\,160922. Left: LSD intensity profiles. Right: Corresponding LSD Stokes V profiles showing the magnetic signatures in one star.} \label{hd160922} \end{figure} At the time of writing, 93 close SB2 systems including at least one O, B, A or early F star have been observed. Among these systems, a magnetic field has been detected in one system: HD\,160922. This is a F4+F5 system with an orbital period of 5.28 days. Only one of the two stars in the system shows a magnetic signature. This target has been transferred to the hTC sample and we have acquired 15 measurements. The simple magnetic signature does not seem to vary much from one observation to another. Therefore this star could host a fossil field rather than a dynamo field (see Fig.~\ref{hd160922}). The detection of only one magnetic system among the 93 systems observed so far raises many questions. If we consider only the 79 systems including at least one O, B or A star, we can compare the occurence of a magnetic field in these systems with the occurence established by the MiMeS collaboration for single massive stars. In single stars, the occurence is 7\% (\cite{wade2013}). If the occurence was the same in binaries, we should have detected between 5 and 11 magnetic systems. Instead we found none. Although the statistics of our current sample is not sufficient to draw firm conclusions, the lack of detections in this sample may be an important result for stellar formation and the origin of magnetism in massive stars. The observations of the complete SC sample will provide firmer conclusions and cast new light on these issues. \section{Conclusions} BinaMIcS is an ambitious project to study magnetic interactions in hot and cool close binary systems, understand the role of magnetism during stellar formation and the origin of magnetic field in massive stars. The first BinaMIcS observations already provide very interesting results.\\ \noindent {\bf Acknowledgements} Based on observations obtained at the Telescope Bernard Lyot (TBL) operated by the Observatoire Midi-Pyr\'en\'ees, Universit\'e de Toulouse, Centre National de la Recherche Scientifique (CNRS) of France, and at the Canada-France-Hawaii Telescope (CFHT) operated by the National Research Council of Canada, the Institut National des Sciences de l'Univers of CNRS, and the University of Hawaii.
\section{Proofs} \noindent \textbf{Lemma \ref{lm:order}.} \textit{ The relation $\prec$ on RMTL formulas is a well-founded partial order. } \begin{proof} We first show that $\prec$ is well-founded. Suppose otherwise: then there is an infinite descending chain of formulas: $$ \cdots \prec_S \phi_n \prec_S \phi_{n-1} \prec_S \cdots \prec_S \phi_2 \prec_S \phi_1. $$ Obviously, none of $\phi_i$'s can be a bottom element (i.e., those that take the form as specified in clause (1) of Definition~\ref{def:order}). Furthermore, there must be an $i$ such that $\phi_i = P(\vec t)$ and $\phi_{i+1} = \phi_P(\vec t)$ where $P$ is a recursive predicate defined by $P(\vec x) := \phi_P(\vec x).$ If no such $i$ exists then all the instances of the relation $\prec_S$ in the chain must be instances of clause (3) and (4) in Definition~\ref{def:order}, and the chain would be finite as those two clauses relate only strict subformulas. So without loss of generality, let us assume that $\phi_1 = P(\vec t)$. We claim that for every $j > 1$, every occurrence of any recursive predicate in $\phi_j$ is guarded. We prove this by induction on $j$. If $j=2$ then we have $\phi_2 \prec_S P(\vec t)$. In this case, $\phi_2$ must be $\phi_P(\vec t)$, and by the guardedness condition, all recursive predicates in $\phi_P$ are guarded. If $j > 2$, then we have $\phi_j \prec_S \phi_{j-1}$. By induction hypothesis, all recursive predicates in $\phi_{j-1}$ are guarded. In this case, the relation $\phi_j \prec_S \phi_{j-1}$ must be an instance of either clause (3) or clause (4) of Definition~\ref{def:order}, and therefore $\phi_j$ also satisfies the guardedness condition. So now we have that none of $\phi_i$'s are recursive predicates. This means that all instances of $\prec_S$ in the chain must be instances of clause (3) and (4) in Definition~\ref{def:order}, and consequently the size of the formulas in the chain must be strictly decreasing. Thus the chain cannot be infinite, contrary to the assumption. Anti-symmetry follows immediately from well-foundedness. Suppose $\prec$ is not anti-symmetric. Then we have a chain $$ \phi = \phi_1 \prec_S \phi_2 \prec_S \phi_3 \prec_S \cdots \prec_S \phi_n = \phi $$ where $n > 1.$ We can repeat this chain to form an infinite descending chain, which contradicts the well-foundedness of $\prec.$ \qed \end{proof} \noindent \textbf{Lemma \ref{lm:minimal} (Minimality).} \textit{ If $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ ($(\rho, i) \models \diamonddot_n \phi$) then there exists an $m \leq n$ such that $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_m \phi_2$ (resp. $(\rho, i) \models \diamonddot_m \phi$) and such that for every $k$ such that $0 < k < m$, we have $(\rho, i) \not \models \phi_1 \mathrel{\mathbb{S}}_k \phi_2$ (resp., $(\rho, i) \not \models \diamonddot_k \phi$). } \begin{proof} We show a case for $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$; the other case is straightforward. We prove this by induction on $i.$ \begin{itemize} \item Base case: $i = 1.$ Since $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$, it must be the case that $(\rho, i) \models \phi_2$. In this case, let $m = 1.$ Obviously $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_m \phi_2$ and $m$ is minimal. \item Inductive case: $i > 1.$ We have $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$. By the definition of $\models$, there exists $j \leq i$ such that $(\rho, j) \models \phi_2$ and $(\rho, k) \models \phi_1$ for every $k$ s.t. $j < k \leq i$ and $\tau_i - \tau_j < n.$ If $(\rho, i) \models \phi_2$, then we have $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_1 \phi_2.$ In this case, let $m = 1.$ If $(\rho, i) \not \models \phi_2$, then it must be the case that $j < i$. It is not difficult to see that in this case we must have $$(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_{n - (\tau_i - \tau_{i-1})} \phi_2.$$ By the induction hypothesis, we have there is an $m'$ such that $$ (\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_{m'} \phi_2 $$ and for every $l$ s.t. $l < m'$, we have $(\rho, i-1) \not \models \phi_1 \mathrel{\mathbb{S}}_l \phi_2.$ In this case, we let $m = m' + \tau_i - \tau_{i-1}.$ It is straightforward to check that $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_m \phi_2.$ Now, we claim that this $m$ is minimal. Suppose otherwise, i.e., there exists $k < m$ s.t. $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_k \phi_2.$ Since $(\rho, i) \not \models \phi_2$, we must have $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_{k - (\tau_i - \tau_{i-1})} \phi_2$. But $(k - (\tau_i - \tau_{i-1})) < m'$, so this contradicts the minimality of $m'.$ \end{itemize} \qed \end{proof} \begin{lemma}[Monotonicity] \label{lm:monotonicity} If $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ (resp., $(\rho, i) \models \bullet_n \phi$ and $(\rho, i) \models \diamonddot_n \phi$) then for every $m \geq n$, we have $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_m \phi_2$ (resp., $(\rho, i) \models \bullet_m \phi$ and $(\rho, i) \models \diamonddot_m \phi$). \end{lemma} \begin{proof} Straightforward from the definition of $\models.$ \qed \end{proof} \noindent \textbf{Theorem \ref{thm:recursive-form} (Recursive forms).} \textit{ For every model $\rho$, every $n \geq 1$, $\phi$, $\phi_1$ and $\phi_2$, and every $1 < i \leq |\rho|$, the following hold:} \begin{enumerate} \item $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ iff $(\rho, i) \models \phi_2$, or $(\rho, i) \models \phi_1$ and $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ and $n - (\tau_i - \tau_{i-1}) \geq \mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2).$ \item $(\rho, i) \models \diamonddot_n \phi$ iff $(\rho, i - 1) \models \phi$ and $\tau_i - \tau_{i-1} < n$, or $(\rho, i-1) \models \diamonddot_n \phi$ and $n - (\tau_i - \tau_{i-1}) \geq \mathfrak{m}(\rho, i-1, \diamonddot_n \phi).$ \end{enumerate} \begin{proof} We show the case for $\mathrel{\mathbb{S}}_n$; the other case is similar. Suppose $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$. By definition, there exists $j \leq i$ such that \begin{equation} \label{eq:re1} (\rho, j) \models \phi_2, \end{equation} \begin{equation} \label{eq:re2} \tau_i - \tau_j \leq n, \end{equation} and for every $k$ s.t. $j < k \leq i$ we have \begin{equation} \label{eq:re3} (\rho, k) \models \phi_2, \end{equation} Suppose that the right-hand side of iff doesn't hold, i.e., we have $(\rho, i) \not \models \phi_2$, and that one of the following hold: \begin{itemize} \item $(\rho, i) \not \models \phi_1$, \item $(\rho, i-1) \not \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$, or \item $n - (\tau_i - \tau_{i-1}) < \mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2).$ \end{itemize} The first case contradicts our assumption in (\ref{eq:re3}), so it cannot hold. Note that since $(\rho, i) \not \models \phi_2$, it must be the case that $j \leq i-1$, so (\ref{eq:re1}), (\ref{eq:re2}), and (\ref{eq:re3}) above entail that $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$, so the second case can't hold either. For the third case: from (\ref{eq:re2}) we have $\tau_i - \tau_j = (\tau_i - \tau_{i-1}) + (\tau_{i-1} - \tau_j) \leq n$ so $$ \tau_{i-1} - \tau_j \leq n - (\tau_{i} - \tau_{i-1}). $$ This, together with (\ref{eq:re1}) and (\ref{eq:re2}), implies that $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_{n - (\tau_{i} - \tau_{i-1})} \phi_2$. If $n - (\tau_i - \tau_{i-1}) < \mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2)$ holds, then it would contradict the fact that $\mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2)$ is the minimal index as guaranteed by Lemma~\ref{lm:minimal}. Hence the third case cannot hold either. For the other direction, suppose that either of the following holds: \begin{itemize} \item $(\rho, i) \models \phi_2$, or \item $(\rho, i) \models \phi_1$ and $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ and $n - (\tau_i - \tau_{i-1}) \geq \mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2).$ \end{itemize} If it is the first case, i.e., $(\rho,i)\models \phi_2$, then trivially $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2.$ So suppose it is the second case. By Lemma~\ref{lm:minimal}, we have $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_m \phi_2$, where $m = \mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2).$ Since $n - (\tau_i - \tau_{i-1}) \geq m$, by Lemma~\ref{lm:monotonicity}, we have $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_{n - (\tau_i - \tau_{i-1})} \phi_2.$ This means there exists $j \leq i-1$ such that \begin{itemize} \item $(\rho, j) \models \phi_2$, \item $(\tau_{i-1} - \tau_j) \leq n - (\tau_i - \tau_{i-1})$, and \item $(\rho, k) \models \phi_1$, for every $k$ s.t. $j < k \leq i-1$. \end{itemize} The second item implies that $(\tau_i - \tau_j) \leq n$. These and the fact that $(\rho, i) \models \phi_1$ entail that $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2.$ \qed \end{proof} \noindent \textbf{Theorem 2.} \textit{ $(\rho, i) \models \phi$ iff $Monitor(\rho,i,\phi)$ returns true. } \begin{proof} The proof is quite straightforward, because each step in the calculation of the truth values of subformulas of $\phi$ corresponds to their (recursive-form) semantics reading and Theorem~\ref{thm:recursive-form}. Moreover, by the well-foundedness of $\prec$, this algorithm terminates, and at each step, the calculation of a subformula $\phi$ uses only values of other subformulas smaller than $\phi$ in the ordering $\prec$, or truth values of subformulas in the previous world (which are already defined). In the case of $\mathrel{\mathbb{S}}_n$, the updates of the value $mcur$ and $mprev$ corresponds exactly to the construction shown in the proof of Lemma~\ref{lm:minimal}. \qed \end{proof} \section{Conclusion, related and future work} \label{conclusion} We have shown a policy language design based on MTL that can effectively describe various scenarios of privilege escalation in Android. Moreover, any policy written in our language can be effectively enforced. The key to the latter is the fact that our enforcement procedure is trace-length independent. We have also given a proof-of-concept implementation on actual Android devices and show that our implementation can effectively enforce RMTL policies. We have already discussed related work in runtime monitoring based on LTL in the introduction. We now discuss briefly related work in Android security. There is a large body of works in this area, more than what can be reasonably surveyed here, so we shall focus on the most relevant ones to our work, i.e., those that deal with privilege escalation. For a more comprehensive survey on other security extensions or analysis, the interested reader can consult~\cite{LastPE}. QUIRE \cite{QUIRE} is an application centric approach to privilege escalation, done by tagging the intent objects with the caller's UID. Thus, the recipient application can check the permission of the source of the call chain. IPC Inspection \cite{IPCInspection} is another application centric solution that works by reducing the privilege of the recipient application when it receives a communication from a less privileged application. XManDroid \cite{LastPE} is a system centric solution, just like ours. Its security monitor maintains a call graph between apps. It is the closest to our solution, except that we are using temporal logic to specify a policy, and our policy can be modified modularly. Our policy language is also more expressive as we can specify both temporal and metric propertes. TaintDroid~\cite{TaintDroid} is another system-centric solution, but it is designed to track data flow, rather than control flow, via taint analysis, so privilege escalation can be inferred from leakage of data. We currently do not deal with quantifiers directly in our algorithm. Such quantifiers are expanded into purely propositional connectives (when the domain is finite), which is exponential in the number of variables in the policy. As an immediate future work, we plan to investigate whether techniques using {\em spawning automata}~\cite{bauer2013rv} can be adapted to our setting to allow a ``lazy'' expansion of quantifiers as needed. It is not possible to design trace-length-independent monitoring algorithms in the unrestricted first-order LTL~\cite{bauer2013rv}, so the challenge here is to find a suitable restriction that can be enforced efficiently. \section{Examples} \label{examples} We provide some basic policies as an example of how we can use this logic to specify security policies. From now on, we shall only quantify over the domain $app$, so in the following we shall omit the sort annotation in the existential quantifier. The predicate $trans$ is the recursive predicate defined in Equation~(\ref{eq:trans}) in the introduction. The constant $sink$ denotes a service or resource that an unprivileged application tries to access via privilege escalation e.g. send SMS, or access to internet. The constant $contact$ denotes the Contact provider app in Android. We also assume the following ``static'' predicates (i.e., their truth values do not vary over time): \begin{itemize} \item $system(x)$: $x$ is a system app or process. \item $hasPermissionToSink(y)$: $y$ has permission to access the sink. \item $trusted(x)$: $x$ is an app that the user trusts. This is not a feature of Android, rather, it is a specific feature of our implementation. We build into our implementation a `trust' management app to allow the user a limited control over apps that he/she trusts. \end{itemize} The following policies refer to access patterns that are forbidden. So given a policy $\phi$, the monitor at each state $i$ make sure that $(\rho,i) \not \models \phi$ holds. Assuming that $(\rho, i) \not \models \phi$, where $i = |\rho|$, holds, then whenever a new event (i.e., the IPC call) $e$ is registered at time $t$, the monitor checks that $(([\pi;e], [\tau;t]), i+1) \not \models \phi$ holds. If it does, then the call is allowed to proceed. Otherwise, it will be terminated. \begin{enumerate} \item $\exists x. (call(x,sink) \wedge \neg system(x) \wedge \neg trusted(x)).$ This is a simple policy where we block a direct call from any untrusted application to the sink. This policy can serve as a privilege manager where we dynamically revoke permission for application to access the sink regardless of the static permission it asked during installation. \item $\exists_{x} (trans(x, sink) \land \neg system(x) \land \neg hasPermissionToSink(x)).$ This policy says that transitive calls to a sink from non-system apps are forbidden, unless the source of the calls already has permission to the sink. This is then a simple privilege escalation detection (for non-system apps). \item $\exists_{x} (trans(x, sink) \wedge \neg system(x) \wedge \neg trusted(x)).$ This is further refinement to the policy in that we also give the user privilege to decide for themselves dynamically whether or not to trust an application. Untrusted apps can not make transitive call to the sink, but trusted apps are allowed, regardless of their permissions. \item $\exists_{x} (trans(x, internet) \wedge \neg system(x) \wedge \neg trusted(x) \wedge \diamonddot(call(x, contact))).$ This policy allows privilege escalation by non-trusted apps as long as there is no potential for data leakage through the sink. That is, as soon as a non-system and untrusted app accesses contact, it will be barred from accessing the internet. Note that the use of non-metric operator $\diamonddot$ ensures that the information that a particular app has accessed contact is persistent. \end{enumerate} \section{Implementation} \label{implementation} \begin{table}[t] \caption{Performance Table} \label{performanceTable} \centering \small{ \begin{tabular}{ c c c c} \hline\hline Policy & Uncached & Cached\\ \hline 1 & 766.4 & 143.6\\ 2 & 936.5 & 423.6\\ 3 & 946.8 & 418.3\\ 4 & 924.3 & 427.5\\ No Monitor & 758 & 169\\ \hline \end{tabular} } \end{table} \begin{figure}[t] \centering \subfloat[]{\includegraphics[trim=0 0 0 15, clip, width=0.5\textwidth]{Cached-gray.pdf}} \subfloat[]{\includegraphics[trim=0 0 0 15, clip, width=0.5\textwidth]{Uncached-gray.pdf}} \caption{Timing of Calls} \label{callTiming} \end{figure} We have implemented the monitoring algorithm presented in the previous section in Android 4.1. Some modifications to the application framework and the underlying Linux kernel in Android are neccessary to ensure our monitor can effectively monitor and stop unwanted behaviours. We have tested our implementation in both Android emulator and an actual device (Samsung Galaxy Nexus phone). We give a brief overview of our implementation; details of the monitor code generators and the modified Android kernel images are available from the authors' website. Our implementation consists of two parts: the codes that generate a monitor given a policy specification, and the modifications of Android framework and its Linux kernel to hook our monitor and to intercepts IPCs and access to Android resources. To improve runtime performance, the monitor generation is done outside Android; it produces C codes that are then compiled into a kernel module, and inserted into Android boot image. The monitor generator takes an input policy, encoded in an XML format extending that of RuleML. The monitor generator works by following the logic of the monitoring algorithm presented in Section~\ref{monitor}. It takes a policy formula $\phi$, and generates the required data structures and determine an ordering between elements of $Sub^*(\phi)$ as described earlier, and produces the codes illustrated in Algorithm~\ref{Init_Algorithm}, \ref{Iter_Algorithm} and \ref{Monitor_Algorithm}. The main body of our monitor lies in the Linux kernel space as a kernel module. The reason for this is that there are some cases where Android leaves the permission checking to the Linux kernel layer e.g., for opening network socket. However, to monitor the IPC events between Android components and apps, we need to place a hook inside the application framework. The IPC between apps is done through passing a data structure called {\em Intent}, which gets broken down into {\em parcels} before they are passed down to the kernel level to be delivered. So intercepting these parcels and reconstructing the original Intent object in the kernel space would be more difficult and error prone. The events generated by apps or components will be passed down to the monitor in the kernel, along with the application's user id. If the event is a call to the sink, then depending on the policy that is implemented in the monitor, it will decide to whether block or allow the call to proceed. We do this through our custom additional system calls to the Linux kernel which goes to this monitor. Our implementation places hooks in four services, namely accessing internet, sending SMS, accessing location, and accessing contact database. For each of this sink, we add a virtual UID in the monitor and treat it as a component of Android. We currently track only IPC calls through the Intent passing mechanism. This is obviously not enough to detect all possible communications between apps, e.g., those that are done through file systems, or side channels, such as vibration setting (e.g., as implemented in SoundComber~\cite{Soundcomber}), so our implementation is currently more of a proof of concept. In the case of SoundComber, our monitor can actually intercepts the calls between colluding apps, due to the fact that they utilise intent broadcast through IPC to synchronize data transfer. We have implemented some apps to test policies we mentioned in Section~\ref{examples}. In Table \ref{performanceTable} and Figure \ref{callTiming}, we provide some measurement of the timing of the calls between applications. The policy numbers in Table~\ref{performanceTable} refer to the policies in Section~\ref{examples}. To measure the average time for each IPC call, we construct a chain of ten apps, making successive calls between them, and measure the time needed for one end to reach the other. We measure two different average timings in miliseconds (ms) for different scenarios, based on whether the apps are in the background cache (i.e., suspended) or not. We also measure time spent on the monitor actually processing the event, which are around 1 ms for policy 1, and around 10 ms for the other three policies. This shows that the time spent in processing the event is quite low, but more overhead comes from the space required to process the event (there is a big jump in overall timing from simple rules with at most 2 free variables to more complex one with 3 free variables). Figure \ref{callTiming} shows that the timing of calls over time for each policy are roughly the same. This backs our claim that even though our monitor implements history-based access control, its performance does not depend on the size of the history. \section{Introduction} Android is a popular mobile operating system (OS) that has been used in a range of mobile devices such as smartphones and tablet computers. It uses Linux as the kernel, which is extended with an application framework (middleware). Most applications of Android are written to run on top of this middleware, and most of Android-specific security mechanisms are enforced at this level. Android treats each application as a distinct user with a unique user ID. At the kernel level, access control is enforced via the standard Unix permission mechanism based on the user id (and group id) of the app. At the middleware level, each application is sandboxed, i.e., it is running in its own instance of Dalvik virtual machine, and communication and sharing between apps are allowed only through an inter-process communication (IPC) mechanism. Android middleware provides a list of resources and services such as sending SMS, access to contacts, or internet access. Android enforces access control to these services via its permission mechanism: each service/resource is associated with a certain unique permission tag, and each app must request permissions to the services it needs at installation time. Everytime an app requests access to a specific service/resource, Android runtime security monitor checks whether the app has the required permission tags for that particular service/resource. A more detailed discussion of Android security architecture can be found in \cite{Enck09}. One problem with Android security mechanism is the problem of {\em privilege escalation}, that is, the possibility of an app to gain access to services or resources that it does not have permissions to access. Obviously privilege escalation is a common problem of every OS, e.g., when a kernel bug is exploited to gain root access. However, in Android, privilege escalation is possible even when apps are running in the confine of Android sandboxes~\cite{defcon18,Davi10,LastPE}. There are two types of attacks that can lead to privilege escalation~\cite{LastPE}: the {\em confused deputy attack} and the {\em collusion attack}. In the confused deputy attack, a legitimate app (the deputy) has permissions to certain services, e.g., sending SMS, and exposes an interface to this functionality without any guards. This interface can then be exploited by a malicious app to send SMS, even though the malicious app does not have the permission. Recent studies \cite{defcon18,Grace12NDSS,DroidChecker} show some system and consumer apps expose critical functionalities that can be exploited to launch confused deputy attacks. The collusion attack requires two or more malicious apps to collaborate. We have yet to encounter such a malware, either in the Google Play market or in the third party markets, although a proof-of-concept malware with such properties, called SoundComber~\cite{Soundcomber}, has been constructed. Several security extensions to Android have been proposed to deal with privilege escalation attacks~\cite{QUIRE,IPCInspection,LastPE}. Unlike these works, we aim at designing a high-level policy language that is expressive enough to capture privilege escalation attacks, but is also able to express more refined policies (see Section~\ref{examples}). Moreover, we aim at designing a lightweight monitoring framework, where policy specifications can be modified easily and enforced efficiently. Thus we aim at an automated generation of security monitors that can efficiently enforce policies written in our specification language. On the specific problem of detecting privilege escalation, it is essentially a problem of tracking (runtime) control flow, which is in general a difficult problem and would require a certain amount static analysis~\cite{Denning,TaintDroid}. So we adopt a `lightweight' heuristic to ascertain causal dependency between IPC calls: we consider two successive calls, say from A to B, followed by a call from B to C, as causally dependent if they happen within a certain reasonably short time frame. This heuristic seems sensible in the presence of confused deputy attacks, but can be of course circumvented by colluding attacks. For the latter there is probably not a general solution that can be effective in all cases, e.g., when covert channels are involved~\cite{Soundcomber}, so we shall restrict to addressing the former. The core of our policy language, called RMTL, is essentially a past-fragment of metric linear temporal logic (MTL)~\cite{Alur90lics,Thati05MTL,Basin08FSTTCS}. We consider only the fragment of MTL with past-time operators, as this is sufficient for our purpose to enforce history-sensitive access control. This also means that we can only express safety properties~\cite{Lichtenstein}, but not policies capturing obligations as in, e.g., \cite{Basin08FSTTCS}. Temporal operators are useful in this setting to enforce access control on apps based on histories of their executions; see Section~\ref{examples}. Such a history-dependent policy cannot be expressed in the policy languages used in \cite{QUIRE,IPCInspection,LastPE}. MTL by itelf is, however, insufficient to express transitive closures of relations, which is needed to specify an IPC call chains between apps, among others. To deal with this, we extend MTL with a recursive definitions, e.g., one would be able to write a definition such as: \begin{equation} \label{eq:trans} trans(x,y) := call(x,y) \lor \exists z. \diamonddot_n trans(x,z) \land call(z,y), \end{equation} where $call$ denotes the IPC event. This equation defines $trans$ the reflexive-transitive closure of $call.$ The metric operator $\diamonddot_n \phi$ means intuitively $\phi$ holds within $n$ time units in the past; we shall see a more precise definition of the operators in Section~\ref{logic}. Readers familiar with modal $\mu$-calculus~\cite{Bradfield07handbook} will note that this is but a syntactic sugar for $\mu$-expressions for (least) fixed points. To be practically enforceable in Android, RMTL monitoring algorithm must satisfy an important constraint, i.e., the algorithm must be {\em trace-length independent}. This is because the number of events generated by Android can range in the thousands per hour, so if the monitor must keep all the events generated by Android, its performance will degrade significantly overtime. Another practical consideration also motivates a restriction to metric operators that we adopt in RMTL. More specifically, MTL allows a metric version of the `since' operator of the form $\phi_1 \mathrel{\mathbb{S}}_{[m,n)} \phi_2$, where $[m, n)$ specifies a half-closed (discrete) time interval from $m$ to $n$. The monitoring algorithm for MTL in \cite{Thati05MTL} works by first expanding this formula into formulas of the form $\phi_1 \mathrel{\mathbb{S}}_{[m',n')} \phi_2$ where $[m',n')$ is any subinterval of $[m,n)$. A similar expansion is also used implicitly in monitoring for first-order MTL in \cite{Basin08FSTTCS}, i.e., in their incremental automatic structure extension in their first-order logic encoding for the `since' and `until' operators. In general, if we have $k$ nested occurrences of metric operators, each with interval $[m,n)$, the number of formulas produced by this expansion is bounded by $O((\frac{(n-m) \times (n - m + 1)}{2})^{k})$. In Android, event timestamps are in milliseconds, so this kind of expansion is not practically feasible. For example, suppose we have a policy that monitors three successive IPC calls that happen within 10 seconds between successive calls. This requires two nested metric operators with intervals $[0,10^4)$ to specify. The above naive expansion would produce around $25 \times 10^{14}$ formulas, and assuming the truth value of each formula is represented with 1 bit, this would require more than $30$ GB of storage to store all their truth values, something which is beyond the capacity of most smartphones today. An improvement to this exponential expansion is proposed in \cite{Basin11RV,Reinbacher12RV}, where one keeps a sequence of timestamps for each metric temporal operator occuring in the policy. This solution, although avoids the exponential expansion, is strictly speaking not trace-length independent. This solution seems optimal so it is hard to improve it without further restriction to the policy language. We show that, if one restricts the intervals of metric operators to the form $[0,n)$, one only needs to keep one timestamp for each metric operator in monitoring; see Section~\ref{monitor}. To summarise, our contributions are as follows: \begin{enumerate} \item In terms of results in runtime verification, our contribution is in the design of a new logic-based policy language that extends MTL with recursive definitions, that avoids exponential expansion of metric operators, and for which the policy enforcement is trace-length independent. \item In terms of the application domain, ours is the first implementation of a logic-based runtime security monitor for Android that can enforce history-based access control policies, including those that concern privilege escalations. Our monitoring framework can express temporal and metric-based policies not possible in existing works \cite{QUIRE,IPCInspection,LastPE,bauer2013rv}. \end{enumerate} The rest of the paper is organized as follows. Section~\ref{logic} introduces our policy language RMTL. Some example policies are described in \ref{examples}. Section~\ref{implementation} discusses our implementation of the monitors for RMTL, and the required modification of Android OS kernel to integrate our monitor into the OS. In Section~\ref{conclusion} we conclude and discuss related and future works. Some proofs of the lemmas and theorems can be found in the appendix. Details of the implementation of the monitor generator and the binaries of the modified Android OS are available online.\footnote{ \url{http://users.cecs.anu.edu.au/\~{}hengunadi}/LogicDroid.html.} \paragraph{Acknowledgment} This work is partly supported by the Australian Research Council Discovery Grant DP110103173. \section{The policy specification language RMTL} \label{logic} Our policy specification language, which we call RMTL, is based on an extension of metric linear temporal logic (MTL)~\cite{MTL}. The semantics of LTL~\cite{Pnueli77} is defined in terms of models which are sequences of states (or worlds). In our case, we restrict to finite sequences of states. MTL extends LTL models by adding timestamps to each state, and adding temporal operators that incorporate timing constraints, e.g., MTL features temporal operators such as $\diamond_{[0,3)} \phi$ which expresses that $\phi$ holds in some state in the future, and the timestamp of that world is within 0 to 3 time units from the current timestamp. We restrict to a model of MTL that uses discrete time, i.e., timestamps in this case are non-negative integers. We shall also restrict to the past-time fragment of MTL. We extend MTL with two additional features: first-order quantifiers and recursive definitions. Our first-order language is a multi-sorted language. For this paper, we consider only two sorts, which we call $prop$ (for `properties') and $app$ (for denoting applications). Sorts are ranged over by $\alpha$. We first fix a {\em signature} $\Sigma$ for our first-order language, which is used to express terms and predicates of the language. We shall consider only constant symbols and predicate symbols, but no function symbols. We distinguish two types of predicate symbols: {\em defined} predicates and {\em undefined} ones. The defined predicate symbols are used to write recursive definitions and to each of such symbols we associate a formula as its definition. We shall come back to this shortly. Constant symbols are ranged over by $a$, $b$ and $c$, undefined predicate symbols are ranged over by $p$, $q$ and $r$, and defined predicate symbols are ranged over by $P$, $Q$ and $R.$ We assume an infinite set of sorted variables $\mathcal{V}$, whose elements are ranged over by $x$, $y$ and $z.$ We sometimes write $x_\alpha$ to say that $\alpha$ is the sort of variable $x.$ A {\em $\Sigma$-term} is either a constant symbol $c\in\Sigma$ or a variable $x\in \mathcal{V}$. We use $s$, $t$ and $u$ to range over terms. To each symbol in $\Sigma$ we associate a sort information. We shall write $c: \alpha$ when $c$ is a constant symbol of sort $\alpha.$ A predicate symbol of arity $n$ has sort of the form $\alpha_1 \times \cdots \times \alpha_n$, and such a predicate can only be applied to terms of sorts $\alpha_1,\dots,\alpha_n.$ Constant symbols are used to express things like permissions in the Android OS, e.g., reading contacts, sending SMS, etc., and user ids of apps. Predicate symbols are used express events such as IPC calls between apps, and properties of an app, such as whether it is a system app, a trusted app (as determined by the user). As standard in first-order logic (see e.g. \cite{fitting}), the semantics of terms and predicates are given in terms of a first-order structure, i.e., a set $\mathcal{D}_\alpha$, called a {\em domain}, for each sort $\alpha$, and an interpretation function $I$ assigning each constant symbol $c : \alpha \in \Sigma$ an element of $c^I \in \mathcal{D}_\alpha$ and each predicate symbol $p : \alpha_1 \times \cdots \times \alpha_n \in \Sigma$ an $n$-ary relation $p^I \subseteq \mathcal{D}_{\alpha_1} \times \cdots \times \mathcal{D}_{\alpha_n}.$ We shall assume constant domains in our model, i.e., every world has the same domain. The formulas of RMTL is defined via the following grammar: $$ \begin{array}{ll} F := & \bot \mid p(t_1,\dots,t_m) \mid P(t_1,\dots,t_n) \mid F \lor F \mid \neg F \mid \bullet F \mid F \mathrel{\mathbb{S}} F \mid \blacklozenge F \mid \diamonddot F \mid \\ & \bullet_{n} F \mid F \mathrel{\mathbb{S}}_{n} F \mid \blacklozenge_{n} F \mid \diamonddot_{n} F \mid \exists_\alpha x.F\\ \end{array} $$ where $m$ and $n$ are natural numbers. The existential quantifier is annotated with a sort information $\alpha$. For most of our examples and applications, we only quantify only over variables of sort $app$. The operators indexed by $n$ are {\em metric temporal operators}. The $n \geq 1$ here denotes the interval $[0,n)$, so these are special cases of the more general MTL operators in \cite{MTL}, where intervals can take the form $[m,n)$, for $n \geq m \geq 0.$ We use $\phi$, $\varphi$ and $\psi$ to range over formulas. We assume that unary operators bind stronger than the binary operators, so $\bullet \phi \lor \psi$ means $(\bullet \phi) \lor \psi$. We write $\phi(x_1,\dots,x_n)$ to denote a formula whose free variables are among $x_1,\dots,x_n$. Given such a formula, we write $\phi(t_1,\dots,t_n)$ to denote the formula obtained by replacing $x_i$ with $t_i$ for every $i \in \{1,\dots,n\}$. To each defined predicate symbol $P : \alpha_1 \times \cdots \times \alpha_n$, we associate a formula $\phi_P$, which we call the {\em definition} of $P$. Notationally, we write $P(x_1,\dots,x_n) := \phi_p(x_1,\dots,x_n).$ We require that $\phi_P$ is {\em guarded}, i.e., every occurrence of any recursive predicate $Q$ in $\phi_P$ is prefixed by either $\bullet$, $\bullet_m$, $\diamonddot$ or $\diamonddot_n$. This guardedness condition is important to guarantee termination of recursion in model checking. Given the above logical operators, we can define additional operators via their negation, e.g., $\top$ is defined as $\neg \bot$, $\phi \land \psi$ is defined as $\neg (\neg \phi \lor \neg \psi)$, $\phi \rightarrow \psi$ is defined as $\neg \phi \lor \psi$, and $\forall_\alpha x.\phi$ is defined as $\neg (\exists_\alpha x.\neg \phi)$, etc. Before proceeding to the semantics of RMTL, we first define a well-founded ordering on formulae of RMTL, which will be used later. \begin{definition} \label{def:order} We define a relation $\prec_S$ on the set RMTL formulae as the smallest relation satisfying the following conditions: \begin{enumerate} \item For any formula $\phi$ of the form $p(\vec t)$, $\bot$, $\bullet \psi$, $\bullet_n \psi$, $\diamonddot \psi$ and $\diamonddot_n \psi$, there is no $\phi'$ such that $\phi' \prec_S \phi$. \item For every recursive definition $P(\vec x) := \phi_P(\vec x)$, we have $\phi_P(\vec t) \prec_S P(\vec t)$ for every terms $\vec t.$ \item $\psi \prec_S \psi \lor \psi'$, $\psi \prec_S \psi' \lor \psi$, $\psi \prec_S \neg \psi$, and $\psi \prec_S \exists x.\psi$. \item $\psi_i \prec_S \psi_1 \mathrel{\mathbb{S}} \psi_2$, and $\psi_i \prec_S \psi_1 \mathrel{\mathbb{S}}_n \psi_2$, for $i \in \{1,2\}$ \end{enumerate} We denote with $\prec$ the reflexive and transitive closure of $\prec_S.$ \end{definition} \begin{lemma} \label{lm:order} The relation $\prec$ on RMTL formulas is a well-founded partial order. \end{lemma} For our application, we shall restrict to finite domains. Moreover, we shall restrict to an interpretation $I$ which is injective, i.e., mapping every constant $c$ to a unique element of $\mathcal{D}_{\alpha}.$ In effect we shall be working in the term model, so elements of $\mathcal{D}_\alpha$ are just constant symbols from $\Sigma.$ So we shall use a constant symbol, say $c : \alpha$, to mean both $c \in \Sigma$ and $c^I \in \mathcal{D}_\alpha$. With this fix interpretation, the definition of the semantics (i.e., the satisfiability relation) can be much simplified, e.g., we do not need to consider valuations of variables. A {\em state} is a set of undefined atomic formulas of the form $p(c_1,\dots,c_n)$. Given a sequence $\sigma$, we write $|\sigma|$ to denote its length, and we write $\sigma_i$ to denote the $i$-th element of $\sigma$ when it is defined, i.e., when $1 \leq i \leq |\sigma|.$ A {\em model} is a pair $(\pi,\tau)$ of a sequence of {\em states} $\pi$ and a sequence {\em timestamps}, which are natural numbers, such that $|\pi| = |\tau|$ and $\tau_i \leq \tau_j$ whenever $i\leq j$. Let $<$ denote the total order on natural numbers. Then we can define a well-order on pairs $(i,\phi)$ of natural numbers and formulas by taking the lexicographical ordering $(<, \prec)$. The satisfiability relation between a model $\rho = (\pi,\tau)$, a {\em world} $i \geq 1$ (which is a natural number) and a {\em closed} formula $\phi$ (i.e., $\phi$ contains no free variables), written $(\rho,i) \models \phi$, is defined by induction on the pair $(i, \phi)$ as follows, where we write $(\rho, i) \not \models \phi$ when $(\rho, i) \models \phi$ is false. \begin{itemize} \item $(\rho, i) \not \models \bot$ \item $(\rho, i) \models \neg \phi$ iff $(\rho, i) \not \models \phi$. \item $(\rho, i) \models p(c_1,\dots,c_n)$ iff $p(c_1,\dots,c_n) \in \pi_i.$ \item $(\rho, i) \models P(c_1,\dots,c_n)$ iff $(\rho,i) \models \phi(c_1,\dots,c_n)$ where $P(\vec x) := \phi(\vec x).$ \item $(\rho, i) \models \phi \lor \psi$ iff $(\rho, i) \models \phi$ or $(\rho,i) \models \psi$. \item $(\rho, i) \models \bullet \phi$ iff $i > 1$ and $(\rho, i-1) \models \phi.$ \item $(\rho, i) \models \blacklozenge \phi$ iff there exists $j \leq i$ s.t. $(\rho, j) \models \phi$. \item $(\rho, i) \models \diamonddot \phi$ iff $i > 1$ and there exists $j < i$ s.t. $(\rho, j) \models \phi.$ \item $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}} \phi_2$ iff there exists $j \leq i$ such that $(\rho, j) \models \phi_2$ and $(\rho,k)\models \phi_1$ for every $k$ s.t. $j < k \leq i.$ \item $(\rho, i) \models \bullet_n \phi$ iff $i > 1$, $(\rho, i-1) \models \phi$ and $\tau_i - \tau_{i-1} < n.$ \item $(\rho, i) \models \blacklozenge_n \phi$ iff there exists $j \leq i$ s.t. $(\rho, j) \models \phi$ and $\tau_i - \tau_j < n$. \item $(\rho, i) \models \diamonddot_n \phi$ iff $i > 1$ and there exists $j < i$ s.t. $(\rho, j) \models \phi$ and $\tau_i - \tau_j < n.$ \item $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ iff there exists $j \leq i$ such that $(\rho, j) \models \phi_2$, $(\rho,k)\models \phi_1$ for every $k$ s.t. $j < k \leq i$, and $\tau_i - \tau_j < n.$ \item $(\rho, i) \models \exists_\alpha x.\phi(x)$ iff there exists $c \in \mathcal{D}_\alpha$ s.t. $(\rho, i) \models \phi(c).$ \end{itemize} Note that due to the guardedness condition in recursive definitions, our semantics for recursive predicates is much simpler than the usual definition as in $\mu$-calculus, which typically involves the construction of a (semantic) fixed point operator. Note also that some operators are redundant, e.g., $\blacklozenge \phi$ can be defined as $\top \mathrel{\mathbb{S}} \phi$, and $\diamonddot \phi$ can be defined as $\bullet \blacklozenge \phi.$ This holds for some metric operators, e.g., $\blacklozenge_n \phi$ and $\diamonddot_n \phi$ can be defined as, respectively, $\top \mathrel{\mathbb{S}}_n \phi$ and \begin{equation} \label{eq:diamonddot} \diamonddot_n \phi = \bigvee_{i+j = n} \bullet_i \blacklozenge_j \phi \end{equation} This operator will be used to specify an active call chain, as we shall see later, so it is convenient to include it in our policy language. In the next section, we shall assume that $\blacklozenge$, $\diamonddot$, $\blacklozenge_n$ as derived connectives. Since we consider only finite domains, $\exists_\alpha x.\phi(x)$ can be reduced to a big disjunction $\bigvee_{c \in \mathcal{D}_\alpha} \phi(c)$, so we shall not treat $\exists$-quantifier explicitly. This can be problematic if the domain of quantification is big, as it suffers the same kind of exponential explosion as with the expansion of metric operators in MTL~\cite{Thati05MTL}. We shall defer the explicit treatment of quantifiers to future work. \section{Trace-length independent monitoring} \label{monitor} The problem of monitoring is essentially a problem of model checking, i.e., to decide whether $(\rho, i) \models \phi$, for any given $\rho = (\pi,\tau)$, $i$ and $\phi.$ In the context of Android runtime monitoring, a state in $\pi$ can be any events of interest that one would like to capture, e.g., the IPC call events, queries related to location information or contacts, etc. To simplify discussions, and because our main interest is in privilege escalation through IPC, the only type of event we consider in $\pi$ is the IPC event, which we model with the predicate $call : app \times app$. Given a policy specification $\phi$, a naive monitoring algorithm that enforces this policy would store the entire event history $\pi$ and every time a new event arrives at time $t$, it would check $(([\pi; e], [\tau; t]), |\rho| + 1) \models \phi.$ This is easily shown decidable, but is of course rather inefficient. In general, the model checking problem for RMTL (with finite domains) can be shown PSPACE hard following the same argument as in \cite{FOTL}. A design criteria of RMTL is that enforcement of policies does not depend on the length of history of events, i.e., at any time the monitor only needs to keep track of a fixed number of states. Following~\cite{bauer2013rv}, we call a monitoring algorithm that satisfies this property {\em trace-length independent.} For PTLTL, trace-length independent monitoring algorithm exists, e.g., the algorithm by Havelund and Rosu~\cite{DYNAMICLTL}, which depends only on two states in a history. That is, satisfiability of $(\rho, i+1) \models \phi$ is a boolean function of satisfiability of $(\rho, i + 1) \models \psi$, for every strict subformula $\psi$ of $\phi$, and satisfiability of $(\rho, i) \models \psi'$, for every subformula $\psi'$ of $\phi.$ This works for PTLTL because the semantics of temporal operators in PTLTL can be expressed in a recursive form, e.g., the semantics of $\mathrel{\mathbb{S}}$ can be equally expressed as~\cite{DYNAMICLTL}: $(\rho, i+1) \models \phi_1 \mathrel{\mathbb{S}} \phi_2$ iff $(\rho, i+1) \models \phi_2$, or $(\rho, i + 1) \models \phi_1$ and $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}} \phi_2.$ This is not the case for MTL. For example, satisfiability of the unrestricted `since' operator $\mathrel{\mathbb{S}}_{[m,n)}$ can be equivalently expressed as: \begin{equation} \begin{array}{ll} (\rho, i+1) \models \phi_1 \mathrel{\mathbb{S}}_{[m,n)} \phi_2 \mbox{ iff } & m = 0, n > 1, \mbox{ and } (\rho, i+1) \models \phi_2, \mbox{ or } \\ & (\rho, i+1) \models \phi_1 \mbox{ and } (\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_{[m', n')} \phi_2 \end{array} \label{eq:since} \end{equation} where $m' = min(0, m - \tau_{i+1} + \tau_i)$ and $n' = min(0, n - \tau_{i+1} + \tau_i).$ Since $\tau_{i+1}$ can vary, the value of $m'$ and $n'$ can vary, depending on the history $\rho.$ We avoid the expansion of metric operators in monitoring by restricting the intervals in the metric operators to the form $[0, n).$ We show that clause (\ref{eq:since}) can be brought back to a purely recursive form. The key to this is the following lemma: \begin{lemma}[Minimality] \label{lm:minimal} If $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ ($(\rho, i) \models \diamonddot_n \phi$) then there exists an $m \leq n$ such that $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_m \phi_2$ (resp. $(\rho, i) \models \diamonddot_m \phi$) and such that for every $k$ such that $0 < k < m$, we have $(\rho, i) \not \models \phi_1 \mathrel{\mathbb{S}}_k \phi_2$ (resp., $(\rho, i) \not \models \diamonddot_k \phi$). \end{lemma} Given $\rho$, $i$ and $\phi$, we define a function $\mathfrak{m}$ as follows: $$ \mathfrak{m}(\rho,i,\phi) = \left\{ \begin{array}{l} m, \mbox{ if $\phi$ is either $\phi_1 \mathrel{\mathbb{S}}_n \phi_2$ or $\diamonddot_n \phi'$ and $(\rho,i) \models \phi$, } \\ 0, \mbox{ otherwise. } \end{array} \right. $$ where $m$ is as given in Lemma~\ref{lm:minimal}; we shall see how its value is calculated in Algorithm~\ref{Iter_Algorithm}. The following theorem follows from Lemma~\ref{lm:minimal}. \begin{theorem}[Recursive forms] \label{thm:recursive-form} For every model $\rho$, every $n \geq 1$, $\phi$, $\phi_1$ and $\phi_2$, and every $1 < i \leq |\rho|$, the following hold: \begin{enumerate} \item $(\rho, i) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ iff $(\rho, i) \models \phi_2$, or $(\rho, i) \models \phi_1$ and $(\rho, i-1) \models \phi_1 \mathrel{\mathbb{S}}_n \phi_2$ and $n - (\tau_i - \tau_{i-1}) \geq \mathfrak{m}(\rho, i-1, \phi_1 \mathrel{\mathbb{S}}_n \phi_2).$ \item $(\rho, i) \models \diamonddot_n \phi$ iff $(\rho, i - 1) \models \phi$ and $\tau_i - \tau_{i-1} < n$, or $(\rho, i-1) \models \diamonddot_n \phi$ and $n - (\tau_i - \tau_{i-1}) \geq \mathfrak{m}(\rho, i-1, \diamonddot_n \phi).$ \end{enumerate} \end{theorem} Given Theorem~\ref{thm:recursive-form}, the monitoring algorithm for PTLTL in \cite{DYNAMICLTL} can be adapted, but with an added data structure to keep track of the function $\mathfrak{m}.$ In the following, given a formula $\phi$, we assume that $\exists$, $\blacklozenge$ and $\diamonddot$ have been replaced with its equivalent form as mentioned in Section~\ref{logic}. Given a formula $\phi$, let $Sub(\phi)$ be the set of subformulas of $\phi$. We define a closure set $S^*(\phi)$ of $\phi$ as follows: Let $Sub^0(\phi) = Sub(\phi)$, and let $$ Sub^{n+1}(\phi) = Sub_n(\phi) \cup \{Sub(\phi_P(\vec c)) \mid P(\vec c) \in Sub_n(\phi), \mbox{ and } P(\vec x) := \phi_P(\vec x)\} $$ and define $Sub^*(\phi) = \bigcup_{n\geq 0} Sub^n(\phi).$ Since $\mathcal{D}_\alpha$ is finite, $Sub^*(\phi)$ is finite, although its size is exponential in the arities of recursive predicates. For our specific applications, the predicates used in our sample policies have at most arity of two (for tracking transitive calls), so this is still tractable. In future work, we plan to investigate ways of avoiding this explicit expansion of recursive predicates. We now describe how monitoring can be done for $\phi$, given $\rho$ and $1 \leq i \leq |\rho|.$ We assume implicitly a preprocessing step where we compute $Sub^*(\phi)$; we do not describe this step here but it is quite straightforward. Let $\phi_1, \phi_2, \ldots, \phi_m$ be an enumeration of $Sub^*(\phi)$ respecting the partial order $\prec$, i.e., if $\phi_i \prec \phi_j$ then $i \leq j.$ Then we can assign to each $\psi \in Sub^*(\phi)$ an index $i$, s.t., $\psi = \phi_i$ in this enumeration. We refer to this index as $idx(\psi).$ We maintain two boolean arrays $prev[1,\dots,m]$ and $cur[1,\dots,m].$ The intention is that given $\rho$ and $i > 1$, the value of $prev[k]$ corresponds to the truth value of the judgment $(\rho, i-1) \models \phi_k$ and the truth value of $cur[k]$ corresponds to the truth value of the judgment $(\rho, i) \models \phi_k.$ We also maintain two integer arrays $mprev[1,\dots,m]$ and $mcur[1,\dots,m]$ to store the value of the function $\mathfrak{m}.$ The value of $mprev[k]$ corresponds to $\mathfrak{m}(\rho, i-1, \phi_k)$, and $mcur[k]$ corresponds to $\mathfrak{m}(\rho, i, \phi_k).$ Note that this preprocessing step only needs to be done once, i.e., when generating the monitor codes for a particular policy, which is done offline, prior to inserting the monitor into the operating system kernel. The main monitoring algorithm is divided into two subprocedures: the initialisation procedure (Algorithm~\ref{Init_Algorithm}) and the iterative procedure (Algorithm~\ref{Iter_Algorithm}). The monitoring procedure (Algorithm~\ref{Monitor_Algorithm}) is then a simple combination of these two. We overload some logical symbols to denote operators on boolean values. In the actual implementation, we do not actually implement the loop in Algorithm~\ref{Monitor_Algorithm}; rather it is implemented as an event-triggered procedure, to process each event as they arrive using $Iter$. {\small \begin{algorithm}[t] \caption{$Monitor(\rho, i, \phi)$} \label{Monitor_Algorithm} \begin{algorithmic} \STATE $Init(\rho,\phi,prev,cur,mprev,mcur)$ \FOR {$j = 1$ to $i$} \STATE $Iter(\rho,j,\phi,prev,cur,mprev,mcur);$ \ENDFOR \RETURN $cur[idx(\phi)];$ \end{algorithmic} \end{algorithm} } {\small \begin{algorithm}[t] \caption{$Init(\rho, \phi, prev, cur, mprev, mcur)$} \label{Init_Algorithm} \begin{algorithmic} \FOR {$k=1,\dots,m$} \STATE $prev[k] := false$, $mprev[k] := 0$ and $mcur[k] := 0;$ \ENDFOR \FOR {$k=1,\dots,m$} \SWITCH {$\phi_k = \bot$} \CASELINE{$\bot$}{$cur[k] := false;$} \CASELINE {$p(\vec c)$}{$cur[k] := p(\vec c) \in \pi_1;$} \CASELINE {$P(\vec c)$}{$cur[k] := cur[idx(\phi_P(\vec c))];$ \COMMENT{Suppose $P(\vec x) := \phi_P(\vec x).$}} \CASELINE {$\neg \psi$}{$cur[k] := \neg cur[idx(\psi)];$} \CASELINE {$\psi_1 \lor \psi_2$}{$cur[k] := cur[idx(\psi_1)] \lor cur[idx(\psi_2)];$} \CASELINE {$\bullet \psi$}{$cur[k] := false;$} \CASELINE {$\diamonddot \psi$}{$cur[k] := false;$} \CASELINE {$\psi_1 \mathrel{\mathbb{S}} \psi_2$}{$cur[k] := cur[idx(\psi_2)];$} \CASELINE {$\bullet_n \psi$}{$cur[k] := false;$} \CASELINE {$\diamonddot_n \psi$}{$cur[k] := false; mcur[k] := 0;$} \CASE {$\phi_k = \psi_1 \mathrel{\mathbb{S}}_n \psi_2$} \STATE $cur[k] := cur[idx(\psi_2)];$ \IFLINE {$cur[k] = true$}{$mcur[k] := 1;$} \ELSELINE {$mcur[k] := 0;$ } \ENDIFLINE \end{ALC@g} \STATE \textbf{end switch} \ENDFOR \RETURN $cur[idx(\phi)];$ \end{algorithmic} \end{algorithm} } {\small \begin{algorithm}[t] \caption{$Iter(\rho, i, \phi, prev, cur, mprev, mcur)$} \label{Iter_Algorithm} \begin{algorithmic} \REQUIRE {$i > 1.$} \STATE $prev := cur;$ $mprev := mcur;$ \FORLINE {$k = 1$ to $m$}{$mcur[k] := 0;$} \FOR {$k = 1$ to $m$} \SWITCH{$\phi_k$} \CASELINE {$\bot$}{$cur[k] := false$;} \CASELINE {$p(\vec c)$}{$cur[k] := p(\vec c) \in \pi_i$;} \CASELINE {$\neg \psi$}{$cur[k] := \neg cur[idx(\psi)];$} \CASELINE {$P(\vec c)$}{$cur[k] := cur[idx(\phi_P(\vec c))];$ \COMMENT{Suppose $P(\vec x) := \phi_P(\vec x).$}} \CASELINE {$\psi_1 \lor \psi_2$}{$cur[k] := cur[idx(\psi_1)] \lor cur[idx(\psi_2)];$} \CASELINE{$\bullet \psi$}{$cur[k] := prev[idx(\psi)];$} \CASELINE{$\diamonddot \psi$}{$curr[k] := prev[idx(\psi)] \lor prev[\diamonddot \psi];$} \CASELINE{$\psi_1 \mathrel{\mathbb{S}} \psi_2$}{$cur[k] := cur[idx(\psi_2)] \lor (cur[idx(\psi_1)] \land prev[idx(\psi_2)]);$} \CASELINE{$\bullet_n \psi$}{$cur[k] := prev[\psi] \land (\tau_i - \tau_{i-1} < n);$} \CASE {$\diamonddot_n \psi$} \STATE $l := prev[idx(\psi)] \land (\tau_i - \tau_{i-1} < n);$ \STATE $r := prev[idx(\diamonddot_n \psi)] \land (n - (\tau_i - \tau_{i-1}) \geq mprev[k]));$ \STATE $cur[k] := l \lor r;$ \IFLINE {$l$}{$mcur[k] := \tau_i - \tau_{i-1} + 1;$} \ELSIFLINE {$r$}{$mcur[k] := mprev[k] + \tau_i - \tau_{i-1};$} \ELSELINE{$mcur[k] := 0;$} \ENDIFLINE \end{ALC@g} \CASE {$\psi_1 \mathrel{\mathbb{S}}_n \psi_2$} \STATE $l := cur[idx(\psi_2)];$ \STATE $r := cur[idx(\psi_1)] \land prev[k] \land (n - (\tau_i - \tau_{i-1}) \geq mprev[k]);$ \STATE $cur[k] := l \lor r;$ \IFLINE {$l$}{$mcur[k] := 1;$} \ELSIFLINE {$r$}{$mcur[k] := mprev[k] + \tau_i - \tau_{i-1};$} \ELSELINE{$mcur[k] := 0;$ } \ENDIFLINE \end{ALC@g} \STATE \textbf{end switch} \ENDFOR \RETURN $cur[idx(\phi)];$ \end{algorithmic} \end{algorithm} } \begin{theorem} $(\rho, i) \models \phi$ iff $Monitor(\rho,i,\phi)$ returns true. \end{theorem} The $Iter$ function only depends on two worlds: $\rho_i$ and $\rho_{i-1}$, so the algorithm is trace-length independent. In principle there is no upperbound to its space complexity, as the timestamp $\tau_i$ can grow arbitrarily large, as is the case in \cite{Basin11RV}. Practically, however, the timestamps in Android are stored in a fixed size data structure, so in such a case, when the policy is fixed, the space complexity is constant (i.e., independent of the length of history $\rho$).
\section{Introduction} A polymer attached to a wall produces a force because of the loss of entropy on the wall. This has been measured experimentally \cite{bijsterbosch1995a-a,carignano1995a-a,currie2003a-a} and recently described theoretically in two dimensions using lattice walk models \cite{jensen2013a-a,janse2013a-a}. There has also been work concerning the entropic pressure of a polymer in the bulk \cite{gassoumov2013a-a}. Lattice walks and polygons on two dimensional lattices have in the past been utilised to model simple vesicles \cite{leibler1987a-a,brak1990a-a,fisher1991a-a,owczarek1993a-:a,brak1994a-:a}, where there can be an internal pressure. Here we explore the competition between the bulk internal pressure and the point pressure caused by entropy loss when a vesicle is fixed to a wall in two dimensions. Our study involves an exactly solved model of vesicles \cite{owczarek1993a-:a,prellberg1995c-:a}, namely, area-weighted Dyck paths \cite{janse2000a-a,owczarek2010c-:a,owczarek2012b-:a}. Let $Z_N$ be the partition function of some lattice model of rooted configurations of size $N$, for example directed or undirected self-avoiding walks or self-avoiding polygons, and let $Z_N^{(Q)}$ be the partition function conditioned on configurations avoiding a chosen point $Q$ in the lattice. Then the pressure on the point $Q$ is given by the difference of the finite-size free energies $-\log Z_N$ and $-\log Z_N^{(Q)}$, that is \begin{equation} \label{generalpressure} P_N^{(Q)}=-\log Z_N^{(Q)}+\log Z_N\;. \end{equation} Here we take $k_BT=1$ for convenience. When the configurations are Dyck paths, which are directed paths above the diagonal of a square lattice starting at the origin and ending on the diagonal, this model was analysed in \cite{janse2013a-a}. The pressure at the point $Q=(m,m)$ for walks of length $2N$ is given exactly as \begin{equation} \label{catalanpressure} P_N^{(m)}=-\log\left(1-\frac{C_mC_{N-m}}{C_N}\right)\;, \end{equation} where $C_k=\frac1{k+1}\binom{2k}k$ is the $k$-th Catalan number counting $2k$-step Dyck paths. For $N$ and $m$ large, this leads to \begin{equation} \label{Dyckpathprofile} P_N^{(m)}=\frac1{\sqrt{\pi x^3(1-x)^3}}\cdot\frac1{N^{3/2}}+O(N^{-5/2})\;, \end{equation} where $x=m/N$ measures the relative distance of the point $Q$ from the origin with respect to the length of the walk. That is, the pressure of the Dyck path decays to zero as $N^{-3/2}$ in the centre of the Dyck path, with an $x$-dependent profile, as shown in Figure \ref{profile}. \begin{figure}[ht] \begin{center} \includegraphics[width=0.4\textwidth]{Pressureprofile.pdf} \end{center} \caption{Pressure profile for long Dyck paths as a function of the relative distance $x$ from the point of tether. \label{profile} } \end{figure} In contrast, near the boundary the pressure tends to an $N$-independent limiting value \begin{equation} P_N^{(m)}\rightarrow-\log\left(1-\frac{C_m}{4^m}\right)\;. \end{equation} \section{The model} In an extension of the work described in the introduction, the configurations of the model studied here are area-weighted Dyck paths, where we have rotated the lattice by $45^{\circ}$ for convenience as shown in Figure \ref{vesicle}. To be more precise, we weight each full square plaquette between the Dyck path and the surface with a weight $q=\exp(\Pi)$, where $\Pi$ is the osmotic pressure. We use these paths to model vesicles adsorbed at the surface, that is to say that we consider them as vesicles with the bottom part of the membrane firmly attached to the surface. As described above, to calculate the pressure we need to consider a slightly modified set of configurations that avoid some point. For our vesicle model this means that the bottom of the vesicle does not include a particular plaquette, see Figure \ref{vesicle}. \begin{figure}[ht] \begin{center} \includegraphics[width=0.9\textwidth]{q-dyck-sketch.pdf} \end{center} \caption{A vesicle of length 9 given by a Dyck path with 18 steps, enclosing the 11 shaded plaquettes and having weight $q^{11}$. When considering the modified configurations that avoid the lighter coloured plaquette, which is at distance $m=6$, the vesicle is considered to enclose 10 plaquettes and have weight $q^{10}$. \label{vesicle} } \end{figure} Denoting the set of all $2N$-step Dyck paths by ${\cal D}_N$, the partition function for unrestricted vesicles of length $N$ is given by \begin{equation} Z_N(q)=\sum_{\varphi\in{\cal D}_N}q^{A(\varphi)}\;, \end{equation} where $A(\varphi)$ is the number of plaquettes enclosed by the configuration $\varphi$. Similarly, denoting the set of all $2N$-step Dyck paths enclosing the surface plaquette at distance $1\leq m\leq N-1$ by ${\cal D}_N^{(m)}\subset{\cal D}_N$, the partition function for the restricted vesicles is given by \begin{equation} Z_N^{(m)}(q)=\sum_{\varphi\in{\cal D}_N^{(m)}}q^{A(\varphi)-1}\;. \end{equation} Note that the distance $m$ is measured as the number of half-plaquettes along the surface to the point of interest. The configurations in ${\cal D}_N\setminus{\cal D}_N^{(m)}$ are precisely the ones that touch the surface at distance $m$, whence \begin{equation} Z_N(q)-qZ_N^{(m)}(q)=\sum_{\varphi\in{\cal D}_N\setminus{\cal D}_N^{(m)}}q^{A(\varphi)}=Z_m(q)Z_{N-m}(q)\;. \end{equation} For $q=1$ this reduces to Equation (\ref{catalanpressure}). Computing the pressure using Eqn.~(\ref{generalpressure}), we find that the pressure $P_N^{(m)}(q)$ of a Dyck vesicle of length $N$ on the surface at distance $m$ is given by \begin{equation} \label{pressureformula} P_N^{(m)}(q)=-\log\left(1-\frac{Z_m(q)Z_{N-m}(q)}{Z_N(q)}\right)+\log q\;. \end{equation} \section{Results} \subsection{Exact results} One can recursively calculate $Z_N(q)$ using \begin{equation} \label{recurrence} Z_0(q)=1\;,\quad Z_{N+1}(q)=\sum\limits_{k=0}^Nq^kZ_k(q)Z_{N-k}(q)\;,\quad N\geq 0\;, \end{equation} which for $q=1$ reduces to a well-known recursion for Catalan numbers. One sees that $Z_N(q)$, and therefore also $P_N^{(m)}(q)$, is computable in polynomial time. There are no closed-form expressions known. However, there are well-known closed-form expressions \cite[Example V.9]{flajolet2009a-a} for the generating function \begin{equation} G(z,q)=\sum_{N=0}^\infty Z_N(q)z^N\;, \end{equation} namely, \begin{equation} G(z,q)=\frac{A_q(z)}{A_q(z/q)}, \end{equation} where \begin{equation} A_q(z)=\sum_{n=0}^\infty\frac{q^{n^2}(-z)^n}{(q;q)_n} \end{equation} is Ramanujan's Airy function. Here, we use the $q$-product notation $(t;q)_n=\prod_{k=0}^{n-1}(1-tq^k)$. Note that the radius of convergence $z_c(q)$ of $G(z,q)$ is simply related to the thermodynamic limit of the partition function via \begin{equation} \log z_c(q)=-\lim_{N\to\infty}\frac1N\log Z_N(q)\;. \end{equation} For $q=1$, $G(z,q)$ is simply the Catalan generating function and hence $z_c(1)=1/4$. At the level of the generating function, the recurrence (\ref{recurrence}) is equivalent to the functional equation \begin{equation} \label{functeqn} G(z,q)=1+zG(z,q)G(qz,q)\;, \end{equation} which in turn leads to a nice continued fraction representation for $G(z,q)$: \begin{equation} \label{confrac} G(z,q)=\cfrac1{1-\cfrac z{1-\cfrac{qz}{1-\cfrac{q^2z}{1-\cfrac{q^3z}{1-\ldots}}}}}\;. \end{equation} We can calculate the surface pressure in the thermodynamic limit from the generating function by relating the pressure to the density of contacts with the surface. In order to do so, we need to introduce a surface weight $\kappa$ for contacts with the surface, which we will set to one after the calculation. Let $G(z,q;\kappa)$ be the generating function for area-weighted Dyck paths where each contact with the surface (except for the origin) is associated with a weight $\kappa$. A simple necklace argument gives \begin{equation} \label{kappaeqn} G(z,q;\kappa)= \frac1{1-z\kappa G(qz,q)}\;. \end{equation} The density of contacts with the surface well away from the ends of the walk is \begin{equation} \label{rhoeqn} \rho(q)=-\left.\frac{\partial\log z_c(q;\kappa)}{\partial\log\kappa}\right|_{\kappa=1}\;, \end{equation} where $z_c(q;\kappa)$ is the location of the closest singularity of $G(z,q;\kappa)$ in $z$ to the origin. The surface pressure in the thermodynamic limit, for any fixed value of $0<x=m/N<1$, is then given as the constant \begin{equation} P(q)=-\log(1-\rho(q))+\log q\;. \end{equation} We shall refer to this as the bulk pressure. \begin{figure}[ht] \begin{center} \includegraphics[width=0.5\textwidth]{Generatingfunctionscaling.pdf} \end{center} \caption{Convergence of the scaled generating function $G(z,q)$ to the scaling function $f(s)$ for $q=0.99$, $0.9999$, and $0.999999$. Numerical evaluation of $G(z,q)$ has been done using the continued fraction representation (\ref{confrac}). \label{Generatingfunctionscaling} } \end{figure} Much work has been done on the computation of the asymptotics for $q$-series such as those involved here \cite{prellberg1994a-a}. In the vicinity of $q=1$ and $z=z_c(1)=1/4$, one can show convergence of suitably scaled generating function. More precisely, one finds that the limit \begin{equation} \label{scaling} f(s)=\lim_{q\to1^-}-\left((1-q)^{-1/3}G(1/4-s(1-q)^{2/3},q)-2\right) \end{equation} exists and is equal to the scaling function \begin{equation} f(s)=-2\frac{\mbox{\rm Ai}'(4s)}{\mbox{\rm Ai}(4s)} \end{equation} The convergence to the scaling function is shown in Figure \ref{Generatingfunctionscaling}. \subsection{Pressure profiles} We now consider the profile of the pressure for finite $N$. In Figure \ref{pressureprofiles} we show pressure profiles as a function of $x=m/N$ for three different values of $q$ with positive, zero, and negative osmotic pressures, and for lengths $N=50$, $100$, and $200$. \begin{figure}[ht] \begin{center} \includegraphics[width=0.3\textwidth]{pressure_q098.pdf} \includegraphics[width=0.3\textwidth]{pressure_q100.pdf} \includegraphics[width=0.3\textwidth]{pressure_q102.pdf} \end{center} \caption{Pressure profiles for $q=0.98$ (left), $1.00$ (center), and $1.02$ (right) and lengths $N=50$, $100$, and $200$, from top to bottom. \label{pressureprofiles} } \end{figure} We see that the pressure converges to a non-zero bulk pressure when the osmotic pressure is non-zero. For $q=1$, one can compare the finite-size data shown in this figure to the exactly known scaled profile shown in Figure \ref{profile}. Moreover, closer inspection shows that the rate of convergence is significantly different for positive and negative osmotic pressure. As we know, for $q=1$ the bulk pressure is zero, and convergence obeys the power law $N^{-3/2}$. For $q\neq1$, the rate of convergence to the bulk pressure is exponential in $N$ and $N^2$ for $q<1$ and $q>1$, respectively, as can be seen in Figure \ref{pressuredecay}. \begin{figure}[ht] \begin{center} \includegraphics[width=0.3\textwidth]{pressure_q090_decay.pdf} \includegraphics[width=0.3\textwidth]{pressure_q110_decay.pdf} \end{center} \caption{The rates of approach to the bulk pressure, calculated via the pressure at the centre of the vesicle for $q=0.9$ (left) and $q=1.1$ (right). Clearly the convergence is exponential in $N$ for $q<1$ and exponential in $N^2$ for $q>1$. The straight line (right) is our theoretical prediction. \label{pressuredecay} } \end{figure} One can derive these results in the following way. Convergence to the thermodynamic limit is encoded in the singularity structure of the generating function. For $q<1$, the leading singularity of the generating function is an isolated simple pole, and the finite-size corrections to scaling are therefore exponential, with the rate of convergence given by the ratio between the magnitudes of the leading singularity and the sub-leading one. A closer analysis reveals that the rate depends on the value $x=m/N$ as \begin{equation} P_N^{(m)}(q)\sim-\log(1-\rho(q))+\log q+O(e^{-D\min(x,1-x)N})\;, \end{equation} where $D$ does not depend on the value of $x$. The singularity structure can also be seen in Figure \ref{Generatingfunctionscaling} for negative values of the scaling parameter $s$. For $q>1$, it is easier to argue directly via the partition functions. For positive osmotic pressure, configurations with large area dominate, and using arguments in \cite{prellberg1999a-:a} one can deduce that \begin{equation} Z_N(q)\sim\frac1{(q^{-1};q^{-1})_\infty}q^{\binom N2} \end{equation} where $(t;q)_\infty=\prod_{k=0}^\infty(1-tq^k)$ is again a $q$-product. The appearance of the factor $1/(q^{-1};q^{-1})_\infty$ is due to fluctuations around configurations of maximal area. Substituting this into Eqn.~(\ref{pressureformula}) shows that \begin{equation} P_N^{(m)}(q)\sim\log q+\frac1{(q^{-1};q^{-1})_\infty}q^{-m(N-m)}=\log q+\frac1{(q^{-1};q^{-1})_\infty}q^{-x(1-x)N^2}\;. \end{equation} Note that the decay rate depends on the value of $x=m/N$. \subsection{Bulk pressure} \begin{figure}[ht] \begin{center} \includegraphics[width=0.4\textwidth]{pressure_vs_q_in_bulk.pdf} \includegraphics[width=0.4\textwidth]{pressure_vs_q_in_bulk_in_TDL.pdf} \end{center} \caption{A plot of the pressure at the centre of the vesicle against $q$ for vesicle lengths $10$, $20$, $40$ and $80$, from top to bottom (left) and of the thermodynamic limit bulk pressure of the vesicle against $q$ (right). \label{Pvsq} } \end{figure} We now turn to the consideration of the pressure in the centre of the vesicle, so as to consider the bulk pressure. Using our results above, we already know that for $q>1$ the bulk pressure is given by $\log q$. In Figure \ref{Pvsq} on the left we have used the recursion (\ref{recurrence}) to calculate the pressure at the centre of the vesicle as a function of $q$ for $N=10$, $20$, $40$, and $80$. One can see that the convergence to a limit is slowest around $q=1$, which of course aligns with our predictions above for the rates of convergence in the different regimes. We have now used the continued fraction expansion (\ref{confrac}) to numerically estimate the thermodynamic limit bulk pressure as a function of $q$, shown on the right in Figure \ref{Pvsq}. It should be clear that the finite-size curves approach the curve shown here in the limit of large $N$. It is interesting to see the competition of the osmotic pressure and the entropic pressure for $q<1$. Clearly the limiting behaviour of the pressure for $q>1$ is $P(q)\sim q-1$ as $q\to1^+$. Extracting the limiting behaviour of the pressure for $q<1$ is considerably more difficult. A calculation starts by considering that from Equations (\ref{functeqn}) and (\ref{kappaeqn}) we find that the critical fugacity $z_c=z_c(q;\kappa)$ satisfies \begin{equation} G(z_c,q)=\frac{\kappa}{\kappa-1}\;. \end{equation} Differentiating this expression with respect to $\kappa$ allows us to write the density $\rho(q)$ in terms of the generating function as \begin{equation} \rho(q)=\frac{G(z_c,q)(G(z_c,q)-1)}{z\frac{\partial}{\partial z_c}G(z_c,q)}\;. \end{equation} Utilising the scaling form (\ref{scaling}) now shows that $\rho(q)\sim2(1-q)$ as $q\to1^-$ and hence $P(q)\sim 1-q$. Put together, this implies that \begin{equation} \label{Pasy} P(q)\sim|1-q|\quad\text{as $q\to 1$.} \end{equation} \begin{figure}[ht] \begin{center} \includegraphics[width=0.4\textwidth]{Pressurescaling.pdf} \end{center} \caption{An estimate of the finite-size scaling function for the pressure, using $N=40$, $80$, and $160$, from bottom to top. \label{pressurescaling} } \end{figure} The existence of the scaling function $f(s)$ in the variable $s=(1/4-z)/(1-q)^{2/3}$ implies by standard Laplace transform that there should be a finite-size scaling form for the pressure in the variable $t=N^{3/2}(1-q)$. Hence we define the scaling function \begin{equation} g(t)=\lim\limits_{N\to\infty}N^{3/2}P_N^{(N/2)}(1-tN^{-3/2})\;. \end{equation} In Figure \ref{pressurescaling}, we numerically estimate the scaling function $g(t)$ for a range of $t$. Asymptotic matching with (\ref{Pasy}) requires that $g(t)\to|t|$ for $t\to\pm\infty$. We note the convergence to the scaling function is poor for $q>1$, which arises because of the unusually differing rates of convergence to the thermodynamic limit in the two regimes. \section*{Acknowledgements} Financial support from the Australian Research Council via its support for the Centre of Excellence for Mathematics and Statistics of Complex Systems is gratefully acknowledged by the authors. A L Owczarek thanks the School of Mathematical Sciences, Queen Mary, University of London for hospitality. \providecommand{\newblock}{}
\section{Introduction} An understanding of the origin of the elements in the universe requires an understanding of the nucleosynthesis in Type IIP supernovae (SNe) and their progenitors, which make up 40\% of all SN explosions in the local universe \citep{Li2011}. These are the explosions of stars that have retained most of their hydrogen envelopes throughout their evolution, which results in a $\sim$100 day hydrogen recombination plateau in the light curve. After this follows the nebular phase, when the inner regions of the ejecta become visible. With recently produced SN ejecta grids from a range of progenitor masses, improvements in atomic data, and the development of NLTE (Non-Local Thermodynamic Equilibrium) radiative transfer codes \citep[e.g.,][]{Dessart2011,Jerkstrand2011,Maurer2011}, self-consistently calculated synthetic spectra for these kind of objects are now achievable. These models can be used to diagnose the nucleosynthesis in individual events and improve our understanding of the role of Type IIP SNe for the cosmic production of abundant intermediate-mass elements such as oxygen. Nucleosynthesis analysis can be performed in the nebular phase of the SN evolution, when the photosphere has receded to reveal the products of hydrostatic and explosive burning. Since the nucleosynthesis depends strongly on the main-sequence mass of the star \citep[e.g.][]{Woosley1995}, comparing the strengths of observed nebular lines to models allows an estimate of the main-sequence mass. In the cases where pre-explosion images of the progenitor are available, another powerful technique for estimating the progenitor mass is by comparing the luminosity of the progenitor with stellar evolutionary models \citep[e.g.][]{Smartt2009a}. Combining such progenitor analysis with nebular phase spectral modelling offers a way of tightly constraining stellar evolution models and nucleosynthesis. A third approach to analyse the link between progenitors and SN observables is through hydrodynamical modeling of the diffusion-phase light curve \citep[see e.g.][]{Utrobin2008, Utrobin2009, Bersten2011, Pumo2011}. In \citet[][J12 hereafter]{Jerkstrand2012}, we developed models for the spectra of Type IIP SNe from the onset of the nebular phase ($\sim140$ days) up to 700 days, to use for spectral analysis. These models solve for the non-thermal energy deposition channels, the statistical and thermal equilibrium for each layer of nuclear burning ashes, and the internal radiation field. The calculations are currently applied to SN ejecta from stars evolved and exploded with KEPLER \citep{Woosley2007}. Here, we apply these models to optical and near-infrared observations of SN 2012aw, a nearby Type IIP SN \citep{Bose2013, dallOra2013} with a potentially luminous and massive progenitor directly identified in pre-explosion images \citep{Fraser2012, vandyk2012,Kochanek2012}. The inferred luminosity of the progenitor was disputed in these separate analyses (see Sect. \ref{sec:discussion}) but is potentially higher than most previous red supergiant progenitor stars and a candidate to be the first star with $M_{\rm ZAMS} > 20$ \msun\ (from progenitor analysis) seen to explode as a Type IIP SN. The lack of any such stars in the sample of detections and upper limits obtained so far has been termed the ``red supergiant problem'' \citep{Smartt2009b}, and is currently challenging our understanding of the final outcomes of $M_{\rm ZAMS}=20-30$ \msun\ stars, and their role in galactic nucleosynthesis. The strong theoretical dependency of the oxygen yield with progenitor mass makes SN 2012aw an interesting candidate to analyze for signs of the high mass of synthesised oxygen expected from a $M_{\rm ZAMS} > 20$ \msun\ progenitor. \section{Observational data} \label{sec:obs} We obtained four optical ($250-451$ days) and one near-infrared (306 days) spectra of SN 2012aw using the William Herschel Telescope (WHT) and the Nordic Optical Telescope (NOT), as detailed in Table \ref{tab:log}. The NIR spectrum at 306 days post-explosion is among the latest near-infrared spectra taken for a Type IIP SN. The optical spectra were taken with ISIS (WHT) and ALFOSC (NOT), while the near-infrared (NIR) WHT spectrum was taken with LIRIS. All ISIS data were reduced within {\sc iraf} using standard techniques. The spectra were bias-subtracted and flat-fielded using a normalised flat field from an internal lamp. Cosmic rays were removed in the 2D spectral images using {\sc lacosmic} \citep{vandokkum2001}, before individual exposures were combined, and the 1D spectrum was optimally extracted. Arc spectra from CuNe and CuAr lamps were used to wavelength calibrate the extracted spectra, while flux calibration was performed using observations of spectrophotometric standard stars taken with the same instrumental configuration on the same night. No attempt was made to correct for telluric absorptions. The NOT spectra were reduced in a similar manner, including corrections for second order contamination \citep{Stanishev2007}. To calibrate the spectra to photometry, we scaled each spectrum with a constant to minimize the RMS difference between synthetic photometry calculated from the spectra and observed Johnson-Cousins $BVRI$ photometry. For the day 250 spectrum we used the day 248 photometry of \citet{Bose2013} and for the day 332 spectrum we used the day 333 photometry of \citet{dallOra2013}. For the day 369 and day 451 spectra, we employed an extrapolation of the \citet{dallOra2013} day 333 photometry assuming decline rates based on the \citet{Sahu2006} observations of SN 2004et (which had a spectral evolution similar to SN 2012aw) between $336-372$ days for the day 369 spectrum and between $336-412$ days for the 451 day spectrum. We obtain RMS errors of 4.6\% (250d), 8.0\% (332d), 7.3\% (369d), and 11\% (451d) for the $BVRI$ scatter, which we regard as representative of the absolute flux errors. \begin{table*} \caption{Log of spectroscopy of SN 2012aw. The phase is with respect to the explosion epoch of March 16 2012 \citep{Fraser2012}. The resolution is the FWHM resolution at the midpoint of the spectral range.} \begin{tabular}{lcccccccc} Date & Phase &Instrument & Disperser & Coverage & Exp. Time & Slit & Dispersion & Resolution \\ & (days) & & & (\AA) & (seconds) & (\arcsec) & (\AA/pixel) & (\AA) \\ \hline 2012-11-21 & +250 & WHT+ISIS & R300B/R158R & 3300-9500 & 1800 & 2.0 & 0.86/1.8 & 8.2/15.4\\ 2013-01-16 & +306 & WHT+LIRIS & zJ+HK & 8870-24000 & 2000/1600 & 1.0 & 6.1/10 & 14/30 \\ 2013-02-11 & +332 & WHT+ISIS & R300B/R158R & 3300-9500 & 1800 & 1.5 & 0.86/1.8 & 6.2/11.6\\ 2013-03-20 & +369 & NOT+ALFOSC & Grism 4 & 3330-9140 & 5400 & 1.0 & 2.95 & 16 \\ 2013-06-10 & +451 & WHT+ISIS & R300B/R158R & 3300-9500 & 2700 & 1.0 & 0.86/1.8 & 4.1/7.7 \\ \hline \end{tabular} \label{tab:log} \end{table*} For the LIRIS NIR spectrum, we alternated the target between two positions in the LIRIS slit. By subtracting pairs of spectra at different slit positions from each other, the sky background and detector bias were removed. LIRIS suffers from a problem where pixels may be shifted during readout, so we first ``descrambled'' the raw data to correct for this, using the {\sc lcpixmap} task with the {\sc liris} data reduction package within {\sc iraf} \footnote{http://www.ing.iac.es/astronomy/instruments/liris/liris\_ql.html}. The 2D spectra were then flat-fielded and shifted to spatially align the spectra. They were then extracted and wavelength calibrated using Xe and Ar arc lamps. A solar analog was observed in between the zJ and HK spectra of SN 2012aw at a similar airmass, and this was used to correct for the telluric absorption. As there are few good spectrophotometric standards in the NIR, we performed an approximate flux calibration, estimated by scaling the flux of the telluric standard to match its catalogued 2MASS {\it JHK} magnitudes. The resulting spectrum showed a $\sim$30\% flux deficiency in the overlap region with both the day 250 and day 332 optical spectra (scaled with exponential decay factors to the same epoch). We therefore applied an additional correction factor of 1.3 to force overlap between the NIR spectrum and the optical spectra. To compare the observed spectra to models, we adopt a distance of 9.9 Mpc \citep{Bose2013}, an extinction of $E_{B-V}=0.074$ mag \citep{Bose2013} and the extinction law of \citet{Cardelli1989}, with $R_V=3.1$. The uncertainties in distance and extinction are small (0.1 Mpc and 0.008 mags, respectively, \citet{Bose2013}). We also adopt a recessional velocity of 778 km s$^{-1}$ \citep{Bose2013}, and an explosion epoch of March 16 2013 (MJD 56002) from \cite{Fraser2012}. The ejected mass of $^{56}$Ni is well constrained to be $0.06\pm 0.01$ \msun\ by the evolution in the early nebular phase when gamma-ray trapping is almost complete \citep{Bose2013}. \section{Modelling} We compute $M_{ZAMS}=12,\ 15\ \mbox{and}\ 19$ \msun\ models at the observational epochs of SN 2012aw, following the same method and model setup as described in J12, apart from a few minor code updates described below. The model setup procedure (as more completely described in J12) involves dividing the ejecta into several chemically distinct zones - Fe/He, Si/S, O/Si/S, O/Ne/Mg, O/C, He/C, He/N, and H, named after their dominant components. To mimic the mixing structures obtained in multidimensional explosion simulations, the metal zones (Fe/He, Si/S, O/Si/S, O/Ne/Mg, O/C) are macroscopically mixed together with parts of the He/C, He/N, and H zones (fractions 0.6, 0.6, and 0.15 respectively) in a core region between 0 and 1800 km s$^{-1}$. Each zone has an individual filling factor in this core, and is distributed over $10^3$ (identical) clumps. The model takes dust formation into account by applying a gray opacity over the core from 250 days, growing with $d \tau_{dust} = 1.8\e{-3}$ per day (a calibration to the observed dust emission of SN 2004et). Over the period covered here, this dust component has only a small effect on the optical/NIR spectrum as $\tau_{dust} < 0.18$ at all times. Outside the core reside the remaining helium layers followed by the hydrogen envelope, whose density profile is determined by the one-dimensional hydrodynamic solutions. The exact zone masses, chemical compositions, and density profiles for the 12, 15 and 19 \msun\ models are given in J12. In addition to these, we here also compute the spectral evolution of the $M_{\rm ZAMS}=25$ \msun\ model of \citet{Woosley2007}, which was set-up following the same method. Table \ref{table:25} shows its zone masses, filling factors, and chemical composition. We make a few minor updates to the code compared to the version used in J12. First, we replace the non-relativistic Doppler formula with the relativistic one \citep[e.g.][]{Rybicki1979} \begin{equation} \lambda = \lambda' \gamma(V) \left(1 - \frac{V}{c}\cos{\theta}\right) \end{equation} where $\lambda$ and $\lambda'$ are the wavelengths in the two frames, $V$ is the relative velocity, $\theta$ is the angle (between the direction of the photon and the velocity), $c$ is the speed of light, and $\gamma=\left(1-V^2/c^2\right)^{-1/2}$. We find this correction to give small but discernible differences in the line profiles, useful for detailed analysis of dust effects etc. Second, we increase the number of photon packets in the Monte Carlo simulations to keep photon noise to a minimum. Finally, we correct a minor error in the J12 calculations where the first photoexcitation rate in each NLTE solution was accidentally set to zero. Recomputation shows that this did not have any noticeable effects on the spectra apart from a 5-10\% shift in the Na I D line luminosity. \begin{table*} \caption{The 25 \msun\ model. The total ejecta mass is 13.7 \msun. The core filling factors are denoted $f_{core}$. Mass fractions below $10^{-9}$ are put to zero in the model.} \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline \hline Zone & Fe/He & Si/S & O/Si/S & O/Ne/Mg & O/C & He/C & He/N & H \\ \hline Mass $(M_\odot$) & 0.064 & 0.18 & 0.62 & 2.7 & 1.4 & 1.0 & 0.35 & 7.4 \\ $f_{core}$ & 0.15 & 0.10 & 0.028 & 0.15 & 0.077 & 0.081 & 0.027 & 0.39\\ \emph{Mass fractions:}\\ $^{56}$Ni + $^{56}$Co & 0.77 & 0.075 & $4.8\e{-7}$ & $1.3\e{-7}$ & $3.9\e{-8}$ & $4.2\e{-8}$ & $2.9\e{-8}$ & 0\\ $^{57}$Co & 0.031 & $1.7\e{-3}$ & $4.4\e{-6}$ & $1.7\e{-7}$ & $2.1\e{-8}$ & $6.8\e{-9}$ & $1.4\e{-9}$ & 0 \\ $^{44}$Ti & $3.4\e{-4}$ & $1.8\e{-5}$ & $1.7\e{-6}$ & 0 & 0 & 0 & 0 & 0\\ H & $7.5\e{-8}$ & $1.3\e{-9}$ & 0 & 0 & 0 & 0 & $6.7\e{-8}$ & 0.53\\ He & 0.12 & $1.1\e{-5}$ & $1.9\e{-6}$ & $1.5\e{-6}$ & $2.6\e{-5}$ & 0.96 & 0.99 & 0.45 \\ C & $2.7\e{-6}$ & $6.4\e{-7}$ & $8.6\e{-5}$ & $8.1\e{-3}$ & 0.20 & 0.017 & $3.2\e{-4}$ & $8.6\e{-4}$ \\ N & $2.8\e{-7}$ & 0 & $2.2\e{-5}$ & $4.2\e{-5}$ & $1.5\e{-5}$ & $7.7\e{-6}$ & $9.0\e{-3}$ & $5.5\e{-3}$ \\ O & $7.9\e{-6}$ & $6.7\e{-6}$ & 0.55 & 0.74 & 0.75 & $4.1\e{-3}$ & $1.6\e{-4}$ & $3.5\e{-3}$ \\ Ne & $8.6\e{-6}$ & $1.7\e{-6}$ & $1.5\e{-4}$ & 0.16 & 0.034 & 0.014 & $1.1\e{-3}$ & $1.2\e{-3}$ \\ Na & $4.8\e{-7}$ & $7.4\e{-7}$ & $5.5\e{-6}$ & $2.2\e{-3}$ & $2.0\e{-4}$ & $1.9\e{-4}$ & $1.8\e{-4}$ & $9.9\e{-5}$ \\ Mg & $1.9\e{-5}$ & $1.7\e{-4}$ & 0.014 & 0.063 & $7.1\e{-3}$ & $5.8\e{-4}$ & $5.7\e{-4}$ & $6.1\e{-4}$ \\ Al & $6.7\e{-6}$ & $2.2\e{-4}$ & $4.5\e{-4}$ & $6.3\e{-3}$ & $1.0\e{-4}$ & $7.5\e{-5}$ & $1.0\e{-4}$ & $6.9\e{-5}$\\ Si & $2.3\e{-4}$ & 0.37 & 0.26 & 0.013 & $8.9\e{-4}$ & $8.3\e{-4}$ & $8.2\e{-4}$ & $8.2\e{-4}$ \\ S & $1.9\e{-4}$ & 0.39 & 0.15 & $3.9\e{-4}$ & $2.3\e{-4}$ & $4.1\e{-4}$ & $4.2\e{-4}$ & $4.2\e{-4}$ \\ Ar & $1.5\e{-4}$ & 0.060 & 0.022 & $8.1\e{-5}$ & $8.6\e{-5}$ & $1.1\e{-4}$ & $1.1\e{-4}$ & $1.1\e{-4}$\\ Ca & $1.7\e{-3}$ & 0.042 & $5.0\e{-3}$ & $2.7\e{-5}$ & $2.8\e{-5}$ & $7.3\e{-5}$ & $7.4\e{-5}$ & $7.4\e{-5}$ \\ Sc & $2.1\e{-7}$ & $2.7\e{-7}$ & $2.7\e{-7}$ & $1.5\e{-6}$ & $7.8\e{-7}$ & $8.5\e{-8}$ & $4.5\e{-8}$ & $4.5\e{-8}$\\ Ti & $1.1\e{-3}$ & $5.8\e{-4}$ & $4.8\e{-5}$ & $7.9\e{-6}$ & $8.0\e{-6}$ & $3.4\e{-6}$ & $3.4\e{-6}$ & $3.4\e{-6}$ \\ V & $1.3\e{-5}$ & $1.5\e{-4}$ & $3.8\e{-6}$ & $5.2\e{-7}$ & $3.2\e{-7}$ & $4.9\e{-7}$ & $4.3\e{-7}$ & $4.3\e{-7}$ \\ Cr & $1.6\e{-3}$ & $7.5\e{-3}$ & $3.3\e{-5}$ & $1.2\e{-5}$ & $1.3\e{-5}$ & $2.0\e{-5}$ & $2.0\e{-5}$ & $2.0\e{-5}$ \\ Mn & $2.1\e{-6}$ & $3.3\e{-4}$ & $2.3\e{-6}$ & $2.8\e{-6}$ & $2.1\e{-6}$ & $1.8\e{-5}$ & $1.5\e{-5}$ & $1.5\e{-5}$\\ Fe & $9.7\e{-4}$ & 0.047 & $4.7\e{-4}$ & $5.7\e{-4}$ & $5.8\e{-4}$ & $1.4\e{-3}$ & $1.4\e{-3}$ & $1.4\e{-3}$ \\ Co & $3.7\e{-8}$ & $3.1\e{-9}$ & $1.7\e{-6}$ & $1.6\e{-4}$ & $1.7\e{-4}$ & $4.4\e{-6}$ & $4.0\e{-6}$ & $4.0\e{-6}$ \\ Ni & 0.037 & $2.7\e{-3}$ & $1.1\e{-3}$ & $6.6\e{-4}$ & $6.8\e{-4}$ & $8.2\e{-5}$ & $8.2\e{-5}$ & $8.2\e{-5}$ \\ \hline \end{tabular} \label{table:25} \end{table*} All model spectra presented in the paper have been convolved with a Gaussian with $FWHM=\lambda/500$, corresponding to the typical resolving power of the instruments. \section{Results} Figs. \ref{fig:optical} and \ref{fig:NIR} show the observed (dereddened and redshift corrected) optical and near-infrared spectra of SN 2012aw, compared to the $M_{\rm ZAMS}=15\ M_{\odot}$ model computed at the corresponding epochs. The model has an initial $^{56}$Ni mass of 0.062 \msun, and the agreement in flux levels during the early radioactive tail phase shows that SN 2012aw has a similar $^{56}$Ni mass, as also found by \citet{Bose2013}. Spectroscopically, there is satisfactory agreement throughout the evolution, showing that the ejecta model is representative of the SN 2012aw ejecta and that the model calculations capture the main physical processes important for the formation of the spectrum. The main shortcomings of the model - too strong H$\alpha$, Pa$\alpha$, He I \wl1.083 $\upmu$m, He I \wl2.058 $\upmu$m, as well as too strong scattering in Ca II \wll8542, 8662, are the same issues that were present in the analysis of SN 2004et (J12). The incorrect Ca II NIR triplet reproduction occurs as Ca II \wl8498 and Ca II \wl8542 scatter in Ca II \wl8662 in the model, presumably due to a too high density in the envelope, or a poorly reproduced ionization balance in calcium (the relative amounts of Ca II and Ca III). The H and He line discrepancies are likely related to the mixing-scheme applied for bringing hydrogen and helium clumps into the radioactive core, but do not strongly affect our analysis of the metal emission lines below. \begin{figure*} \includegraphics[trim=10mm 0mm 18mm 0mm, clip, width=0.95\linewidth]{fig2.eps} \caption{Observed (dereddened and redshift corrected) spectra of SN 2012aw (red) compared to the 15 $M_{\odot}$ model (blue). The gray band shows the strongest telluric absorption band.} \label{fig:optical} \end{figure*} \begin{figure*} \includegraphics[trim=10mm 0mm 18mm 0mm, clip, width=0.95\linewidth]{fig1.eps} \caption{Observed (dereddened and redshift corrected) near-infrared spectrum of SN 2012aw at 306 days (red) compared to the 15 $M_{\odot}$ model (blue). Inset is a zoom-in on the CO overtone, and a comparison with the overtone in SN 2004et at the same epoch \citep{Maguire2010}, scaled to the same distance assuming $D=5.5$ Mpc and $E_{B-V}=0.41$ mag for SN 2004et. The gray bands show the strongest telluric absorption bands.} \label{fig:NIR} \end{figure*} There is a unique emission line that provides a direct link to the nucleosynthesis - the [O I] \wll 6300, 6364 doublet. The usefulness of this line stems from several factors. Oxygen production is highly sensitive to the helium-core mass, which in turn depends on the zero-age main sequence (ZAMS) mass \citep[e.g.][]{Woosley1995, Thielemann1996}. Additionally, most of the oxygen is synthesized during hydrostatic helium and carbon burning, with little creation or destruction in the explosive process. Finally, the carbon burning ashes (the O/Ne/Mg zone) that contain most of the synthesised oxygen are devoid of much carbon and silicon, which prevents the formation of any large amounts of CO and SiO (at least for $t < 500$\ days, \citet{Sarangi2013}) that would damp out the thermal [O I] \wll 6300, 6364 emission, as happens in the helium and neon burning ashes \citep[e.g][]{Liu1995}. For SN 2004et, the [O I] \wll 6300, 6364 luminosity was about as strong as the combined molecular emission in fundamental and overtone bands of CO and SiO \citep[][J12]{Kotak2009}, showing that the bulk of the oxygen is not molecularly cooled. In SN 2012aw, the CO overtone at 2.3 $\upmu$m is similar in strength to SN 2004et (Fig. \ref{fig:NIR}), consistent with the amount of molecular cooling being similar in these two objects. In the absence of significant molecular cooling, the [O I] \wll 6300, 6364 doublet becomes the main cooling agent of the carbon burning ashes, reemitting $50-70$\% of the deposited thermal energy in our models in the $300-700$ day interval. Since a larger mass of oxygen-rich ashes absorbs a larger amount of gamma-rays, the line is a direct tracer of the oxygen production. The analysis is aided by the oxygen density constraints set by the evolution of the [O I] \wl 6300/[O I] \wl 6364 line ratio \citep{Spyromilio1991, Li1992}, although for low-mass stars ($M_{\rm ZAMS} \lesssim 12$ \msun)\ contributions by primordial oxygen in the hydrogen envelope necessitates a multi-component analysis \citep{Maguire2012}. The density used in the model is the one derived from observations of the [O I] \wl 6300/[O I] \wl 6364 line ratio in SN 1987A. It is clear from Fig. \ref{fig:optical} that the observed line ratio in SN 2012aw evolves in a manner consistent with the model, and thus the oxygen density in SN 2012aw must be similar to the one in SN 1987A. Also many other Type IIP SNe show a similar evolution of this line ratio, suggesting that the oxygen density shows little variation within the Type IIP class \citep{Maguire2012}. Another important aspect of the [O I] \wll 6300, 6364 modelling is to have a representative morphological structure for how the gamma-ray emitting $^{56}$Co clumps are hydrodynamically mixed with the oxygen clumps, since this determines the amount of gamma-ray energy deposited into the oxygen clumps. The model we use assumes a uniform distribution of $^{56}$Ni and O clumps between zero and 1800 km s$^{-1}$, a scenario that emerges by Rayleigh-Taylor mixing in the explosion \citep[e.g.][]{Kifonidis2006} and is empirically supported by the observed similarity between iron and oxygen emission line profiles (J12, see also Fig. \ref{fig:oi5577} here). A check on this mixing treatment is nevertheless desirable and the ratio of [O I] \wl 5577 to [O I] \wll 6300, 6364 can help. Whereas [O I] \wll 6300, 6364 has been widely discussed and used in previous nebular analyses, [O I] \wl 5577 has rarely been used since it is about an order of magnitude weaker than [O I] \wll 6300, 6364 and lies in a more blended spectral region. But if a luminosity can be extracted, the [O I] \wl 5577 line is useful as the [O I] \wl 5577/[O I] \wll 6300,~6364 ratio is sensitive to the temperature and therefore the gamma-ray deposition per unit mass. Both lines are excited by thermal collisions, but as they have different excitation energies their line ratio depends on the temperature (and density) \citep[see e.g.][]{Fransson1989}. An emission line that we identify with [O I] \wl 5577 is clearly detected in SN 2012aw (Fig. \ref{fig:oi5577}), although it is partially blended with another line that we identify with [Fe II] \wl5528. To measure the luminosity in [O I] \wl 5577 we simultaneously fit the [Fe II] \wl5528 and [O I] \wl 5577 lines with two Gaussians, with the results reported in Table \ref{table:oilines}. \begin{figure} \includegraphics[trim=0mm 10mm 17mm 15mm, clip, width=1\linewidth]{fig4.eps} \caption{SN 2012aw (red, solid) between 5400-5700 \AA\ (dereddened and redshift corrected), the double Gaussian fit to [Fe II] \wl 5528 and [O I] \wl 5577 (black, solid), the [O I] \wl 5577 component to the fit (black, dashed), the 15 \msun\ model (blue, solid) and the [O I] \wl 5577 component in the 15 \msun\ model (blue, dashed). The Gaussian central wavelengths are fixed at 5533 \AA\ (the 270 km s$^{-1}$ offset from 5528 \AA\ may be due to contribution by other lines or possibly by an asymmetry in the iron distribution) and 5577 \AA, and the widths are fixed at $FWHM = 40$ \AA.} \label{fig:oi5577} \end{figure} At the temperatures of interest here, the critical densities (above which the lines form in LTE) for the [O I] \wl5577 and [O I] \wll6300, 6364 lines are $n_e^{5577} = 10^8 \beta_{5577}$ cm$^{-3}$ and $n_e^{6300,6364} = 3\e{6} \beta_{6300,6364}$ cm$^{-3}$, respectively, where $\beta_\lambda = \left(1-\exp{\left(-\tau_\lambda\right)}\right)/\tau_\lambda$ is the Sobolev escape probability and $\tau_\lambda$ is the Sobolev optical depth \citep{Sobolev1957}. The [O I] \wll 6300, 6364 critical density is well below the oxygen-zone electron density in the model for several years, implying that the line is formed in LTE. For the [O I] \wl 5577 line, the LTE approximation is accurate up to $\sim$250 days, but after that the model luminosity starts falling below the LTE value. In LTE, the line ratio is \begin{flalign} &\frac{L_{5577}}{L_{6300, 6364}} = \frac{g_{2p^4(^1S)}}{g_{2p^4(^1D)}} e^{\frac{-\Delta E}{kT}}\frac{A_{5577}\beta_{5577}}{A_{6300,6364} \beta_{6300,6364}}\frac{\lambda_{6300,6364}}{\lambda_{5577}} \nonumber \\ &= 38 \times \exp{\left(\frac{-25790\ K}{T}\right)}\frac{\beta_{5577}}{\beta_{6300, 6364}}~. \label{eq:ratio} \end{flalign} where we have used $\Delta E = E(2p^4(^1S))-E(2p^{4}(^1D)) = 2.22\ \mbox{eV}$, $g(2p^4(^1S))=1$, $g(2p^{4}(^1D))=5$, $A_{6300,6364}=7.4\e{-3}$ s$^{-1}$, and $A_{5577}=1.26$ s$^{-1}$. If the [O I] \wl 5577 line is formed in NLTE, the line ratio will be smaller than the quantity on the RHS, still allowing a \emph{lower limit} to $T$ to be set. This in turn translates to an \emph{upper limit} in the oxygen mass, which we will find useful. A direct and model-independent determination of $T$ is complicated by the factor $\beta_{5577}/\beta_{6300,6364}$, but fortunately this can be constrained. Since the observed [O I] \wl 6364 line is always weaker than [O I] \wl 6300, the doublet is at least partially transitioning to optical thinness (in the optically thick limit the two lines have equal strength whereas in the optically thin limit the [O I] \wl6364 line is three times weaker than the [O I] \wl6300 line), so $\tau_{6300,6364} \lesssim 2$, or equivalently $\beta_{6300,6364} \gtrsim 0.5$. The optical depth in [O I] \wl 5577 is always smaller than in [O I] \wll 6300, 6364 (at least for the $T \lesssim 5000$ K regime that is relevant here), so also $\beta_{5577} \gtrsim 0.5$. Thus, the $\beta_{5577}/\beta_{6300,6364}$ ratio must be in the range $1-2$. In our models, the ratio is between $1.3 - 1.6$ over the evolution covered here ($250 - 450$ days). If we use a value of 1.5 throughout, the observed line ratios correspond to the temperatures listed in column five of Table \ref{table:oilines}. These temperatures are $\sim$500 K lower than the ones computed in the model. Part of this may be due to NLTE in [O I] \wl 5577, which causes the temperatures computed assuming LTE to be lower than the true temperatures. Inspection of our statistical equilibrium solutions indeed shows that the parent state of [O I] \wl 5577 (2p$^4$($^1$S)) deviates from LTE after 250 days, with a departure coefficient $n_{2p^4(1S)}/n_{2p^4(1S)}^{LTE}=0.8-0.3$ over the $250-450$ day interval. Nevertheless, the [O I] \wl 5577 line in the model is somewhat brighter than the observed line (Fig. \ref{fig:oi5577}), so part of the temperature discrepancy may also be due to a too high gamma-ray deposition into the oxygen clumps in the model, in turn caused by a too strong mixing between $^{56}$Ni and oxygen clumps. Although the mixing affects the thermal conditions in the oxygen zones, its influence via the ionization balance is small as almost all oxygen is neutral. Given a temperature estimate, the O I mass can be estimated from the luminosity in either line. The weaker exponential dependency of the [O I] \wll 6300, 6364 luminosity with temperature propagates a smaller error from temperature errors, and in addition $L_{6300,6364}$ can be measured to higher accuracy than $L_{5577}$ since it is always stronger. We thus use the equation \begin{align} M_{\rm OI} = L_{6300,6364} 16 m_p \frac{Z(T)}{g_{2p^4(^1D)}}e^{\frac{E_{2p^4(^1D)}}{kT}} \left(A \beta h \nu\right)^{-1}_{6300,6364} \nonumber \\ = \frac{L_{6300,6364} / \beta_{6300, 6364} }{9.7\e{41}\ \mbox{erg s}^{-1}}\times \exp{\left(\frac{22720\ K}{T}\right)}\ M_\odot\ , \label{eq:oimasses} \end{align} to estimate the oxygen mass (where $m_p$ is the proton mass, and we approximate the partition function $Z(T)$ with the ground state statistical weight, $g_{2p^4(^3P)}=9$). If we use $\beta_{6300,6364}=0.5$ throughout (its minimum value based on the [O I] \wl 6300/[O I] \wl 6364 line ratio , we obtain the O I masses listed in the last column of Table \ref{table:oilines} ($\sim0.6$ \msun). These are close to the total masses of thermally emitting oxygen as the gas is mostly neutral ($x_{\rm OII} \lesssim 0.1$ at all times in the models). The derived masses are in the range $0.4-0.9$ \msun\ at all epochs, which corresponds to the amount of oxygen in the O/Ne/Mg layer of a $16-17$ \msun\ progenitor as computed by \citet{Woosley2007}. Using a $\beta_{5577}/\beta_{6300,6364}$ ratio of 2 instead of 1.5 shifts the oxygen mass range to $0.6-1.2$ \msun\ ($M_{\rm ZAMS}=17-19\ M_\odot$), and using a ratio of 1 gives $0.3-0.7$ \msun\ ($M_{\rm ZAMS}=14-17\ M_\odot$). The oxygen mass in the O/Ne/Mg layer of the 15 \msun\ progenitor star is 0.3 \msun, on the lower end of these estimates, for the reasons discussed above (some combination of departure from LTE and a too strong mixing between $^{56}$Ni and oxygen clumps in the model). The power of these analytical arguments is that the [O I] \wl 6300/[O I] \wl 6364 line ratio combined with the [O I] \wl 5577/[O I] \wll 6300,~6364 line ratio provide checks on the density and energy deposition for the oxygen clumps in any model. Since breakdown of the LTE assumption can only lead to overestimated oxygen masses using Eqs. \ref{eq:ratio} and \ref{eq:oimasses}, it would be difficult to reconcile any ejecta which has more than 1 \msun\ of oxygen in the O/Ne/Mg zone with the observed oxygen line strengths, \emph{independent} of how the oxygen is mixed hydrodynamically and compositionally. For example, at $M_{\rm ZAMS}= 20$ \msun, the oxygen mass in the O/Ne/Mg layer is over 1.5 \msun \ in the \citet{Woosley2007} models, and the spectrum from such an ejecta would be inconsistent with the observed line strengths of SN 2012aw given the constraints on temperature derived here (see J12 for model spectra from such a high-mass star). It also seems clear that a hypothetical inmixing of coolants like calcium, carbon, silicon, CO, or SiO into the O/Ne/Mg zone cannot save such a model; apart from the temperature constraints considered above, Figs. \ref{fig:optical} and \ref{fig:NIR} show that there is not sufficient observed luminosity from either of these elements ([Ca II] \wll 7291, 7323, Ca II \wl \wl 8498, 8542, 8662, [C I] \wl8727, [C I] \wll9824, 9850, [Si I] \wl1.10 $\upmu$m, [Si I] \wll 1.60, 1.64 $\upmu$m, CO overtone band), to account for any significant cooling of a large oxygen mass. For SN 2004et, which had a very similar spectral evolution in the optical and NIR, observations in the MIR also ruled out CO and SiO fundamental bands as strong enough cooling channels for such a scenario. The models here therefore represent calculations with no obvious discrepancy with observations regarding the thermal emission from the oxygen zones. We find, both here for SN 2012aw and in J12 for SN 2004et, an amount of thermally emitting oxygen $<1$ \msun. As the oxygen mass rises steeply with progenitor mass, this sets a constraint on the upper mass for the progenitor star model of 18 \msun, assuming nucleosynthesis as calculated by \citet{Woosley2007}. \begin{table*} \centering \caption{The line luminosities of [O I] \wl 5577 and [O I] \wll 6300, 6364, their ratio, the single-zone LTE temperature corresponding to this ratio, and the LTE O I mass corresponding to this temperature. For all line luminosity measurements we estimate a $\pm$30\% (relative) error in [O I] \wl5577 and a $\pm$20\% error in [O I] \wll 6300, 6364.} \begin{tabular}{|c|c|c|c|c|c|} \hline Time & $L_{5577}$ & $L_{6300, 6364}$ & $L_{5577}/L_{6300,6364}$ & $T^{LTE}_{\beta_{ratio}=1.5}$ & $M(O I)^{LTE}_{\beta_{ratio}=1.5,\beta_{6300,6364}=0.5}$\\ (days) & ($10^{38}$ erg s$^{-1}$) & ($10^{38}$ erg s$^{-1}$) & & (K) & (\msun)\\ \hline 250 & $1.5\pm 0.45$ & $13 \pm{2.6}$ & $0.12 \pm 0.042$ & $4170_{-280}^{+220}$ & $0.6_{-0.2}^{+0.3}$\\ 332 & $0.68\pm 0.20$ & $8.6 \pm 1.7$ & $0.079 \pm 0.029$ & $3920_{-250}^{+190}$ & $0.6_{-0.1}^{+0.3}$\\ 369 & $0.39\pm 0.12$ & $6.7 \pm 1.4$ & $0.057 \pm 0.021$ & $3740_{-230}^{+170}$ & $0.6_{-0.2}^{+0.3}$ \\ 451 & $0.082\pm 0.025$ & $3.3 \pm {0.66}$ & $0.025 \pm 0.0090$ & $3330_{-180}^{+140}$ & $0.6_{-0.1}^{+0.3}$ \\ \hline \end{tabular} \label{table:oilines} \end{table*} The only other clearly detected oxygen line is O I \wl 1.130 $\upmu$m (O I \wl7771 may be detected in the 250-day spectrum but is not to be confused with the (unidentified)\footnote{A plausible identification of this line is the ground state resonance line of potassium - K I 4s($^2$S)-4p($^2$P$^o$) \wll 7665, 7699 \citep{Spyromilio1991b} which is not included in our model.} line seen at $\sim$7700 \AA). The O I \wl 1.130 $\upmu$m line is radiatively pumped by Ly$\beta$ line overlap with O I \wl1025.76, much of which is zone-crossing between hydrogen clumps and oxygen clumps\footnote{The Ly$\beta$ photon is created in a hydrogen clump but performs a random walk into an oxygen clump while still resonantly trapped. This leads to a breakdown of the Sobolev approximation where all physical conditions are assumed to stay constant over the resonance region.} \citep[][J12]{Oliva1993}. Our model treats the \emph{local} line-overlap in this line, producing an emission feature from the primordial oxygen in the hydrogen clumps (Fig. \ref{fig:NIR}). As is clear from Fig. \ref{fig:NIR} though, this is not enough to reproduce the line luminosity, in agreement with the analysis of this line in SN 1987A \citep{Oliva1993}. That the [O I] \wl 1.130 $\upmu$m line is not a recombination line can be inferred from the lack of similar strength emission from other lines that are expected to have similar recombination luminosities; O I \wl7771, O I \wl9263 and O I \wl 1.316 $\upmu$m. The oxygen recombination lines are all observationally weak or absent. In the model this arises as many O II ions are neutralized by rapid charge transfer reactions, with for example C I and Mg I, rather than by radiative recombination. These charge transfer reactions usually occur to the ground state or first excited state in O I, thereby preventing the radiative cascade that follows radiative recombinations to higher levels. Two other important elements reside in the carbon burning ashes; magnesium and sodium. A successful model must be able to account for the observed luminosities in these lines as well as in the oxygen lines, but as discussed in J12, these lines are more challenging to model accurately. Fig. \ref{fig:linefluxes} shows the measured evolution in the line luminosities of [O I] \wll6300, 6364, Mg I] \wl4571, and Na I-D, compared to the models for 12, 15, 19 and 25 \msun~progenitors. \begin{figure*} \includegraphics[trim=1mm 1mm 1mm 1mm, clip, width=0.8\linewidth]{fig3.eps} \caption{Observed luminosities of [O I] \wll6300, 6364, Mg I] \wl4571, and Na I-D lines compared to models for 12 (blue, dot-dashed), 15 (black, solid), 19 (green, dashed), and 25 \msun\ (magenta, dotted) progenitors. The line fluxes are extracted by an automated algorithm applied in the same way to observed and modeled spectra, as described in J12. We estimate the statistical errors by (quadrature) adding the RMS error of the photometric flux calibration (Sect. \ref{sec:obs}) to the estimated error from the continuum fit in the algorithm (taken as $\pm$20\% here).} \label{fig:linefluxes} \end{figure*} As seen in Fig. \ref{fig:linefluxes}, the observed Mg I] \wl4571 and Na I-D line strengths are bracketed by the $15-19$ \msun\ model range, which overlaps with the 14 - 18 \msun \ mass range determined from the oxygen lines. The 25 \msun\ model shows some interesting behaviour. It has an O/Ne/Mg zone of 2.7 \msun, compared to 1.9 \msun\ of the 19 \msun\ model and 0.45 \msun\ of the 15 \msun\ model. There is thus a more dramatic increase in oxygen production between 15 and 19 \msun\ than between 19 and 25 \msun, which is reflected in the oxygen line luminosities. The 25 \msun\ model still produces brighter O I lines, but only by $\sim$20\% between $140-350$ days and almost not at all after that. In addition to the smaller difference in oxygen production, the metal content in the core of the 25 \msun\ ejecta starts to take up such a large fraction of the volume (50\%), that some line blocking occurs even around 6300 \AA, further reducing the line luminosity. The Mg I] 4571 line luminosity increases more linearly throughout the $12-25$ \msun\ range, and is therefore an important complement to the oxygen lines for the analysis. Finally, the Na I-D lines are actually weaker in the 25 \msun\ model compared to the 19 \msun\ model. This is because sodium nucleosynthesis is not a strictly monotonic function of progenitor mass, and there is less synthesized sodium in the 25 \msun\ ejecta ($6\e{-3}$ \msun) than in the 19 \msun\ ejecta ($1.3\e{-2}$ \msun) In the NIR, the model reproduces all the distinct emission lines qualitatively, which includes lines from H I, He I, C I, N I, O I, Na I, Mg I, Si I, S II, Fe I, Fe II, Co II, as labeled in Fig. \ref{fig:NIR}. The model predicts a strong He I \wl2.058 $\upmu$m line which is not observed (it was also absent/weak in SN 2004et), as well as a too strong He I \wl 1.083 $\upmu$m line. Of the NIR lines, J12 found C I \wl1.454 $\upmu$m, Mg I \wl1.504 $\upmu$m and [Si I] \wll1.607, 1.645 $\upmu$m to be the lines most sensitive to the nucleosynthesis. In general these lines are in rough agreement with the 15 \msun~model (Fig. \ref{fig:NIR}), but we cannot with confidence distinguish between different models in the $12-19$ \msun\ range. For instance the [Si I] \wll 1.607, 1.645 $\upmu$m lines are in best agreement with a 12 \msun\ model whereas Mg I \wl 1.504 $\upmu$m is in best agreement with a 19 \msun\ model. The uncertainty in the absolute flux calibration of the NIR spectrum (as NIR photometry is lacking) should also be kept in mind. All observed line fluxes appear weaker than in the 19 \msun\ model (see J12 for the 19 \msun\ spectrum), and the NIR spectrum constraints therefore agree with the optical constraints in ruling out a high-mass progenitor. \begin{figure} \includegraphics[trim=16mm 3mm 8mm 10mm, clip, width=1\linewidth]{myplot-endpoints.eps} \caption{The final pre-SN luminosity as function of stellar mass, from three sets of stellar evolution models discussed in the text. The KEPLER models are those used as input for the spectral modelling in this paper. The range of luminosities of the progenitor star of SN 2012aw calculated by \citet*{Kochanek2012} ($\log{L/L_\odot} = 4.8-5.0$) are shown, corresponding to the end point luminosities of $13-16$ \msun\ KEPLER progenitors.} \label{fig:endpoints} \end{figure} \section{Discussion} \label{sec:discussion} The link between progenitor mass and nucleosynthesis depends on how some uncertain physical processes are treated in the stellar evolution model. Whereas mass-loss, convection, and metallicity have weak influence on the oxygen nucleosynthesis \citep{Dessart2013}, semi-convection \citep{Langer1991}, overshooting \citep{Langer1991, Schaller1992, Dessart2013} and rotation \citep{Hirschi2004, Dessart2013} can alter the oxygen nucleosynthesis by more than a factor of two. The KEPLER models we use employ efficient semi-convection \citep[see discussions in][]{Langer1985, Woosley1988, Langer1991} and overshooting \citep[see][]{Woosley1988}, but no rotation. Models with overshooting produce about twice as much oxygen as models without and fast-rotating models produce a factor $\sim$1.5 more oxygen than non-rotating ones \citep{Dessart2013}. These effects are also reflected in the stellar luminosities, with non-rotating models with efficient overshooting tending to match the luminosities of rotating models \citep[see discussion in][]{Smartt2009b}. Reviewing a variety of models for nucleosynthesis yields \citep{Thielemann1996, Nomoto1997, Limongi2003, Hirschi2005, Woosley2007}, the factor two variety in oxygen yield for a given ZAMS mass is confirmed. For instance, at $M_{\rm ZAMS}=15$ \msun, the ejected masses vary from $0.4-1.0$ \msun, and at $M_{\rm ZAMS}=25$ \msun from $2.2-3.6$ \msun. This latter range is clearly higher than the oxygen mass in SN 2012aw derived here. Whereas these issues in stellar evolution physics still remain to be resolved, any model can be checked for its ability to reproduce \emph{both} the progenitor appearance and the derived nucleosynthesis of an individual event. As described in Sect. 3, the KEPLER supernova progenitor models with masses in the range $M_{\rm ZAMS} = 14-18$ \msun\ have nucleosynthesis that satisfactorily reproduces the observed nebular spectrum of SN 2012aw. The model progenitor luminosities from this range can then be compared with the derived luminosity of the progenitor star. The \citet{Woosley2007} final luminosities (from private communication with S. Woosley) are shown in Fig.\,\ref{fig:endpoints}, along with those from the STARS models \citep{Eldridge2004}\footnote{http://www.ast.cam.ac.uk/$\sim$stars/archive/} and the Geneva rotating models \citep{Ekstrom2012}. The KEPLER model luminosities are at the pre-SN silicon burning stage, whereas the other two are at the end of core C-burning, but the surface luminosity of a red supergiant is not expected to change after this stage. The progenitor star of SN 2012aw was originally identified from optical and near-infrared images by \cite{Fraser2012} and \cite{vandyk2012} as a red supergiant with luminosity of $\log \left(L/L_\odot\right) = 5.0 - 5.6$ and $\log \left(L/L_\odot\right) = 5.18-5.24$, respectively. Both papers argued for significant extinction of the progenitor, most likely due to circumstellar dust that was later destroyed in the explosion, as the SN itself had a low line of sight extinction. An extended reanalysis of the progenitor photometry was carried out by \citet{Kochanek2012} who performed a more detailed calculation of the extinction by computing the transfer through a realistic circumstellar density distribution and computing optical properties of the dust (in comparison, \citet{Fraser2012} and \citet{vandyk2012} treated the dust as a distant dust screen which removes all photons that scatter on it from the observer). With the \citet{Kochanek2012} treatment, the derived stellar luminosity is lowered by several tenths of a dex, as many photons scatter on the circumstellar dust but still reach the observer, and the authors find the observed progenitor SED to be consistent with a red supergiant star with $\log \left(L/L_\odot\right) = 4.8-5.0$. As shown in Fig. \ref{fig:endpoints}, this progenitor luminosity corresponds to a KEPLER progenitor in the $13-16$ \msun\ range, consistent with the results from the nebular analysis in this paper. \citet{Kochanek2012} also noted that the X-ray and radio observations of SN 2012aw constrained the progenitor's mass-loss rate to $-\log \left({\dot{M}/M_\odot \mbox{yr}^{-1}}\right) = 5.5 - 5.0$. The KEPLER models employ mass loss rates from the empirical parametrisation of \cite{Nieu1990} which are $-\log \left({\dot{M}/M_\odot \mbox{yr}^{-1}}\right) = 5.4-5.2$ for $\log \left(L/L_\odot\right) = 5.0$, $M_{\rm ZAMS} = 15$ \msun\ and $3500\ \mbox{K} < T_{\rm eff} < 4500\ \mbox{K}$. There is therefore encouraging consistency between the required mass-loss rate from X-ray and radio observations, the progenitor luminosity from pre-explosion imaging, and the nucleosynthesis derived from the nebular spectra. However, hydrodynamical modeling of the optically thick phase \citep{dallOra2013} favors an ejecta mass $>20$ \msun, and thus a higher-mass progenitor. Understanding the differences in results between progenitor imaging, hydrodynamical modeling, and nebular phase spectral analysis is a high priority in the Type IIP research field. We hope in the future to continually increase the accuracy of our nebular-phase spectral models, and apply them to grids of multidimensional explosions with self-consistently calculated mixing. Combined with further high signal-to-noise optical and near-infrared nebular phase spectra of nearby SNe, we will eventually be able to constrain the yields of individual elements to still higher accuracy and improve our understanding of stellar and supernova nucleosynthesis. \section{Conclusions} We have obtained optical and near-infrared spectra of the Type IIP SN 2012aw in the nebular phase, and analyzed these with spectral model calculations. The observed spectral evolution shows close agreement with a model for a $M_{\rm ZAMS}= 15\ M_\odot$ progenitor (evolved and exploded with KEPLER), with additional analysis of magnesium and sodium emission lines consistently pointing to a progenitor in the $14-18$ \msun\ range. The endpoint luminosities of $14-16$ \msun\ models are also consistent with the estimate of the luminosity of the directly detected progenitor star by \citet{Kochanek2012}, and hence a self-consistent solution is found. We have demonstrated how combining the individual evolution in the three oxygen lines excited by thermal collisions ([O I] \wl5577, [O I] \wl 6300, and [O I] \wl6364) can be used to constrain the mass of the oxygen present in the carbon burning ashes (here limiting it to $<$ 1 \msun), independent of uncertainties in the hydrodynamical and chemical mixing. Reviewing the literature of published nebular spectra of Type IIP SNe (with ``normal'' explosion energies and $^{56}$Ni masses)\footnote{Some objects in the ``subluminous'' class of Type IIP SNe \citep[e.g.][]{Turatto1998, Benetti2001, Pastorello2004, Fraser2011}, which have extremely low kinetic energies and $^{56}$Ni masses, may still be associated with such massive stars (although those with directly identified progenitors so far have $M_{ZAMS}<12$ \msun), but a viewpoint on this from their nebular spectra would need models calculations with significantly different parameters than those used here.} \citep{Turatto1993, Schmidt1993, Benetti1994, Elmhamdi2003, Pozzo2006, Sahu2006, Quimby2007, Maguire2010, Andrews2011, Inserra2011, Roy2011, Tomasella2013}, we find no observations where the [O I] \wll 6300, 6364 lines are significantly stronger (relative to the optical spectrum as a whole) than in SN 2012aw. We conclude that no Type IIP SN has yet been shown to eject nucleosynthesis products expected from stars more massive than 20 \msun. \section*{Acknowledgments} The research leading to these results has received funding from the European Research Council under the European Union's Seventh Framework Programme (FP7/2007-2013)/ERC Grant agreement n$^{\rm o}$ [291222] (PI : S. J. Smartt). The results are based on observations made with the William Herschel Telescope (PATT and Service time, proposal ID SW2012b27) and the Nordic Optical Telescope operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof\'{i}sica de Canarias. We thank Cosimo Inserra, Ting-Wan Chen, and Elisabeth Gall for discussion and observing assistance, and Andy Lawrence, Alex Bruce, Massimo Dall'Ora, Stan Woosley, and Christopher Kochanek for discussion. \bibliographystyle{mn2e3}
\section{Introduction} \setcounter{equation}{0} \label{s1} Kuhn-Tucker optimality conditions and Fritz John ones for scalar and vector nonsmooth problems with inequality constraints are among the most important directions of investigation in optimization. In the present paper we deal with optimality conditions of Kuhn-Tucker type for the following problem: \medskip Minimize$\quad f(x)\quad$subject to$\quad x\in X,\; g(x)\leqq0$,\hfill (P) \medskip \noindent where $f:X\to\mathbb R^n$ and $g:X\to\mathbb R^m$ are given vector real-valued functions, defined on an open set $X\subset\mathbb R^s$. Denote by $S$ the set of feasible points, that is $S:=\{x\in X\mid g(x)\leqq 0\}$. All results given here are obtained for nonsmooth problems in terms of the usual second-order directional derivative. For every vector $\lambda$ from the set \[ \Lambda:=\{\lambda=(\lambda_1,\lambda_2,\ldots,\lambda_n)\in\mathbb R^n\mid\sum_{i=1}^n\lambda_i=1,\;\lambda_i\geqq 0,\; i=1,2,\ldots, n\} \] we consider the following weighting scalar problem: \medskip Minimize$\quad\scalpr{\lambda}{f(x)}\quad$subject to$\quad x\in X,\; g(x)\leqq0$.\hfill ${\rm (P_\lambda)}$ \medskip \noindent Here we denote by $\scalpr{a}{b}$ the scalar product of the vectors $a$ and $b$. For every $\lambda\in\Lambda$ we define the scalar Lagrange function \[ L_\lambda(x,\mu)=\scalpr{\lambda}{f(x)}+\scalpr{\mu}{g(x)}. \] For all vectors $\lambda\in\Lambda$ and $\mu\geqq 0$ we consider the following unconstrained problem \medskip Minimize$\quad\scalpr{\lambda}{f(x)}+\scalpr{\mu}{g(x)}\quad$subject to$\quad x\in X$. \hfill ${\rm (P_{\lambda,\mu})}$ \medskip These problems are formulated and discussed by Geoffrion \cite{geo68}. It is easy to prove the following claims: \begin{proposition} Let $\bar x\in S$ be a global solution of ${\rm (P_\lambda)}$ for some $\lambda\in\Lambda$. Then $\bar x$ is a weakly efficient solution of {\rm (P)}. \end{proposition} \begin{proposition} Let $\bar x\in S$ be a global minimizer of ${\rm (P_{\lambda,\mu})}$ for some $\lambda\in\Lambda$ and $\mu\geqq 0$ with $\mu_j g_j(\bar x)=0$ for all $j=1,2,\dots,m$. Then $\bar x$ is a global solution solution of ${\rm (P_\lambda)}$. \end{proposition} \begin{proposition}\label{pr3} Let $\bar x\in S$ be a global minimizer of ${\rm (P_{\lambda,\bar\mu})}$ for some $\lambda\in\Lambda$ and $\bar\mu\geqq 0$ with $\bar\mu_j g_j(\bar x)=0$. Then the pair $(\bar x,\bar\mu)$ is a Kuhn-Tucker saddle point of ${\rm L_\lambda}$. \end{proposition} \begin{proposition}\label{pr4} Let the pair $(\bar x,\bar\mu)\in S\times [0,+\infty)^m$ be a Kuhn-Tucker saddle point of ${\rm L_\lambda}$ for some $\lambda\in\Lambda$. Then $\bar x\in S$ is a global minimizer of ${\rm (P_{\lambda,\bar\mu})}$ and $\bar\mu_j g_j(\bar x)=0$. \end{proposition} Denote the solution sets of (P), (P$_\lambda$) and (P$_{\lambda,\mu}$) respectively by WE$\,$(P), Argmin$\,$(P$_\lambda$) and Argmin$\,$(P$_{\lambda,\mu}$), and the set of points which satisfy Kuhn-Tucker necessary optimality conditions by KT$\,$(P). We have the following implications under additional assumptions: \[ \bar x\in {\rm Argmin\,(P_{\lambda,\mu})}\;\Rightarrow\; \bar x\in {\rm Argmin\,(P_\lambda)}\;\Rightarrow\;\bar x\in {\rm WE\,(P)} \;\Rightarrow\; \bar x\in {\rm KT\,(P)}. \] It is interesting when the following converse conditions are satisfied: 1) Every Kuhn-Tucker stationary point is a saddle point of the Lagrange function. 2) Every Kuhn-Tucker stationary point is a weak global Pareto minimizer. 3) Every Kuhn-Tucker stationary point is a solution of the weighting problem. We answer the first question in the first- and second-order cases. The answers of the second question and the third one are known. \cite{ara08,osu99,osu98} In this paper we find the largest class of functions such that every second-order Kuhn-Tucker stationary point is weakly efficient, and when it is a solution of weighting problem. We define the following new classes of vector problems: KTSP-invex, second-order KTSP-invex, second-order KT-pseudoinvex I, second-order KT-pseudoinvex II, second-order KT-invex. We prove that in contrast of the second-order generalized invex vector problems, applied in Duality Theory, the classes of second-order generalized invex vector problems, introduced in the paper, include the respective class of first-order generalized invex vector problem. In Section \ref{s2} we prove a theorem of the alternative, which we apply later several times. In Section \ref{s3} we derive second-order necessary conditions for a local minimum of problems with C$^1$ (i.e. continuously differentiable) data. The necessary conditions are not enough to determine the minimizer. In the next sections we show that the same conditions without constraint qualification are sufficient for a global minimum in some problems of generalized convexity type. In Section \ref{s6} we introduce notions of KTSP-invex and second-order KTSP-invex problems with inequality constraints. We prove that these classes are the largest ones with the following properties: every (second-order) Kuhn-Tucker stationary point is saddle for the respective scalar Lagrange function. In other words we find the largest classes of problems (P) such that the saddle points of the Lagrange function and the (second-order) Kuhn-Tucker stationary points coincide with the weakly efficient solution of (P). In Section \ref{s4} we define a notion of a second-order KT-pseudoinvex-I vector problem with inequality constraints. We obtain that these problems are the largest class such that every second-order Kuhn-Tucker stationary point is a weak global Pareto minimizer of (P). At last, in Section \ref{s5} we define a notion of a second-order invex vector problem with inequality constraints. We prove that these problems are the largest class such that every second-order Kuhn-Tucker stationary point is a global solution of the weighting problem. In other words we find the largest class of vector problems with inequality constraints such that the set of weakly efficient global solutions of (P) coincides with the global solutions of the weighting problem and the second-order Kuhn-Tucker stationary points. The notions of invexity that we study in this paper are generalizations of some notions for scalar and vector problems \cite{ara08,osu99,osu98,mar85,JOGO,Optimization,OptLett-2}. \section{A Theorem of the Alternative} \label{s2}\setcounter{equation}{0} Throughout this paper we use the following notations comparing the vectors $x$ and $y$ with components $x_i$ and $y_i$ in finite-dimensional spaces: \[ x>y\quad\textrm{iff}\quad x_i>y_i\quad\textrm{for all indeces }i; \] \[ x\geqq y\quad\textrm{iff}\quad x_i\geqq y_i\quad\textrm{for all indeces }i; \] \[ x\ge y\quad\textrm{iff}\quad x_i\geqq y_i\quad\textrm{for all indeces }i,\; x\ne y. \] We denote by $e$ the vector $e=(1,1,\ldots,1)$, and by $A^T$ the transpose of the matrix $A$. We apply several times the following lemma: \begin{lemma}\label{th-alternative} Let $A$, $B$, $C$, and $D$ be given matrices. Then either the system \begin{equation}\label{7} A^Tx+C^Tu<0,\quad B^Tx+D^Tu\leqq 0,\quad u\geqq 0 \end{equation} has a solution $(x,u)$ or the system \begin{equation}\label{8} y^TA+z^TB=0,\quad y^TC+z^TD\geqq 0,\quad y\ge 0,\quad z\geqq 0 \end{equation} has a solution $(y,z)$, but never both. \end{lemma} \begin{proof} Consider the following primal linear programming problem \medskip \begin{tabular}{ll} Minimize & 0 \\ subject to & $y^TA+z^TB=0$ \\ & $y^TC+z^TD\geqq 0$ \\ & $\scalpr{y}{e}=1$ \\ & $y\geqq 0,\quad z\geqq 0$ \end{tabular} \medskip \noindent and its dual one \medskip \begin{tabular}{lll} Maximize & $v$ & \\ subject to & $A^Tx+C^Tu+ve$ & $\leqq 0$ \\ & $B^Tx+D^Tu$ & $\leqq 0$ \\ & $u\geqq 0$. & \end{tabular} \medskip Suppose that the system (\ref{7}) has no solutions. We prove that the system (\ref{8}) is solvable. The dual problem is feasible, because $x=0$, $u=0$, $v=0$ is a feasible point. It follows from the assumption (\ref{7}) is unsolvable that there is no a number $v>0$ and vectors $x$ and $u$ such that the triple $(x,u,v)$ is feasible for the dual problem. Therefore, the dual problem is solvable and its optimal value is 0. According to Duality Theorem in Linear Programming the primal problem is solvable with the same optimal value. Hence, the system (\ref{8}) is solvable. Suppose that (\ref{7}) is solvable. We prove that (\ref{8}) has no solutions. Assume the contrary, that is (\ref{8}) is solvable. Therefore, the primal problem is solvable. It follows from Duality Theorem that the dual problem is also solvable and its optimal value is 0. Therefore, there is no a positive number $v$ and vectors $x$, $u$ such that the triple $(x,u,v)$ is feasible for the dual problem, which is a contradiction, because we suppose that (\ref{7}) is solvable. \end{proof} Motzkin's, Farkas', and Gordan's theorems of the alternative are particular cases of Lemma \ref{th-alternative} (see \cite{man69}). \section{Second-Order Necessary Conditions for \\ Weak Local Minimum} \label{s3}\setcounter{equation}{0} In this section we derive necessary optimality conditions for the problem {\rm{(P)}} with continuously differentiable data. We begin this section with some preliminary definitions. \smallskip Denote by $\mathbb R$ the set of reals and $\overline{\mathbb R}=\mathbb R\cup\{-\infty\}\cup\{+\infty\}$. Let the function $h:X\to\mathbb R$ with an open domain $X\subset\mathbb R^s$ be differentiable at the point $x\in X$. Then the second-order directional derivative $h^{\prime\pr}(x,u)$ of $h$ at the point $x\in X$ in direction $u\in\mathbb R^n$ is defined as an element of $\overline{\mathbb R}$ by \begin{displaymath} h^{\prime\pr}(x,u)=\lim_{t\to +0}\,2\,t^{-2}\,[h(x+tu)-h(x)-t\nabla h(x)u]. \end{displaymath} The function $h$ is called second-order directionally differentiable at the point $x$ if the derivative $h^{\prime\pr}(x,u)$ exists for each direction $u\in\mathbb R^n$ and $-\infty<h^{\prime\pr}(x,u)<+\infty$. We say that $h^{\prime\pr}(x,u)$ exists, if it is finite. Consider the problem {\rm{(P)}}. For every feasible point $x\in S$ let $A(x)$ be the set of active constraints \[ A(x):=\{j\in\{1,2,\ldots,m\}\mid g_j(x)=0\}. \] \begin{definition} A direction $d$ is called critical at the point $x\in S$ if \[ \nabla f_i(x)d\leqq0,\; i=1,2,\ldots,n,\,\quad \nabla g_j(x)d\leqq 0,\;\forall j\in A(x). \] For every critical direction $d$ at the feasible point $x$ denote by $I(x,d)$ and $J(x,d)$ the following index sets: \[ I(x,d):=\{i\in\{1,2,\ldots,n\}\mid\nabla f_i(x)d=0\}, \] \[ J(x,d):=\{j\in A(x)\mid\nabla g_j(x)d=0\}. \] \end{definition} \begin{definition} A feasible point $\bar x\in S$ is called a (weak) local Pareto minimizer, or (weakly) efficient if there exists a neigbourhood $U\ni\bar x$ such that there is no $x\in U\cap S$ with $f(x)\le f(\bar x)$ ($f(x)<f(\bar x)$). The point $\bar x\in S$ is called a (weak) global Pareto minimizer if there does not exist $x\in S$ with $f(x)\le f(\bar x)$ ($f(x)<f(\bar x)$). \end{definition} The problem (P) is said to be Fr\'echet differentiable if $f$ and $g$ are Fr\'echet differentiable. The following result, which contains first-order necessary optimality conditions under various constraint qualifications, is known as Kuhn-Tucker Theorem: \begin{proposition} Let the problem {\rm (P)} be Fr\'echet differentiable on some open set $X\subseteq\mathbb R^s$. Suppose that $\bar x$ is a weak local Pareto minimizer, and any of the constraint qualifications holds. Then there exist vectors $\lambda=(\lambda_1,\lambda_2,\ldots,\lambda_n)\ge 0$ and $\mu=(\mu_1,\mu_2,\ldots,\mu_m)\geqq 0$ such that \[ \sum_{i=1}^n\lambda_i\nabla f_i(\bar x)+\sum_{j=1}^m\mu_j\nabla g_j(\bar x)=0,\;\lambda_j\, g_j(\bar x)=0,\; j=1,2,...,m. \] \end{proposition} The next two theorems are generalizations of the primal and dual necessary optimality conditions by Ginchev, Ivanov \cite{JMAA2008} where scalar problems with ineuality constraints are treated. \begin{theorem}\label{th1} Let $X$ be an open set in the space $\mathbb R^s$, the functions \[ f_i\; (i=1,2,\ldots,m),\quad g_j\; (j=1,2,\ldots,m) \] be defined on $X$. Suppose that $\bar x$ is a weak local Pareto minimizer of the problem {\rm{(P)}}, the functions $g_j$ $(j\notin A(\bar x))$ are continuous at $\bar x$, the functions $f$ and $g_j$ $(j\in\ A(\bar x))$ are continuously differentiable, and the functions $f_i$ $(i\in I(\bar x,d))$ and $g_j$ $(j\in J(\bar x,d))$ are second-order directionally differentiable at $\bar x$ in any critical direction $d\in\mathbb R^s$. Then for every critical direction $d\in\mathbb R^s$, it follows that there is no $z\in\mathbb R^s$ which solves the system \begin{equation}\label{1} \nabla f_i(\bar x)z+f_i^{\prime\pr}(\bar x,d)<0,\quad i\in I(\bar x,d). \end{equation} \begin{equation}\label{2} \nabla g_j(\bar x)z+g_j^{\prime\pr}(\bar x,d)<0,\quad j\in J(\bar x,d). \end{equation} \end{theorem} \begin{proof} Suppose the contrary that there exists a critical direction $d$ such that the system (\ref{1}), (\ref{2}) has a solution $z\in\mathbb R^s$. Obviously, the case $I(\bar x,d)\cup J(\bar x,d)=\emptyset$ is impossible, because $\bar x$ is a weak minimimer. Let $i\in \{1,2,\ldots,n\}$ be arbitrary fixed. Consider the function of one variable \[ \varphi_i(t)=f_i(\bar x+td+0.5t^2z). \] Since $X$ is open and $\bar x$ is feasible, there exists $\varepsilon_i>0$ such that $\varphi_i$ is defined for $0\leqq t<\varepsilon_i$. We have \[ \varphi_i^\prime(t)=\nabla f_i(\bar x+td+0.5t^2z)(d+tz). \] Therefore, $\varphi_i^\prime(0)=\nabla f_i(\bar x)d$. Consider the differential quotient \[ 2\,t^{-2}\,[\varphi_i(t)-\varphi_i(0)-t\varphi^\prime_i(0)] =2\,t^{-2}\,[f_i(\bar x+td+0.5t^2 z)-f_i(\bar x)-t\nabla f_i(\bar x)d]. \] According to the Mean-Value Theorem there exists $\theta_i\in(0,1)$ such that \[ f_i(\bar x+td+0.5t^2 z)=f_i(\bar x+td)+\nabla f_i(\bar x+td+0.5t^2\theta_i z)(0.5t^2 z). \] By $f_i\in\rm{C}^1$, we obtain that there exists the second-order directional derivative $\varphi_i^{\prime\pr}(0,1)$ and \[ \nabla f_i(\bar x)z+f_i^{\prime\pr}(\bar x,d)= \lim_{t\to+0}\nabla f_i(\bar x+td+0.5t^2\theta_i z)z \] \[ +\lim_{t\to+0}\;2t^{-2}(f_i(\bar x+td)-f_i(\bar x)- t\nabla f_i(\bar x)d)=\varphi_i^{\prime\pr}(0,1). \] Since $z$ is a solution of the system (\ref{1}) with a direction $d$, we conclude that for every $i\in \{1,2,\ldots,n\}$ there exists $\varepsilon_i>0$ such that \[ \varphi_i(t)-\varphi_i(0)-t\varphi^\prime_i(0))<0\quad {\rm for all}\quad t\in(0,\varepsilon_i) \] that is \begin{equation}\label{3} f_i(\bar x+t d+0.5t^2 z)-f_i(\bar x)-t\nabla f_i(\bar x)d<0 \quad\textrm{ for all }\quad t\in(0,\varepsilon_i). \end{equation} Consider the function of one variable \[ \psi_j(t)=g_j(\bar x+td+0.5t^2z). \] Using similar arguments we prove that there exists $\delta_j>0$, $j=1,2,\ldots, m$ with \begin{equation}\label{5} g_j(\bar x+t d+0.5t^2 z)-g_j(\bar x)-t\nabla g_j(\bar x)d<0 \quad\textrm{ for all }\quad t\in(0,\delta_j). \end{equation} Consider the following cases: 1) For every $j\in\{1,2,...,m\}\setminus A(\bar x)$ we have $g_j(\bar x)<0$. Hence, by continuity, there exists $\delta_j>0$ such that $g_j(\bar x+t d+0.5t^2 z)<0$ for all $t\in[0,\delta_j)$. 2) For every $j\in A(\bar x)\setminus J(\bar x,d)$ we have $\nabla g_j(\bar x)d= \psi_j^\prime(0)<0$. Therefore, there exists $\delta_j>0$ such that $\psi_j(t)<\psi_j(0)$ for all $t\in(0,\delta_j)$. Hence we have $g_j(\bar x+t d+0.5t^2 z)<g_j(\bar x)=0$ for all $t\in(0,\delta_j)$. 3) For all $j\in J(\bar x,d)$, by $\nabla g_j(\bar x)d=0$, it follows from (\ref{5}) that there exist $\delta_j>0$ such that $g_j(\bar x +t d+0.5t^2 z)<g_j(\bar x)=0$ for all $t\in(0,\delta_j)$. 4) For all $i\notin I(\bar x,d)$ we have $\nabla f_i(\bar x)d<0$ that is $\varphi_i^\prime(0)<0$ and therefore, for some $\varepsilon_i>0$ it holds $f_i(\bar x+t d+0.5t^2 z)<f_i(\bar x)$ for all $t\in(0,\varepsilon_i)$. 5) For all $i\in I(\bar x,d)$ we have $\nabla f_i(\bar x)d=0$ and according to (\ref{3}) there exists $\varepsilon_i>0$ such that $f_i(\bar x +t d+0.5t^2 z)<f_i(\bar x)$ for all $t\in(0,\varepsilon_i)$. It is seen that $\bar x$ is not a weak local minimizer, contradicting our hypothesis. \end{proof} \begin{theorem}\label{NKT} If all hypotheses of Theorem \ref{th1} are satisfied, then corresponding to any critical direction $d$ there exist nonnegative multipliers \[ \lambda=(\lambda_1,\lambda_2,\ldots,\lambda_n)\quad\textrm{and}\quad \mu=(\mu_1,\mu_2,\ldots,\mu_m)\quad\textrm{with}\quad (\lambda,\mu)\ne 0 \] such that \begin{equation}\label{6} \mu_j g_j(\bar x)=0,\;j=1,2,\ldots,m,\quad\nabla L(\bar x)=0, \end{equation} \begin{equation}\label{4} \lambda_i\nabla f_i(\bar x)d=0,\quad i=1,2,\ldots,n,\quad\mu_j\nabla g_j(\bar x)d=0,\quad j=1,2,\ldots,m, \end{equation} \begin{equation}\label{10} L^{\prime\pr}(\bar x,d)=\sum_{i=1}^n\lambda_i f_i^{\prime\pr}(\bar x,d)+ \sum_{j\in A(\bar x)}\mu_j g_j^{\prime\pr}(\bar x,d)\geqq 0, \end{equation} where $L=\scalpr{\lambda}{f}+\scalpr{\mu}{g}$ is the Lagrange function. Suppose further that a constraint qualification holds. Then we have $\lambda\ne 0$. \end{theorem} \begin{proof} Consider the matrix $A$ whose columns are \[ \{\nabla f_i(\bar x)\mid i\in I(\bar x,d)\}\quad\textrm{and}\quad \{\nabla g_j(\bar x)\mid j\in J(\bar x,d)\}, \] and the row-matrix $C$ whose components are \[ \{f_i^{\prime\pr}(\bar x,d)\mid i\in I(\bar x,d)\}\}\quad\textrm{and}\quad \{g_j^{\prime\pr}(\bar x,d)\mid j\in J(\bar x,d)\}\}. \] With these notations Theorem \ref{th1} claims that the linear system $A^Tz+C^Tu<0$, $u\geqq 0$, where $u\in\mathbb R$, has no solutions. It follows from Lemma \ref{th-alternative} that there exist vectors $\lambda\in\mathbb R^n$ and $\mu\in\mathbb R^m$ with $\lambda\geqq 0$, $\mu\geqq 0$, $(\lambda,\mu)\ne (0,0)$ such that conditions (\ref{6}) and (\ref{10}) are satisfied. According to the assumptions $d$ is critical and $\nabla L(\bar x)=0$ we obtain that conditions (\ref{4}) are satisfied. \end{proof} \section{The Saddle Points of the Lagrange Function and Weak Efficiency} \label{s6}\setcounter{equation}{0} Consider the vector problem (P) and the scalar Lagrange function \[ L_\lambda(x,\mu)=\scalpr{\lambda}{f(x)}+\scalpr{\mu}{g(x)},\quad\lambda\in\Lambda,\;\mu\geqq 0. \] \begin{definition} A point $(\bar x,\bar\mu)\in X\times [0,+\infty)^m$ is called a Kuhn-Tucker saddle point of the Lagrange function if \begin{equation}\label{9} L_{\lambda}(\bar x,\mu)\leqq L_{\lambda}(\bar x,\bar\mu)\leqq L_{\lambda}(x,\bar\mu),\quad\forall x\in X,\;\forall\mu\geqq 0. \end{equation} \end{definition} Denote the vector $(\nabla f_1,\nabla f_2,\ldots,\nabla f_n)$ by $\nabla f$. The following definition is an extension to vector problems of the respective notion by Ivanov \cite{OptLett-2} which holds for scalar minimization problems. \begin{definition} Let {\rm (P)} be a Fr\'echet differentiable problem. We call {\rm (P)} Kuhn-Tucker saddle point invex (for short, KTSP-invex) if there exists a vector function $\eta: X\times X\to\mathbb R^s$ such that the following implication holds: \medskip \noindent $\left. \begin{array}{l} x\in X,\; \\ u\in X,\; \\ g(u)\leqq 0 \end{array} \right] \quad$ imply $\quad \left[ \begin{array}{l} \textrm{either } f(x)-f(u)\geqq\nabla f(u)\eta(x,u), \\ g_{j}(x)\geqq\nabla g_{j}(u)\eta(x,u),\; j\in A(u) \\ \textrm{or } \nabla f(u)\eta(x,u)<0, \\ \nabla g(u)\eta(x,u)\leqq 0. \end{array} \right.$ \hfill {\rm (KTSPI)} \end{definition} \medskip \begin{theorem}\label{th3} Let the problem {\rm (P)} be Fr\'echet differentiable. Then {\rm (P)} is KTSP-invex if and only if, for every triple $(\bar x,\bar\lambda,\bar\mu)$ which satisfies Kuhn-Tucker optimality conditions \begin{equation}\label{12} \scalpr{\bar\lambda}{\nabla f(\bar x)}+\sum_{j\in A(\bar x)}\bar\mu_j\nabla g_j(\bar x)=0,\quad\bar x\in S,\; \bar\lambda\ge 0,\; \bar\mu\geqq 0, \end{equation} the pair $(\bar x,\bar\mu)$ is a saddle point of the Lagrange function $L_{\bar\lambda}$. \end{theorem} \begin{proof} Suppose that (P) is KTSP-invex. Let the triple $(\bar x,\bar\lambda,\bar\mu)$ satisfy Kuhn-Tucker conditions. We prove that $(\bar x,\bar\mu)$ is a saddle point of Lagrange function. Suppose that $x$ is an arbitrary point from $X$. The left inequality in (\ref{9}) follows from $\bar x\in S$, $\bar\mu\geqq 0$, $\mu\geqq 0$, and $\bar\mu_j g_j(\bar x)=0$, $j=1,2,...,m$. Therefore, it is enough to prove that \begin{equation}\label{11} \scalpr{\bar\lambda}{f(x)-f(\bar x)}+\scalpr{\bar\mu}{g(x)}\geqq 0. \end{equation} It follows from Lemma \ref{th-alternative}, by (\ref{12}), that the system \begin{equation}\label{18} \nabla f(\bar x)\eta<0,\quad \nabla g(\bar x)\eta\leqq 0 \end{equation} is inconsistent. According to KTSP-invexity there exists $\eta\in\mathbb R^s$ such that \begin{equation}\label{13} f_i(x)-f_i(\bar x)\geqq \nabla f_i(\bar x)\eta,\quad i=1,2,\ldots, n \end{equation} \begin{equation}\label{14} g_j(x)\geqq \nabla g_j(\bar x)\eta,\quad j\in A(\bar x). \end{equation} Let us multiply (\ref{13}) by $\bar\lambda_i$ and (\ref{14}) by $\bar\mu_i$ and add all obtained inequalities. Then, we can see that (\ref{11}) is satisfied by Kuhn-Tucker conditions (\ref{12}), which implies that $(\bar x,\bar\mu)$ is a saddle point of $L_{\bar\lambda}$. Suppose that every pair $(\bar x,\bar\mu)$, which satisfies conditions (\ref{12}) together with the Lagrange multiplier $\bar\lambda$, is a saddle point of the Lagrange function $L_{\bar\lambda}$. We prove that (P) is KTSP-invex. Choose an arbitrary point $x\in X$. Consider the linear programming problem with variables $\lambda_1, \lambda_2,\ldots,\lambda_n$, $\mu_j$, $j\in A(\bar x)$ \medskip \noindent \begin{tabular}{@{}ll} Minimize & $\sum_{i=1}^n\lambda_i\left[f_i(x)-f_i(\bar x)\right]+$ $\sum_{j\in A(\bar x)} \mu_j\,g_j(x)$ \\ \smallskip\noindent subject to & $\sum_{i=1}^n\lambda_i \nabla f_i(\bar x)+\sum_{j\in A(\bar x)} \mu_j\nabla g_j(\bar x)=0$, \\ & $ \sum_{i=1}^n\lambda_i=1$, \\ & $\lambda_i\geqq 0$, $i=1,2,...,m$,\quad $\mu_j\geqq 0$, $j\in A(\bar x)$. \end{tabular} \medskip It follows from the hypothesis that the objective function of this problem is non-negative over the feasible set or the feasible set is empty. Therefore this problem is solvable with a non-negative minimal value or it is unsolvable when its feasible set is empty. The dual problem is the following one: \medskip \noindent \begin{tabular}{@{}lll} Maximize & $v$ & \\ subject to & $\nabla f(\bar x)\eta+v$ & $\leqq f(x)-f(\bar x)$ \\ & $\nabla g_j(\bar x)\eta$ & $\leqq g_j(x)$, $j\in A(\bar x)$. \\ \end{tabular} \medskip Consider the case when the primal problem is solvable. According to Duality Theorem in Linear Programming the dual problem is solvable, too, with the same optimal value. Therefore, there exist $v\ge 0$ and $\eta\in\mathbb R^s$, which satisfy the constraints of the dual problem. Consider the case when the primal problem is infeasible. By Lemma \ref{th-alternative} the infeasibility of the primal problem implies that the system (\ref{18}) for $\eta$ is consistent. Both cases bring that (P) is KTSP-invex. \end{proof} We introduce the following definition: \begin{definition} Let {\rm (P)} be a Fr\'echet differentiable problem. We call {\rm (P)} second-order Kuhn-Tucker saddle point invex (for short, second-order KTSP-invex) if for every $x\in X$, $u\in S$, and for every critical direction $d$ at $u$ such that there exist the derivatives $f^{\prime\pr}(u,d)$, $g^{\prime\pr}(u,d)$ with finite values at least one of the systems for $\eta$ and $\omega$ \medskip $\left\{ \begin{array}{l} \nabla f_i(u)\eta+\omega f^{\prime\pr}_i(u,d)\leqq f_i(x)-f_i(u),\; i=1,2,\ldots,n \\ \nabla g_j(u)\eta+\omega g^{\prime\pr}_j(u,d) \leqq g_j(x),\; j\in A(u), \\ \omega\geqq 0 \end{array}\right.$ \medskip \noindent and \medskip $\left\{ \begin{array}{l} \nabla f(u)\eta+\omega f^{\prime\pr}(u,d)<0, \\ \nabla g(u)\eta +\omega g^{\prime\pr}(u,d)\leqq 0, \\ \omega\geqq 0 \end{array}\right.$ \medskip \noindent has a solution. \end{definition} \begin{theorem}\label{th7} Every Fr\'echet differentiable KTSP-invex problem with inequality constraints is second-order KTSP-invex. \end{theorem} \begin{proof} Let (P) be KTSP-invex. Therefore, for all points $x\in X$ and $u\in X$ with $g(u)\leqq 0$ there exists $v\in\mathbb R^s$ such that either \[ \nabla f(u)v\leqq f(x)-f(u),\quad\nabla g_{j}(u)v\leqq g_{j}(x),\; j\in A(u), \] or \[ \nabla f(u)v<0,\quad\nabla g(u)v\leqq 0. \] The choice $\eta=v$, $\omega=0$, and $d=0$ shows that (P) is second-order KTSP-invex, because $f^{\prime\pr}(u,0)=0$ and $g_j^{\prime\pr}(u,0)=0$. \end{proof} \begin{theorem}\label{th4} Let the problem {\rm (P)} be Fr\'echet differentiable. Then {\rm (P)} is second-order KTSP-invex if and only if for every triple $(\bar x,\bar\lambda,\bar\mu)$, which satisfies the second-order Kuhn-Tucker necessary optimality conditions {\rm (\ref{6})} and \rm{(\ref{10})}, the pair $(\bar x,\bar\mu)$ is a saddle point of the Lagrange function $L_{\bar\lambda}$. \end{theorem} \begin{proof} Suppose that (P) is second-order KTSP-invex. Let the triple $(\bar x,\bar\lambda,\bar\mu)$ satisfy the Kuhn-Tucker conditions with respect to an arbitrary direction $d$ which is critical at $\bar x$. We prove that $(\bar x,\bar\mu)$ is a saddle point of the Lagrange function. Suppose that $x$ is an arbitrary point from $X$. The left inequality in (\ref{9}) follows from $\bar x\in S$, $\bar\mu\geqq 0$, $\mu\geqq 0$, and $\bar\mu_j g_j(\bar x)=0$, $j=1,2,...,m$. Therefore, it is enough to prove that inequality (\ref{11}) holds It follows from Lemma \ref{th-alternative}, by (\ref{6}) and (\ref{10}) that the system \[ \nabla f(\bar x)\eta+\omega f^{\prime\pr}(\bar x,d)<0,\quad \nabla g(\bar x)\eta +\omega g^{\prime\pr}(\bar x,d)\leqq 0,\quad\omega\ge0 \] is inconsistent. According to KTSP-invexity there exists $\eta\in\mathbb R^s$ and $\omega\ge 0$ such that \begin{equation}\label{16} f_i(x)-f_i(\bar x)\geqq \nabla f_i(\bar x)\eta+\omega f^{\prime\pr}_i(\bar x,d),\quad i=1,2,\ldots, n \end{equation} \begin{equation}\label{17} g_j(x)\geqq \nabla g_j(\bar x)\eta +\omega g^{\prime\pr}_j(\bar x,d),\quad j\in A(\bar x). \end{equation} Let us multiply (\ref{16}) by $\bar\lambda_i$ and (\ref{17}) by $\bar\mu_i$ and add all obtained inequalities. Then, we can see that (\ref{11}) is satisfied by Kuhn-Tucker conditions (\ref{6}) and (\ref{10}), which implies that $(\bar x,\bar\mu)$ is a saddle point of $L_{\bar\lambda}$. Suppose that $d$ is an arbitrary direction, and every pair $(\bar x,\bar\mu)$, which satisfies conditions (\ref{6}) and (\ref{10}) together with the Lagrange multiplier $\bar\lambda$, is a saddle point of the Lagrange function $L_{\bar\lambda}$. We prove that (P) is second-order KTSP-invex. Choose an arbitrary point $x\in X$. Consider the linear programming problem with variables $\lambda_1, \lambda_2,\ldots,\lambda_n$, $\mu_j$, $j\in A(\bar x)$ \medskip \noindent \begin{tabular}{@{}ll} Minimize & $\sum_{i=1}^n\lambda_i\left(f_i(x)-f_i(\bar x)\right)+$ $\sum_{j\in A(\bar x)} \mu_j\,g_j(x)$ \\ subject to & $\sum_{i=1}^n\lambda_i \nabla f_i(\bar x)+\sum_{j\in A(\bar x)} \mu_j\nabla g_j(\bar x)=0$, \\ & $\sum_{i=1}^n\lambda_i f_i^{\prime\pr}(\bar x,d)+ \sum_{j\in A(\bar x)}\mu_j g_j^{\prime\pr}(\bar x,d)\geqq 0$, \\ & $ \sum_{i=1}^n\lambda_i=1$, \\ & $\lambda_i\geqq 0$, $i=1,2,...,m$,\quad $\mu_j\geqq 0$, $j\in A(\bar x)$. \end{tabular} \medskip It follows from the hypothesis that the objective function of this problem is non-negative over the feasible set or the feasible set is empty. Therefore this problem is solvable with a non-negative minimal value or it is unsolvable when its feasible set is empty. The dual problem is the following one: \medskip \noindent \begin{tabular}{@{}lll} Maximize & $v$ & \\ subject to & $\nabla f(\bar x)\eta+\omega f^{\prime\pr}(\bar x,d)+v$ & $\leqq f(x)-f(\bar x)$, \\ & $\nabla g_j(\bar x)\eta +\omega g^{\prime\pr}_j(\bar x,d)$ & $\leqq g_j(x)$, $j\in A(\bar x)$, \\ & $ \omega\geqq 0$. \end{tabular} \medskip Consider the case when the primal problem is solvable. According to Duality Theorem in Linear Programming the dual problem is solvable, too, with the same optimal value. Therefore, there exist $v\ge 0$, $\eta\in\mathbb R^s$ and $\omega\in[0,+\infty)$, which satisfy the constraints of the dual problem. Consider the case when the primal problem is infeasible. By Lemma \ref{th-alternative} the infeasibility of the primal problem implies that the system \[ \nabla f(\bar x)\eta +\omega f^{\prime\pr}(\bar x,d)<0,\quad \nabla g(\bar x)\eta +\omega g^{\prime\pr}(\bar x,d)\leqq 0,\quad \omega\geqq 0. \] for $\eta$ and $\omega$ is consistent. Both cases bring that (P) is second-order KTSP-invex. \end{proof} The following example shows that the converse claim in Theorem \ref{th7} is not satisfied. \begin{example}\label{ex3} Consider the problem (P) where $X\equiv\mathbb R^2$, $f=(f_1,f_2)$ is the vector function of two variables such that \[f_1(x_1,x_2)=(x_1^2+x_2^2)^2-2x_1^2+2x_2^2,\quad f_2(x_1,x_2)=(x_1^2-1)^2+2x_2^2.\] and $g$ is the scalar function \[ g(x_1,x_2)=x_1^2+x_2^2-1. \] The KT stationary points are the pairs $(0,0)$, $(1,0)$, $(-1,0)$ with Lagrange multipliers $\bar\mu=0$, $\lambda\in\Lambda$. The pair $(\bar x,\bar\mu)$ with $\bar x=(0,0)$ and $\bar\mu=0$ is not a saddle point of $L_\lambda$ for every $\lambda=(\lambda_1,\lambda_2)\in\Lambda$, because \[ L_\lambda(x_{\varepsilon},\bar\mu)=\lambda_1(\varepsilon^4-2\varepsilon^2)+\lambda_2(\varepsilon^2-1)^2<L_\lambda(\bar x,\bar\mu),\; x_{\varepsilon}=(\varepsilon,0) \] for all $\varepsilon\in(0,\sqrt{2})$. It follows from Theorem \ref{th3} that this problem is not KTSP-invex. The point $(0,0)$ is not second-order stationary, because every direction is critical at $(0,0)$, and \[ \scalpr{d}{\nabla^2 L_\lambda(\bar x,\bar\mu)d}=-4d_1^2+4d_2^2<0,\;\bar x=(0,0),\;\bar \mu=0 \] if $d_2^2<d_1^2$, where $d=(d_1,d_2)$ is a direction. The points $(1,0)$ and $(-1,0)$ are second-order stationary, because the multipliers $\lambda_1=1$, $\lambda_2=0$, $\mu=0$, where $L=\lambda_1 f_1+\lambda_2 f_2+\mu g$ is the Lagrange function, satisfy the second-order necessary dual optimality conditions. Since these points compose with the multiplier $\mu=0$ saddle points of $L_\lambda$, then by Theorem \ref{th4}, the problem is second-order KTSP-invex. \end{example} \section{Optimality Conditions with Second-Order KT-Pseudoinvex Problems} \label{s4}\setcounter{equation}{0} Consider the problem (P). We denote the vector \[ (f^{\prime\pr}_1(x,d), f^{\prime\pr}_2(x,d),\dots, f^{\prime\pr}_n(x,d)) \] by $f^{\prime\pr}(x,d)$, and the set of all critical directions at $x$ by $D(x)$. The following notion was introduced by Osuna-Gomez, Beato-Moreno, Rufian-Lizana \cite{osu99}: \begin{definition}\label{df2} Let the problem {\rm (P)} be Fr\'echet differentiable. Then it is called KT-pseudoinvex-I if there exists a map $\eta: X\times X\to\mathbb R^s$ such that the following implication holds: \[ \left. \begin{array}{l} x\in X,\; y\in X,\; f(y)<f(x) \\ g(x)\leqq 0,\; g(y)\leqq 0 \\ \end{array}\right] \quad\Rightarrow \quad \left[ \begin{array}{l} \nabla f(x)\eta(x,y)<0 \\ \nabla g_{j}(x)\eta(x,y)\leqq 0, j\in A(x). \end{array} \right. \] \end{definition} Definition \ref{df2} is a generalization of the notion of KT-invex problems due to Martin \cite{mar85} for scalar problems, because for such problems KT-invexity is equivalent to KT-pseudoinvexity. The following proposition is the main result in the paper of Osuma-Gomez, Rufian-Lizana, Ruiz-Canales \cite{osu98} (see Theorem 2.3). It is a generalization of Theorem 2.1 due to Martin \cite{mar85}. \begin{proposition}\label{CharKT-I} Let the problem (P) be Fr\'echet differentiable. Then it is KT-pseudoinvex-I if and only if every KT stationary point is a weak global minimizer. \end{proposition} Here we study the second-order case. Second-order KT-invex scalar problems are defined by Ivanov. \cite{JOGO} We introduce the following notion, which is a second-order analog of Definition \ref{df2}. \begin{definition} Let {\rm (P)} be a Fr\'echet differentiable problem. Then we call {\rm (P)} second-order KT-pseudoinvex-I if there exist maps $d: X\times X\to\mathbb R^s$, $\eta: X\times X\to\mathbb R^s$ and a function $\omega: X\times X\to [0,+\infty)$ such that the following implication holds: \[ \left. \begin{array}{l} x\in X,\; y\in X \\ g(x)\leqq 0,\\ g(y)\leqq 0, \\ f(y)<f(x) \end{array}\right] \Rightarrow \left[ \begin{array}{l} \exists f^{\prime\pr}(x,d(x,y)),\;\exists g^{\prime\pr}_j(x,d(x,y)),\; j\in A(x) \\ \nabla f(x)\eta(x,y)+ \omega(x,y) f^{\prime\pr}(x,d(x,y))<0 \\ \nabla g_{j}(x)\eta(x,y)+\omega(x,y) g^{\prime\pr}_{j}(x,d(x,y))\leqq 0,\; j\in A(x),\\ \textrm{where}\quad d(x,y)\in D(x). \end{array} \right. \] We suppose here that there exist $f^{\prime\pr}_i(x,d(x,y))$ and $g^{\prime\pr}_j(x,d(x,y))$ with finite values. \end{definition} \begin{theorem}\label{KT-I} Every Fr\'echet differentiable KT-pseudoinvex-I problem with inequality constraints is second-order KT-pseudoinvex-I. \end{theorem} \begin{proof} Let (P) be KT-pseudoinvex-I. Therefore, for all feasible points $x$ and $y$ with $f(y)<f(x)$ there exists $u\in\mathbb R^s$ such that \[ \nabla f(x)u<0,\quad\nabla g_{j}(x)u\leqq 0,\; j\in A(x). \] The choice $\eta=u$, $\omega=0$, and $d=0$ shows that (P) is second-order KT-pseudoinvex-I, because $f^{\prime\pr}(x,0)=0$ and $g_j^{\prime\pr}(x,0)=0$. \end{proof} \begin{definition} A point $x\in S$ is called second-order Kuhn-Tucker stationary point (for short, second-order KT point) if for every critical direction $d\in\mathbb R^s$ at $x$ such that there exist the second-order derivatives $f^{\prime\pr} (x,d)$, $g_j^{\prime\pr} (x,d)$, $j\in A(x)$ with finite values, there are Lagrange multipliers $ \lambda= (\lambda_1, \lambda_2,\ldots, \lambda_n)\ge 0$, $\mu=(\mu_1,\mu_2,\ldots,\mu_m)\geqq 0$, satisfying equations (\ref{6}) and (\ref{10}). \end{definition} Obviously, every second-order KT stationary point $x$ is a first-order KT stationary point. Indeed, for every Fr\'echet differentiable problem there is at least one critical direction $d\in\mathbb R^n$ such that there exist the second-order derivatives $f^{\prime\pr} (x,d)$, $g_j^{\prime\pr} (x,d)$, $j\in A(x)$. It is the direction $d=0$. Therefore, conditions (\ref{6}) and (\ref{10}) are satisfied when $d=0$. It follows from (\ref{6}) that $x$ is a KT stationary point. The converse implication does not hold. \begin{theorem}\label{th2} Let the problem {\rm (P)} be Fr\'echet differentiable. Then every second-order KT stationary point is a weak global Pareto minimizer if and only if {\rm (P)} is second-order KT-pseudoinvex-I. \end{theorem} \begin{proof} Let (P) be second-order KT-pseudoinvex-I. We prove that every se\-cond-order KT point is a weak global Pareto minimizer. Suppose the contrary that $x$ is a second-order KT stationary point, but there exists $y\in S$ with $f(y)<f(x)$. It follows from second-order pseudoinvexity that there exist a critical direction $d\in\mathbb R^s$, $\eta\in\mathbb R^s$, and a number $\omega\ge 0$ such that \begin{equation}\label{21} \nabla f_i(x)\eta(x,y)+ \omega(x,y) f^{\prime\pr}_i(x,d(x,y))<0,\; i=1,2,\ldots, n, \end{equation} \begin{equation}\label{22} \nabla g_j(x)\eta(x,y)+ \omega(x,y) g^{\prime\pr}_j(x,d(x,y))\leqq 0,\; \forall j\in A(x). \end{equation} We conclude from $x$ is a second-order KT point and $d$ is critical that there exist $\lambda\ge 0$ and $\mu\geqq 0$ with \begin{equation}\label{23} \sum_{i=1}^n\lambda_i\nabla f_i(x)+\sum_{j\in A(x)}\mu_j\nabla g_j(x)=0 \end{equation} \begin{equation}\label{24} \sum_{i=1}^n\lambda_i f^{\prime\pr}_i(x,d)+\sum_{j\in A(x)}\mu_j g^{\prime\pr}_j(x,d)\geqq 0. \end{equation} Let us multiply (\ref{21}) by $\lambda_i$, (\ref{22}) by $\mu_j$ and add all obtained inequalities. Taking into account (\ref{23}), (\ref{24}) and $\lambda\ne 0$ we get the impossible inequality $0<0$. Let each second-order KT point be a weak global Pareto minimizer. We prove that (P) is second-order KT-pseudoinvex-I. Take $x\in X$, $y\in X$ with $g(x)\leqq 0$, $g(y)\leqq 0$, $f(y)<f(x)$. Therefore, $x$ is not a weak global Pareto minimizer. Hence, $x$ is not a second-order KT point. Consequently, there is a critical direction $d$ such that the derivatives $f^{\prime\pr}(x,d)$, $g^{\prime\pr}(x,d)$ exist with finite values, but there are not $(\lambda,\mu)\geqq 0$, $\lambda\ne 0$ which satisfy (\ref{23}) and (\ref{24}). Therefore, the linear programming problem \medskip \begin{tabular}{ll} Maximize & $\sum_{i=1}^n\lambda_i$ \\ subject to & $\sum_{i=1}^n\lambda_i\nabla f_i(x)+\sum_{j\in A(x)}\mu_j\nabla g_j(x)=0$, \\ & $\sum_{i=1}^n\lambda_i f^{\prime\pr}_i(x,d)+\sum_{j\in A(x)}\mu_j g^{\prime\pr}_j(x,d)\ge 0,$ \\ & $\lambda_i\ge 0$,\quad $i=1,2,\ldots, n$,\quad $\mu_j\ge 0,\quad\forall j\in A(x)$ \end{tabular} \medskip \noindent has non-positive maximal value. If we take $\lambda_i=0$, $\mu_j=0$, $j\in A(x)$, then we obtain that the exact optimal value of this problem is 0. The dual linear problem is \medskip \begin{tabular}{ll} Minimize & 0 \\ subject to & $\nabla f_i(x)u-v f^{\prime\pr}_i(x,d)\geqq 1$, $i=1,2,\ldots,n$ \\ & $\nabla g_j(x)u-v g^{\prime\pr}_j(x,d)\geqq 0,\; j\in A(x)$ \\ & $v\geqq 0$. \end{tabular} \medskip \noindent It follows from Duality Theorem that the dual problem is also solvable. Therefore, there exist $\eta\in\mathbb R^s$ and $\omega\geqq 0$ ($\eta=-u$, $\omega=v$), which satisfy inequalities (\ref{21}) and (\ref{22}). Hence, (P) is second-order KT-pseudoinvex-I. \end{proof} The following examples show that the converse claim in Theorem \ref{KT-I} is not satisfied. \begin{example}\label{ex2} Consider the vector function of two variables $f=(f_1,f_2)$ such that \[f_1(x_1,x_2)=(x_1^2+x_2^2)^2-2x_1^2+2x_2^2,\quad f_2(x_1,x_2)=(x_1^2-1)^2+2x_2^2.\] The KT stationary points are the pairs $(0,0)$, $(1,0)$, $(-1,0)$. The vector function is not KT-pseudoinvex-I, because $(0,0)$ is not weakly efficient. For instance, we have $f(\varepsilon,0)<f(0,0)$ for all $\varepsilon\in(0,\sqrt{2})$. On the other hand $(0,0)$ is not a second-order stationary point, because $\scalpr{d}{\nabla^2 L(0,0)\, d}<0$ for all critical directions $d=(d_1,d_2)$ such that $|d_1|>|d_2|$ where $L=\lambda_1 f_1+\lambda_2 f_2$ is the Lagrange function with a multiplier $\lambda=(\lambda_1,\lambda_2)$. The other stationary points $(1,0)$ and $(-1,0)$ are second-order stationary ones and they are global minimizers of the component $f_2$ and weakly efficient for $f$. Therefore, by Theorem \ref{th2} the unconstrained problem is second-order KT-pseudoinvex-I. It seem that the unconstrained minimization of $f$ could be reduced to the problem \medskip \begin{tabular}{ll} Minimize & $f(x,y)$ \\ subject to & $g_1(x,y)=-x\leqq 0,\; g_2(x,y)=-y\leqq 0$ \end{tabular}\hfill {\rm (P$^\prime$)} \medskip \noindent by the substitution $x=x_1^2$, $y=x_2^2$, where \[ f=(f_1,f_2),\quad f_1(x,y)=(x+y)^2-2x+2y,\quad f_2(x,y)=(x-1)^2+2y. \] The only stationary point of {\rm (P$^\prime$)} is $(x,y)=(1,0)$ and it is weakly efficient. Therefore, {\rm (P$^\prime$)} is KT-pseudoinvex-I in contrast of the unconstrained minimization of $f(x_1,x_2)$. \end{example} \begin{example}\label{ex4} Consider the problem (P) where $X\equiv\mathbb R^2$, $f=(f_1,f_2)$ is the vector function of two variables such that \[f_1(x_1,x_2)=2x_1x_2-2x_1^2-x_2^2+8x_1-6x_2,\quad f_2(x_1,x_2)=-x_1+x_2,\] and $g$ is the scalar function of two variables \[ g(x_1,x_2)=x_1-x_1^2+x_2. \] The set of the KT stationary points is the segment whose endpoints are (1,0) and (1,-2): \[ \{(x_1,x_2)\mid x_1=1,-2\le x_2\le 0\}. \] All points from the segment are not weakly efficient, because both functions $f_1$ and $f_2$ are strictly monotone over the ray \[ \{(x_1,x_2)\mid x_1=1, x_2\le -2\}, \] they approach $-\infty$ when $x_2$ tends to $-\infty$, and all points $(x_1,x_2)$ such that $x_1=1$, $x_2\leqq 0$ are feasible. Therefore, by Proposition \ref{CharKT-I}, the problem is not KT-pseudoinvex-I. All KT stationary points do not satisfy the second-order necessary optimality conditions in Theorem \ref{NKT}. Therefore, the problem has no second-order KT stationary points. By Theorem \ref{th2} the problem is second-order KT-pseudoinvex-I. \end{example} The next definition was introduced by Arana-Jimenez, Rufian-Lizana, Osuna-Gomez and Ruiz-Garzon \cite{ara08}: \begin{definition}\label{df5} Let the problem {\rm (P)} be Fr\'echet differentiable. Then it is called KT-pseudoinvex-II if there exists a map $\eta: X\times X\to\mathbb R^s$ such that the following implication holds: \[ \left. \begin{array}{l} x\in X,\; y\in X,\; f(y)\le f(x) \\ g(x)\leqq 0,\; g(y)\leqq 0 \\ \end{array}\right] \quad\Rightarrow \quad \left[ \begin{array}{l} \nabla f(x)\eta(x,y)<0 \\ \nabla g_{j}(x)\eta(x,y)\leqq 0,\; j\in A(x). \end{array} \right. \] \end{definition} Definition \ref{df5} is an another generalization of KT-invexity (see Martin \cite{mar85}) to vector problems. We introduce the following notion which is a second-order analog of the last definition. \begin{definition} Let the problem {\rm (P)} be Fr\'echet differentiable. Then we call {\rm (P)} second-order KT-pseudoinvex-II if there exist maps $d: X\times X\to\mathbb R^s$, $\eta: X\times X\to\mathbb R^s$ and a function $\omega: X\times X\to [0,+\infty)$ such that the following implication holds: \[ \left. \begin{array}{l} x\in X,\; y\in X \\ g(x)\leqq 0,\\ g(y)\leqq 0, \\ f(y)\le f(x) \end{array}\right] \;\Rightarrow \; \left[ \begin{array}{l} \exists f^{\prime\pr}(x,d(x,y)),\;\exists g^{\prime\pr}_{j}(x,d(x,y)),\; j\in A(x) \\ \nabla f(x)\eta(x,y)+ \omega(x,y) f^{\prime\pr}(x,d(x,y))<0 \\ \nabla g_{j}(x)\eta(x,y)+ \omega(x,y) g^{\prime\pr}_{j}(x,d(x,y))\leqq 0, \; j\in A(x) \\ \textrm{where}\quad d(x,y)\in D(x). \end{array} \right. \] We suppose here that $f^{\prime\pr}_i(x,d(x,y))$ and $g^{\prime\pr}_j(x,d(x,y))$ are finite. \end{definition} The proofs of the following results are similar to the proofs of Theorems \ref{KT-I} and \ref{th2}. \begin{proposition} Every Fr\'echet differentiable KT-pseudoinvex-II problem with inequality constraints is second-order KT-pseudoinvex-II. \end{proposition} \begin{proposition} Let {\rm (P)} be a Fr\'echet differentiable problem. Then each second-order KT stationary point is a global Pareto minimizer if and only if {\rm (P)} is second-order KT-pseudoinvex-II. \end{proposition} \section{The Weighting Problem and Second-Order Kuhn-Tucker Stationary Points} \label{s5}\setcounter{equation}{0} For every vector $\lambda\in\Lambda$ consider the weighting scalar problem P$_\lambda$. The following notion was introduced by Osuna-Gomez, Beato-Moreno, Rufian-Lizana \cite{osu99}: \begin{definition}\label{df7} Let the problem {\rm (P)} be Fr\'echet differentiable. Then it is called KT-invex if there exists a map $\eta: X\times X\to\mathbb R^s$ such that the following implication holds: \[ \left. \begin{array}{l} x\in X,\; y\in X,\; \\ g(x)\leqq 0,\; g(y)\leqq 0 \\ \end{array}\right] \quad\Rightarrow \quad \left[ \begin{array}{l} f(y)-f(x)\geqq\nabla f(x)\eta(x,y) \\ 0\geqq\nabla g_{j}(x)\eta(x,y),\; j\in A(x). \end{array} \right. \] \end{definition} Definition \ref{df7} is a generalization of the notion of KT-invex problems due to Martin \cite{mar85} for scalar problems. It is shown by Osuna-Gomez, Beato-Moreno, Rufian-Lizana \cite{osu99} that a Fr\'echet differentiable problem (P) is KT-invex if and only if every KT stationary point is a solution of the weighting problem. We introduce the following notion, which is a generalization of the notion of second-order KT-invex problems with inequality constraints for scalar problems (see Ivanov \cite{JOGO,Optimization}). \begin{definition}\label{df1} Let the problem {\rm (P)} be Fr\'echet differentiable. Then we call {\rm (P)} second-order KT-invex if there are maps $d: X\times X\to\mathbb R^s$, $\eta: X\times X\to\mathbb R^s$ and a function $\omega: X\times X\to [0,+\infty)$ such that $f^{\prime\pr}(x,d(x,y))$ and $g^{\prime\pr}_{j}(x,d(x,y))$, $j\in A(x)$ exist as finite values for all $x\in S$, $y\in S$ and the following inequalities hold: \begin{equation}\label{25} f_i(y)-f_i(x)\geqq\nabla f_i(x)\eta(x,y)+\omega(x,y) f^{\prime\pr}_i(x,d(x,y)),\; i=1,2,\ldots,n, \end{equation} \begin{equation}\label{26} 0\geqq\nabla g_j(x)\eta(x,y)+\omega(x,y) g^{\prime\pr}_j(x,d(x,y)),\; j\in A(x). \end{equation} We suppose here that the direction $d(x,y)$ is critical at $x$. \end{definition} In this definition the second-order directional derivatives of $f$ and $g$ are not required to exist for every direction. \begin{theorem}\label{th5} Every Fr\'echet differentiable KT-invex vector problem with inequality constraints is second-order KT-invex. \end{theorem} \begin{proof} Let (P) be KT-invex. Therefore, for all $x\in S$, $y\in S$ there exists $u\in\mathbb R^n$ such that \[ f(y)-f(x)\geqq \nabla f(x)u,\quad 0\geqq\nabla g_{j}(x)u,\; j\in A(x). \] The choice $\eta=u$, $\omega=0$, and $d=0$ shows that (P) is second-order KT invex, because $f^{\prime\pr}(x,0)=0$ and $g_{j}^{\prime\pr}(x,0)=0$, $j\in A(x)$. \end{proof} \begin{theorem}\label{th6} Let {\rm (P)} be Fr\'echet differentiable. Then each second-order KT stationary point $x$ is a global solution of the weighting problem ${\rm (P_\lambda)}$ with the same vector $\lambda$ if and only if {\rm (P)} is second-order KT-invex. \end{theorem} \begin{proof} Let (P) be second-order KT-invex. We prove that every second-order KT stationary point $x$ is a global solution of the respective weighting problem. Choose arbitrary feasible point $y\in S$. Suppose that $x$ is a second-order KT point. It follows from KT-invexity that there exists a direction $d\in\mathbb R^s$ which is critical at $x$, a map $\eta\in\mathbb R^s$, and a number $\omega\geqq 0$ such that the derivatives $f^{\prime\pr}(x,d)$, $g^{\prime\pr}_{j}(x,d)$, $j\in A(x)$ exist and equalities (\ref{25}) and (\ref{26}) are satisfied. We conclude from $x$ is a second-order KT point and $d$ is critical that there exist $\lambda\ge 0$ and $\mu\geqq 0$ satisfying (\ref{23}) and (\ref{24}). Let us multiply (\ref{25}) by $\lambda_i$, (\ref{26}) by $\mu_j$ and add all obtained inequalities. Taking into account (\ref{23}) and (\ref{24}) we conclude that $\scalpr{\lambda}{f(y)}\geqq\scalpr{\lambda}{f(x)}$. Let each second-order KT stationary point be a solution of the weighting problem. We prove that (P) is second-order KT invex. Suppose the contrary that there exist $x\in S$, $y\in S$ such that for every critical direction $d$ with the property that, if the derivatives $f^{\prime\pr}(x,d)$, $g^{\prime\pr}_{j}(x,d)$, $j\in A(x)$ exist with finite values, then the system \[ \left\{ \begin{array}{l} \nabla f(x)\eta+\omega f^{\prime\pr}(x,d)\leqq f(y)-f(x) \\ \nabla g_{j)}(x)\eta+\omega g^{\prime\pr}_{j)}(x,d)\leqq 0,\; j\in A(x) \end{array}\right. \] has no solution for $\eta\in\mathbb R^s$ and $\omega\geqq 0$. Therefore, the linear programming problem \medskip \begin{tabular}{ll} Maximize & 0 \\ subject to &$\nabla f_i(x)\eta +\omega f^{\prime\pr}_i(x,d)\leqq f_i(y)-f_i(x)$, $i=1,2,\ldots,n$, \\ & $\nabla g_j(x)\eta +\omega g^{\prime\pr}_j(x,d)\leqq 0,\; j\in A(x)$, \\ & $\omega\geqq 0$. \end{tabular} \medskip \noindent is infeasible. Its dual linear programming problem is the following one \medskip \begin{tabular}{ll} Minimize & $\sum_{i=1}^n\lambda_i[f_i(y)-f_i(x)]$ \\ subject to & $\sum_{i=1}^n\lambda_i\nabla f_i(x)+\sum_{j\in A(x)}\mu_j\nabla g_j(x)=0$, \\ & $\sum_{i=1}^n\lambda_i f^{\prime\pr}_i(x,d)+\sum_{j\in A(x)}\mu_j g^{\prime\pr}_j(x,d)\ge 0,$ \\ & $\lambda_i\geqq 0,\; i=1,2,\dots, n,\quad \mu_j\geqq 0,\; j\in A(x)$. \end{tabular} \medskip It follows from Duality Theorem that both problems are simultaneously solvable or unsolvable. The dual problem is feasible ($\lambda=0$, $\mu=0$ is a feasible point). Therefore, the dual problem is not solvable, because it is unbounded. Hence, there exists a feasible point $(\lambda,\mu)$ for the dual problem such that $\lambda\ne 0$. It follows from $d$ is critical and $(\lambda,\mu)$ with $\lambda\ne 0$ is feasible for the dual problem that $x$ is a second-order Kuhn-Tucker point for the problem (P). According to our assumption $x$ is a solution of the weighting problem, that is $\scalpr{\lambda}{f(y)-f(x)}\geqq 0$. This conclusion is satisfied for for every $(\lambda,\mu)$ with $\lambda\ne 0$. Hence the optimal value of the dual problem is bounded from below by 0, which contradicts our indirect conclusion that it is unbounded. \end{proof} Example \ref{ex2} show that the converse claim in Theorem \ref{th5} is not satisfied. The stationary point $x=(0,0)$ is not weakly efficient. Therefore, it is not a global solution of the scalar problem $P_\lambda$ for every $\lambda\in\Lambda$, and by Theorem \ref{th6}, the unconstrained problem from this example is not KT-invex. The set of second-order stationary points consists of the points $(1,0)$ and $(-1,0)$. Both points are global minimizers of $f_1$ and $f_2$. Therefore, these points are global solutions of $P_\lambda$.
\section{Introduction} The Brouwer--Heyting--Kolmogorov interpretation for intuitionistic logic gives an informal relation between proofs and constructions. Since computations are a special kind of construction, it therefore seems reasonable to suspect that there is also a relation between constructive proofs and computations. There are several approaches to making such a connection in a mathematically rigorous way. Probably the best known of these is Kleene realisability \cite{kleene-1945}, which turns out to correspond to a proper extension of intuitionistic logic. Both Kleene realisability and variants of it have been well-studied, see e.g.\ van Oosten \cite{vanoosten-2008}. Medvedev \cite{medvedev-1955} followed an alternative path, in an attempt to formalise Kol\-mo\-go\-rov's calculus of problems. He introduced the \emph{Medvedev lattice} $\mathpzc{M}$, which is a lattice arising from computability-theoretic considerations. Furthermore, it is a Brouwer algebra and therefore provides a semantics for an intermediate propositional logic, i.e.\ a propositional logic lying between intuitionistic propositional logic (IPC) and classical logic. Unfortunately, this approach also turns out to capture a proper extension of IPC: namely, IPC plus the weak law of the excluded middle $\neg p \vee \neg\neg p$. The same holds for the closely related \emph{Muchnik lattice} $\mathpzc{M}_w$, which was introduced by Muchnik in \cite{muchnik-1963}. However, this does not mean it is impossible to capture IPC using the Medvedev lattice. For any Brouwer algebra $\mathscr{B}$ and any $x \in \mathscr{B}$, the factor $\mathscr{B} / \{y \in \mathscr{B} \mid y \geq x\}$ (which we will denote by $\mathscr{B} / x$) is also a Brouwer algebra. Thus one might ask if the next-best thing holds for the Medvedev lattice: is there an $\mathcal{A} \in \mathpzc{M}$ such that the theory of $\mathpzc{M} / \mathcal{A}$ is exactly $\mathrm{IPC}$? Quite impressively, Skvortsova \cite{skvortsova-1988-en} showed that there is such a principal factor of the Medvedev lattice which captures IPC. Unfortunately, the class $\mathcal{A}$ generating this factor is unnatural in the sense that it is constructed in an ad hoc manner. This leads to the natural question, posed in Terwijn \cite{terwijn-2006}: are there any natural principal factors of the Medvedev lattice which have IPC as their theory? For the Muchnik lattice one can ask a similar question. Sorbi and Terwijn \cite{sorbi-terwijn-2012} showed that there is also a principal factor of the Muchnik lattice with IPC as its theory, but it suffers from the same problem as Skvortsova's factor of the Medvedev lattice. In \cite{kuyper-2013-2}, the author has shown that there are natural principal factors of the Muchnik lattice which capture IPC. These factors are defined using common notions from computability theory, such as lowness, 1-genericity, hyperimmune-freeness and computable traceability. In this paper we present progress towards an affirmative answer to the question formulated above, by showing that there are principal factors of the Medvedev lattice capturing IPC which are more natural than the one given by Skvortsova. These factors arise from the computability-theoretic notion of a \emph{computably independent set}: that is, a set $A$ such that for every $i \in \omega$ we have that $\bigoplus_{j \neq i} A^{[j]} \not\geq_T A^{[i]}$, where $A^{[i]}$ is the $i^\textrm{th}$ column of $A$, i.e.\ $A^{[i]}(n) = A(\langle i,n \rangle)$ . We can now state the main theorem of this paper. \begin{restatable}{thm}{mainthm} \label{thm-main} Let $A$ be a computably independent set. Then \[\mathrm{Th}\left(\mathpzc{M} / \left\{i \conc f \mid f \geq_T A^{[i]}\right\}\right) = \mathrm{IPC}.\] \end{restatable} The existence of computably independent sets was first proven by Kleene and Post \cite{kleene-post-1954}. In fact, almost all sets are computably independent: both in the measure-theoretic sense, because every $1$-random is computably independent by van Lambalgen's theorem (see e.g.\ Downey and Hirschfeldt \cite[Theorem 6.9.1]{downey-hirschfeldt-2010}), and also in the Baire category sense, because every $1$-generic is computably independent by the genericity analogue of van Lambalgen's theorem (see e.g.\ \cite[Theorem 8.20.1]{downey-hirschfeldt-2010}). We note that the factor from Theorem \ref{thm-main} is not nearly as natural as the factors for the Muchnik lattice from \cite{kuyper-2013-2}, where for example it is shown that \[\mathrm{Th}\left(\mathpzc{M}_w / \left\{f \mid f \text{ is not low}\right\}\right) = \mathrm{IPC}.\] (Note that this factor does not work for the Medvedev lattice by \cite[p.\ 138]{skvortsova-1988-en}.) On the other hand, the factor from Theorem \ref{thm-main} is far more natural than the one given by Skvortsova: our factor is easily definable from just a computably independent set, which occurs naturally in computability theory. Furthermore, while Skvortsova used a deep result by Lachlan, we manage to work around this and therefore our proof is more elementary. We also study a question posed by Sorbi and Terwijn in \cite{sorbi-terwijn-2008}. As mentioned above, the theory of the Medvedev lattice is equal to Jankov's logic $\mathrm{Jan}$, the deductive closure of $\mathrm{IPC}$ plus the weak law of the excluded middle $\neg p \vee \neg\neg p$. Let $0'$ be the mass problem consisting of all non-computable functions. Recall that we say that a mass problem is \emph{Muchnik} if it is upwards closed under Turing reducibility. In \cite{sorbi-terwijn-2008} it is shown that for all Muchnik $\mathcal{B} >_\mathpzc{M} 0'$ the theory of the factor $\mathpzc{M} / \mathcal{B}$ is contained in $\mathrm{Jan}$. Therefore, Sorbi and Terwijn asked: is $\mathrm{Th}(\mathpzc{M} / \mathcal{B})$ contained in $\mathrm{Jan}$ for all mass problems $\mathcal{B} >_\mathpzc{M} 0'$? Sorbi and Terwijn also proposed a connected question: does every $\mathcal{B} >_\mathpzc{M} 0'$ bound a join-irreducible Medvedev degree $>_\mathpzc{M} 0'$? By their results, this would imply that $\mathrm{Th}(\mathpzc{M} / \mathcal{B})$ is always contained in $\mathrm{Jan}$. However, they conjectured the answer to this connected question to be negative, a fact which was later proven by Shafer \cite{shafer-2011}. Nonetheless, in the same paper, Shafer widened the class of mass problems $\mathcal{B}$ for which $\mathrm{Th}(\mathpzc{M} / \mathcal{B}) \subseteq \mathrm{Jan}$ holds to those $\mathcal{B}$ which bound a `pseudo-meet' of a countable sequence of join-irreducible degrees. Unfortunately, Shafer also showed that this still does not cover all $\mathcal{B} >_\mathpzc{M} 0'$. We give a positive answer to Sorbi and Terwijn's question. This is accomplished by showing that a relativisation of Theorem \ref{thm-main} holds, i.e.\ that for every $\mathcal{B} > 0'$ there is in fact a factor $\mathcal{C} \leq_\mathpzc{M} \mathcal{B}$ such that $\mathrm{Th}(\mathpzc{M} / \mathcal{C}) = \mathrm{IPC}$. Our notation is mostly standard. We denote the natural numbers by $\omega$, Cantor space by $2^\omega$ and Baire space by $\omega^\omega$. For any set $X \subseteq \omega^\omega$ we denote by $C(X)$ the upper cone $\{f \in \omega^\omega \mid \exists g \in X (f \geq_T g)\}$. By $\conc$ we denote concatenation of strings. For any set $\mathcal{A} \subseteq \omega^\omega$ we denote by $\overline{\mathcal{A}}$ its complement in $\omega^\omega$. For unexplained notions from computability theory, we refer to Odifreddi \cite{odifreddi-1989}, for the Muchnik and Medvedev lattices, we refer to the surveys of Sorbi \cite{sorbi-1996} and Hinman \cite{hinman-2012} (but we use the notation from Sorbi and Terwijn \cite{sorbi-terwijn-2008}), and finally for unexplained notions from lattice theory we refer to Balbes and Dwinger \cite{balbes-dwinger-1975}. \section{Preliminaries} First, let us recall the definition of the Medvedev lattice. \begin{defi}{\rm (Medvedev \cite{medvedev-1955})} Let $\mathcal{A},\mathcal{B} \subseteq \omega^\omega$ (we will call subsets of $\omega^\omega$ \emph{mass problems}). We say that $\mathcal{A}$ \emph{Medvedev reduces to} $\mathcal{B}$ (denoted by $\mathcal{A} \leq_\mathpzc{M} \mathcal{B}$) if there exists a Turing functional $\Phi$ such that $\Phi(\mathcal{B}) \subseteq \mathcal{A}$. If both $\mathcal{A} \leq_\mathpzc{M} \mathcal{B}$ and $\mathcal{B} \leq_\mathpzc{M} \mathcal{A}$ we say that $\mathcal{A}$ and $\mathcal{B}$ are \emph{Medvedev equivalent} (denoted by $\mathcal{A} \equiv_\mathpzc{M} \mathcal{B}$). The equivalence classes of mass problems under Medvedev equivalence are called \emph{Medvedev degrees}, and the class of all Medvedev degrees is denoted by $\mathpzc{M}$. \end{defi} Instead of the usual notation $\vee$ for joins (least upper bounds) and $\wedge$ for meets (greatest lower bounds) in lattices, we use $\oplus$ respectively $\otimes$. The reason for this is that we will shortly see that $\oplus$ corresponds to logical conjunction $\wedge$, while $\otimes$ corresponds to logical disjunction $\vee$. \begin{defi}{\rm(McKinsey and Tarski \cite{mckinsey-tarski-1946})} A \emph{Brouwer algebra} is a bounded distributive lattice together with a binary \emph{implication operator} $\to$ satisfying: \[a \oplus c \geq b \text{ if and only if } c \geq a \to b\] i.e.\ $a \to b$ is the least element $c$ satisfying $a \oplus c \geq b$. \end{defi} As the name suggests, the Medvedev lattice is a lattice. In fact, it is also a Brouwer algebra, as the next proposition shows. \begin{prop}{\rm (\cite{medvedev-1955})} The Medvedev lattice is a Brouwer algebra under the operations induced by: \begin{align*} \mathcal{A} \oplus \mathcal{B} &= \{f \oplus g \mid f \in \mathcal{A} \text{ and } g \in \mathcal{B}\}\\ \mathcal{A} \otimes \mathcal{B} &= \{0 \conc f \mid f \in \mathcal{A}\} \cup \{1 \conc g \mid g \in \mathcal{B}\}\\ \mathcal{A} \to \mathcal{B} &= \{n \conc f \mid \forall g \in \mathcal{A} (\Phi_n(f \oplus g) \in \mathcal{B}\}. \end{align*} Furthermore, the bottom element $0$ is the Medvedev degree of $\omega^\omega$, while the top element $1$ is the Medvedev degree of $\emptyset$. \end{prop} The main reason Brouwer algebras are interesting is because we can use them to give algebraic semantics for IPC, as witnessed by the next definition and the results following after it. \begin{defi}{\rm(\cite{mckinsey-tarski-1948})} Let $\phi(x_1,\dots,x_n)$ be a propositional formula with free variables among $x_1,\dots,x_n$, let $\mathscr{B}$ be a Brouwer algebra and let $b_1,\dots,b_n \in \mathscr{B}$. Let $\psi$ be the formula in the language of Brouwer algebras obtained from $\phi$ by replacing logical disjunction $\vee$ by $\otimes$, logical conjunction $\wedge$ by $\oplus$, logical implication $\to$ by Brouwer implication $\to$ and the false formula $\bot$ by $1$ (we view negation $\neg\alpha$ as $\alpha \to \bot$). We say that $\phi(b_1,\dots,b_n)$ \emph{holds in $\mathscr{B}$} if $\psi(b_1,\dots,b_n) = 0$. Furthermore, we define the \emph{theory} of $\mathscr{B}$ (notation: $\mathrm{Th}(\mathscr{B})$) to be the set of those formulas which hold for every valuation, i.e.\ \[\mathrm{Th}(\mathscr{B}) = \{\phi(x_1,\dots,x_m) \mid \forall b_1,\dots,b_m \in \mathscr{B}(\phi(b_1,\dots,b_m) \text{ holds in } \mathscr{B})\}.\] \end{defi} The following soundness result is well-known and directly follows from the observation that all rules in some fixed deduction system for IPC preserve truth. \begin{prop}{\rm(\cite[Theorem 4.1]{mckinsey-tarski-1948})} For every Brouwer algebra $\mathscr{B}$: $\mathrm{IPC} \subseteq \mathrm{Th}(\mathscr{B})$. \end{prop} \begin{proof} See e.g.\ Chagrov and Zakharyaschev \cite[Theorem 7.10]{chagrov-zakharyaschev-1997}. \end{proof} Conversely, the class of Brouwer algebras is complete for $\mathrm{IPC}$. \begin{thm}{\cite[Theorem 4.3]{mckinsey-tarski-1948})} \[\bigcap\{\mathrm{Th}(\mathscr{B}) \mid \mathscr{B} \text{ a Brouwer algebra}\} = \mathrm{IPC}\] \end{thm} Thus, Brouwer algebras can be used to provide algebraic semantics for $\mathrm{IPC}$. Therefore, it would be nice if the computationally motivated Medvedev lattice has $\mathrm{IPC}$ as its theory, so that it would provide computational semantics for $\mathrm{IPC}$. Unfortunately the weak law of the excluded middle holds in the Medvedev lattice, as can be easily verified. However, as mentioned in the introduction we can still recover $\mathrm{IPC}$ by looking at principal factors of the Medvedev lattice. \begin{prop} Let $\mathscr{B}$ be a Brouwer algebra and let $x,y \in \mathscr{B}$. Then the interval $[x,y]_\mathscr{B} = \{z \in \mathscr{B} \mid x \leq z \leq y\}$ is a sublattice of $\mathscr{B}$. Furthermore, it is a Brouwer algebra under the implication \[u \to_{[x,y]_\mathscr{B}} v = (u \to_\mathscr{B} v) \oplus x.\] \end{prop} \begin{prop} Let $\mathscr{B}$ be a Brouwer algebra and let $x \in \mathscr{B}$. Then $\mathscr{B} / \{z \in \mathscr{B} \mid z \geq x\}$, which we will denote by $\mathscr{B} / x$, is isomorphic as a bounded distributive lattice to $[0,x]_\mathscr{B}$. In particular, $\mathscr{B} / x$ is a Brouwer algebra. \end{prop} Thus, looking at a principal factor $\mathpzc{M} / \mathcal{A}$ is the same as restricting to $[\omega^\omega,\mathcal{A}]_{\mathpzc{M}}$. This means that, when looking at the theory of this factor, we interpret $\bot$ as $\mathcal{A}$ instead of as $\emptyset$. So one might interpret looking at such a factor by replacing the problem $\emptyset$, which is `too hard', by an easier problem $\mathcal{A}$. Finally, we mention one easy lemma which we will use in this paper. \begin{lem}\label{surj-th} Let $\mathscr{B}, \mathscr{C}$ be Brouwer algebras and let $\alpha: \mathscr{B} \to \mathscr{C}$ be a surjective homomorphism. Then $\mathrm{Th}(\mathscr{B}) \subseteq \mathrm{Th}(\mathscr{C})$. \end{lem} \begin{proof} Let $\phi(x_1,\dots,x_n) \not\in \mathrm{Th}(\mathscr{C})$. Fix $c_1,\dots,c_n \in \mathscr{C}$ such that $\phi(c_1,\dots,c_n) \not= 0$. Fix $b_1,\dots,b_n \in \mathscr{B}$ such that $\gamma(b_i) = c_i$. Then \[\alpha(\phi(b_1,\dots,b_n)) = \phi(\alpha(b_1),\dots,\alpha(b_n)) = \phi(c_1,\dots,c_n) \not=0\] because $\alpha$ is a homomorphism. Thus $\phi(b_1,\dots,b_n) \not= 0$ and therefore $\phi \not\in \mathrm{Th}(\mathscr{B})$. \end{proof} \section{Upper implicative semilattice embeddings of $\mathcal{P}(I)$ into $\mathpzc{M}$} As a first step, we will describe a method to embed Boolean algebras of the form $\mathcal{P}(I)$, ordered under reverse inclusion $\supseteq$, into the Medvedev lattice $\mathpzc{M}$ as an upper implicative semilattice (i.e.\ preserving $\oplus$, $\to$, $0$ and $1$). It should be noted that we will only need this for finite $I$, and Skvortsova \cite[Lemma 7]{skvortsova-1988-en} already showed that such embeddings exist. However, Skvortsova used Lachlan's result \cite{lachlan-1968} that every countable distributive lattice can be order-theoretically embedded as an initial segment of the Turing degrees. Because we want natural factors of the Medvedev lattice, we want to avoid the use of this theorem. Our main result of this section will show that there are various natural embeddings of $\mathcal{P}(I)$ into $\mathpzc{M}$. These embeddings are induced by so-called \emph{strong upwards antichains}, where the notion of a strong upwards antichain is the order-dual of the notion of an antichain normally used in forcing. \begin{defi} Let $\mathcal{A} \subseteq \omega^\omega$ be downwards closed under Turing reducibility and let $(f_i)_{i \in I} \in \mathcal{A}^I$. Then we say that $(f_i)_{i \in I}$ is a \emph{strong upwards antichain in $\mathcal{A}$} if for all $i \neq j$ we have that $f_i \oplus f_j \not\in \mathcal{A}$. \end{defi} Henceforth we will mean by \emph{antichain} a \emph{strong upwards antichain}. \begin{ex} We give some examples of countably infinite antichains. \begin{enumerate}[{\rm (i)}] \item Take $\mathcal{A}$ to be the computable functions together with the functions of minimal degree, and $f_0,f_1 \dots$ any sequence of functions of distinct minimal Turing degree. \item Let $f_0,f_1,\dots$ be pairwise incomparable under Turing reducibility and take $\mathcal{A}$ to be the lower cone of $\{f_i \mid i \in \omega\}$. \end{enumerate} \end{ex} The next theorem shows that each antichain induces an upper implicative semilattice embedding of $\mathcal{P}(I)$ in a natural way. \begin{thm}\label{thm-pow-embed} Let $\mathcal{A} \subseteq \omega^\omega$ be downwards closed under Turing reducibility, let $(f_i)_{i \in I}$ be an antichain in $\mathcal{A}$, and let $\mathcal{B} = \overline{\mathcal{A}} \cup C\left(\{f_i \mid i \in I\}\right)$. Then the map $\alpha$ given by $\alpha(X) = \overline{\mathcal{A}} \cup C\left(\{f_i \mid i \in X\}\right)$ is an upper implicative semilattice embedding of $(\mathcal{P}(I),\supseteq)$ into $\Big[\mathcal{B},\overline{\mathcal{A}}\Big]_\mathpzc{M}$. \end{thm} \begin{proof} For ease of notation, if $X \subseteq I$ we will denote by $C(X)$ the set $C\left(\{f_i \mid i \in X\}\right)$. We have: \[\alpha(X \cap Y) = \overline{\mathcal{A}} \cup C(X \cap Y).\] On the other hand, because $\alpha(X)$ and $\alpha(Y)$ are upwards closed their join is just intersection (see Skvortsova \cite[Lemma 5]{skvortsova-1988-en}), and therefore: \[\alpha(X) \oplus \alpha(Y) \equiv_\mathpzc{M} \overline{\mathcal{A}} \cup (C(X) \cap C(Y)).\] Clearly, $\alpha(X \cap Y) \subseteq \overline{\mathcal{A}} \cup (C(X) \cap C(Y))$. Conversely, let $g \in \overline{\mathcal{A}} \cup (C(X) \cap C(Y))$. If $g \not\in \mathcal{A}$ then clearly $g \in \alpha(X \cap Y)$. So, assume $g \in \mathcal{A}$. Let $i \in X, j \in Y$ be such that $g \geq_T f_i$ and $g \geq_T f_j$. Then $f_i \oplus f_j \leq_T g \in \mathcal{A}$ so $f_i \oplus f_j \in \mathcal{A}$. Since $(f_i)_{i \in I}$ is an antichain in $\mathcal{A}$ this can only be the case if $i = j$, so we see that $g \in \alpha(X \cap Y)$. We also have, again by \cite[Lemma 5]{skvortsova-1988-en}: \begin{align*} \alpha(X) &\to_{\big[\mathcal{B},\overline{\mathcal{A}}\big]_\mathpzc{M}} \alpha(Y)\\ &\equiv_\mathpzc{M} \mathcal{B} \oplus \{g \mid \forall h \in \alpha(X) (g \oplus h \in \alpha(Y))\}\\ &\equiv_\mathpzc{M} \{g \in \mathcal{B} \mid \forall i \in X \forall h \geq_T f_i \exists j \in Y (g \oplus h \in \mathcal{A} \to g \oplus h \geq_T f_j)\}\\ &= \overline{\mathcal{A}} \cup \{g \in C\left(\{f_i \mid i \in I\}\right)\notag\\ &\quad\quad\quad\quad\quad\mid \forall i \in X \forall h \geq_T f_i \exists j \in Y (g \oplus h \in \mathcal{A} \to g \oplus h \geq_T f_j)\}. \end{align*} Fix any $g \in \mathcal{A} \cap C\left(\{f_i \mid i \in I\}\right)$ such that \begin{equation}\label{eqn1} \forall i \in X \forall h \geq_T f_i \exists j \in Y (g \oplus h \in \mathcal{A} \to g \oplus h \geq_T f_j). \end{equation} Then we know that there is some $k \in I$ such that $g \geq_T f_k$. We claim: $k \not\in X$ or $k \in Y$. Namely, assume $k \in X$ and $k \not\in Y$. Then, by \eqref{eqn1} (with $h=g$) there exists some $j \in Y$ such that $g \geq_T f_j$, and since $k \not\in Y$ we know that $j \neq k$. But then $f_k \oplus f_j \leq_T g \in \mathcal{A}$ so $f_k \oplus f_j \in \mathcal{A}$, a contradiction with the fact that $(f_i)_{i \in I}$ is an antichain in $\mathcal{A}$. Conversely, if $g \in \mathcal{A}$ is such that $g \geq_T f_k$ for some $k \not\in X$ or some $k \in Y$, then \eqref{eqn1} holds: namely, if $k \not\in X$ then we have for all $i \in X$ that $g \oplus f_i \not\in \mathcal{A}$ because $(f_i)_{i \in I}$ is an antichain in $\mathcal{A}$, while if $k \in Y$ we have that $g \oplus f_i \geq_T f_k$. So, from this we see: \begin{align*} \alpha(X) \to_{\big[\mathcal{B},\overline{\mathcal{A}}\big]_\mathpzc{M}} \alpha(Y) &\equiv_\mathpzc{M} \overline{\mathcal{A}} \cup C((I \setminus X) \cup Y)\\ &= \alpha((I \setminus X) \cup Y)\\ &= \alpha(X \to_{\mathcal{P}(I)} Y).\qedhere \end{align*} \end{proof} \section{From embeddings of $\mathcal{P}(\omega)$ to factors capturing IPC}\label{sec-to-factors} In this section we will show how to construct a more natural factor of the Medvedev lattice with IPC as its theory; that is, we will prove Theorem \ref{thm-main}. For this proof we will use several ideas from Skvortsova's construction of a factor of the Medvedev lattice which has IPC as its theory, given in Skvortsova \cite{skvortsova-1988-en}. We combine these ideas with our own to get to the factor in Theorem \ref{thm-main}. First, let us discuss canonical subsets of a Brouwer algebra. \begin{defi}{\rm (\cite[p.\ 134]{skvortsova-1988-en})} Let $\mathscr{B}$ be a Brouwer algebra and let $\mathscr{C} \subseteq \mathscr{B}$. Then we call $\mathscr{C}$ \emph{canonical} if: \begin{enumerate}[\rm (i)] \item \label{canon-1}All elements in $\mathscr{C}$ are meet-irreducible, \item \label{canon-2}$\mathscr{C}$ is closed under joins and implications (i.e.\ it is a sub-upper implicative semilattice), \item \label{canon-3}For all $a \in \mathscr{C}$ and $b,c \in \mathcal{B}$ we have $a \to (b \otimes c) = (a \to b) \otimes (a \to c)$. \end{enumerate} \end{defi} \begin{prop}{\rm (\cite[Corollary to Lemma 6]{skvortsova-1988-en})}\label{prop-muchnik-can} The set of Muchnik degrees is a canonical subset of $\mathpzc{M}$. \end{prop} \begin{cor}\label{cor-range-canonical} The range of $\alpha$ from Theorem \ref{thm-pow-embed} is canonical in $[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}$. \end{cor} \begin{proof} The range of $\alpha$ consists of Muchnik degrees, so \eqref{canon-1} holds by Proposition \ref{prop-muchnik-can}. Furthermore, $\alpha$ is an upper implicative semilattice embedding, and therefore \eqref{canon-2} also holds. Finally, if $\mathcal{C}_0,\mathcal{C}_1 \in [\alpha(I),\alpha(\emptyset)]_\mathpzc{M}$ and $X \subseteq I$, then we see, using Proposition \ref{prop-muchnik-can}: \begin{align*} \alpha(X) &\to_{[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}} (\mathcal{C}_0 \otimes \mathcal{C}_1)\\ &= (\alpha(X) \to_\mathpzc{M} (\mathcal{C}_0 \otimes \mathcal{C}_1)) \oplus \alpha(I)\\ &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} \mathcal{C}_0) \otimes (\alpha(X) \to_\mathpzc{M} \mathcal{C}_1)) \oplus \alpha(I)\\ &\equiv_\mathpzc{M} (\alpha(X) \to_{[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}} \mathcal{C}_0) \otimes (\alpha(X) \to_{[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}} \mathcal{C}_1).\qedhere \end{align*} \end{proof} \begin{prop}{\rm (\cite[Lemma 2]{skvortsova-1988-en})}\label{prop-free-canonical} If $\mathscr{C}$ is a canonical set in a Brouwer algebra $\mathscr{B}$, then the smallest sub-Brouwer algebra of $\mathscr{B}$ containing $\mathscr{C}$ is $\{a_1 \otimes \dots \otimes a_n \mid a_i \in \mathscr{C}\}$, and it is isomorphic to the free Brouwer algebra over the upper implicative semilattice $\mathscr{C}$ through an isomorphism fixing $\mathscr{C}$. \end{prop} In particular, we see: \begin{cor}\label{cor-emb-ext} If we let $\alpha$ be the embedding of $(\mathcal{P}(I),\supseteq)$ from Theorem \ref{thm-pow-embed}, then $\{\alpha(X_1) \otimes \dots \otimes \alpha(X_n) \mid X_i \in \mathcal{P}(I)\}$ is a sub-Brouwer algebra of $[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}$ which is isomorphic to the free Brouwer algebra over the upper implicative semilattice $(\mathcal{P}(I),\supseteq)$. \end{cor} \begin{proof} From Corollary \ref{cor-range-canonical} and Proposition \ref{prop-free-canonical}. \end{proof} Let $\mathscr{B}_n$ be the Brouwer algebra of the upwards closed subsets of $(\mathcal{P}(\{1,\dots,n\}) \setminus \{\emptyset\},\supseteq)$ ordered under reverse inclusion $\supseteq$, i.e.\ the elements of $\mathscr{B}_n$ are those $A \subseteq \mathcal{P}(\{1,\dots,n\}) \setminus \emptyset$ such that if $X \in A$ and $Y \in \mathcal{P}(\{1,\dots,n\}) \setminus \{\emptyset\}$ is such that $X \supseteq Y$, then $Y \in A$. We can use $\mathscr{B}_n$ to capture IPC in the following way: \begin{prop}{\rm (\cite[the remark following Lemma 3]{skvortsova-1988-en})}\label{prop-free-n-ipc} $\bigcap_{n > 0} \bigcap_{x \in \mathscr{B}_n} \mathrm{Th}(\mathscr{B}_n / x) = \mathrm{IPC}$. \end{prop} \begin{proof} Let $\mathrm{LM} = \bigcap_{n > 0} \mathrm{Th}(\mathscr{B}_n)$, the \emph{Medvedev logic of finite problems}. Given a set of formulas $X$, let $X^+$ denote the set of positive (i.e.\ negation-free) formulas in $X$. Then $\mathrm{LM}^+ = \mathrm{IPC}^+$, see Medvedev \cite{medvedev-1962}. Now, let $\phi(x_1,\dots,x_m)$ be any formula. Let $\phi'(x_1,\dots,x_{m+1})$ be the formula where $x_{m+1}$ is a fresh variable and where $\bot$ is replaced by $x_1 \wedge \dots \wedge x_{m+1}$, so $\phi'$ is negation-free. Then, if $\phi \not\in \mathrm{IPC}$, we have $\phi' \not\in \mathrm{IPC}^+$ (see Jankov \cite{jankov-1968-2}), so there are $n \in \omega$ and $x_1,\dots,x_{m+1} \in \mathscr{B}_n$ such that $\phi'(x_1,\dots,x_{m+1}) \not= 0$. Let $x = x_1 \oplus \dots \oplus x_{m+1}$, then $\phi \not\in \mathrm{Th}(\mathscr{B}_n / x)$. \end{proof} Furthermore, it is easy to obtain these $\mathscr{B}_n$ as free distributive lattices over upper implicative semilattices, as expressed by the following proposition. \begin{prop}{\rm (\cite[Lemma 3]{skvortsova-1988-en})}\label{prop-free-n-isom} The Brouwer algebra $\mathscr{B}_n$ is isomorphic to the free distributive lattice over the upper implicative semilattice $(\mathcal{P}(\{1,\dots,n\}),\supseteq)$. \end{prop} \begin{cor}\label{cor-n-isom} Let $I$ be a set of size $n$. If we let $\alpha$ be the embedding of $(\mathcal{P}(I),\supseteq)$ from Theorem \ref{thm-pow-embed}, then $\{\alpha(X_1) \otimes \dots \otimes \alpha(X_m) \mid m \in \omega \wedge \forall i \leq m (X_i \in \mathcal{P}(I))\}$ is a sub-Brouwer algebra of $[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}$ isomorphic to $\mathscr{B}_n$. \end{cor} \begin{proof} From Corollary \ref{cor-emb-ext} and Proposition \ref{prop-free-n-isom}. \end{proof} The following lemma allows us to compare the theories of different intervals. \begin{lem}{\rm (\cite[Lemma 4]{skvortsova-1988-en})}\label{lem-intervals} In any Brouwer algebra $\mathscr{B}$: if $x,y,z \in \mathscr{B}$ are such that $x \oplus z = y$, then $\mathrm{Th}([0,z]_\mathscr{B}) \subseteq \mathrm{Th}([x,y]_\mathscr{B})$. \end{lem} \begin{proof} Let $\gamma: [0,z]_\mathscr{B} \to [x,y]_\mathscr{B}$ be given by $\gamma(u) = x \oplus u$. This map is well-defined, since if $u \leq z$, then $x \oplus u \leq x \oplus z = y$. Clearly $\gamma$ preserves $\oplus$ and $\otimes$, while for $\to$ we have: \[\gamma(u \to_{[0,z]_\mathscr{B}} v) = (u \to_\mathscr{B} v) \oplus x = ((u \oplus x) \to_\mathscr{B} (v \oplus x)) \oplus x = \gamma(u) \to_{[x,y]_\mathscr{B}} \gamma(v).\] Furthermore, $\gamma$ is surjective, so the result now follows from Lemma \ref{surj-th}. \end{proof} Before we get to the proof of Theorem \ref{thm-main} we need one theorem from computability theory. \begin{thm}\label{thm-splitting-ext} Let $A,E \in 2^\omega$ be such that $E \geq_T A'$. Let $B_0,B_1,\dots \in 2^\omega$ be uniformly computable in $E$ and such that $A \not\geq_T B_i$ . Then there exists a set $D \geq_T A$ such that $D' \leq_T E$ and such that for all $i \in \omega$ we have $D \oplus B_i \geq_T E$. \end{thm} \begin{proof} This follows from relativising Posner and Robinson \cite[Theorem 3]{posner-robinson-1981} to $A$. \end{proof} Finally, we need an easy lemma on extending computably independent sets. For ease of notation, let us assume that our pairing function is such that $(A \oplus B)^{[2i]} = A^{[i]}$ and $(A \oplus B)^{[2i+1]} = B^{[i]}$. \begin{lem}\label{lem-extend} Let $A$ be a computably independent set. Then there exists a set $B$ such that $A \oplus B$ is computably independent. \end{lem} \begin{proof} Our requirements are as follows: \begin{align*} R_{\langle e,2n \rangle}&: A^{[n]} \not= \{e\}^{\bigoplus_{i \not= 2n} (A \oplus B)^{[i]}}\\ R_{\langle e,2n+1 \rangle}&: B^{[n]} \not= \{e\}^{\bigoplus_{i \not= 2n+1} (A \oplus B)^{[i]}}. \end{align*} We build $B$ by the finite extension method, i.e.\ we define strings $\sigma_0 \subseteq \sigma_1 \subseteq \dots$ and let $B = \bigcup_{s \in \omega} \sigma_s$. For ease of notation, define $\sigma_{-1} = \emptyset$. At stage $s$, we deal with requirement $R_s$. There are two cases: \begin{itemize} \item $s = \langle e,2n \rangle$: if there is a string $\sigma$ extending $\sigma_{s-1}$ and an $m \in \omega$ such that $\{e\}^{\bigoplus_{i \not= 2n} (A \oplus \sigma)^{[i]}}(m){\downarrow} \not= A^{[n]}(m)$, take $\sigma_s$ to be the least such $\sigma$. Otherwise, let $\sigma_s = \sigma_{s-1}$. \item $s = \langle e,2n+1 \rangle$: if there exists a string $\sigma$ extending $\sigma_{s-1}$ such that we have $\{e\}^{\bigoplus_{i \not= 2n+1} (A \oplus \sigma)^{[i]}}(|\sigma_{s-1}| + 1){\downarrow}$, take the least such $\sigma$ and let $\sigma_s$ be the least string extending $\sigma_{s-1}$ which coincides with $\sigma$ outside the $n^\mathrm{th}$ column and such that $\sigma_s^{[n]}(|\sigma_{s-1}| + 1) = 1 - \{e\}^{\bigoplus_{i \not= 2n+1} (A \oplus \sigma)^{[i]}}(|\sigma_{s-1}| + 1)$. Otherwise, let $\sigma_s = \sigma_{s-1}$. \end{itemize} We claim: $B$ is as required. To this end, we verify the requirements: \begin{itemize} \item $R_{\langle e,2n \rangle}$: towards a contradiction, assume $A^{[n]} = \{e\}^{\bigoplus_{i \not= 2n} (A \oplus B)^{[i]}}$. Let $s = \langle e,2n \rangle$. By construction we then know for every $\sigma$ extending $\sigma_{s-1}$ and every $m \in \omega$ that, if $\{e\}^{\bigoplus_{i \not= 2n} (A \oplus \sigma)^{[i]}}(m){\downarrow}$, we have $\{e\}^{\bigoplus_{i \not= 2n} (A \oplus \sigma)^{[i]}}(m) = A^{[n]}(m)$. Furthermore, for every $m \in \omega$ there is a string $\sigma$ extending $\sigma_{s-1}$ such that $\{e\}^{\bigoplus_{i \not= 2n} (A \oplus \sigma)^{[i]}}(m){\downarrow}$: just take a suitably long initial segment of $B$. However, this means that $\bigoplus_{i \not= n} A^{[i]} \geq_T A^{[n]}$, which contradicts $A$ being computably independent. \item $R_{\langle e,2n+1 \rangle}$: let $s = \langle e,2n+1 \rangle$. Then by our construction we know that, if $\{e\}^{\bigoplus_{i \not= 2n+1} (A \oplus B)^{[i]}}(|\sigma_{s-1}| + 1) \downarrow$, then it differs from $B^{[n]}(|\sigma_{s-1}| + 1)$. \end{itemize} \end{proof} We can now prove Theorem \ref{thm-main}. \mainthm* \begin{proof} Fix $n \in \omega$ and $x \in \mathscr{B}_n$. Let $I = \{1,\dots,n\}$. For now assume we have some downwards closed $\mathcal{A}$ and an antichain $D_1,\dots,D_n \in \mathcal{A}$. Then Corollary \ref{cor-n-isom} tells us that \[\{\alpha(Y_1) \otimes \dots \otimes \alpha(Y_m) \mid m \in \omega \wedge \forall i \leq m(Y_i \in \mathcal{P}(I))\}\] is a subalgebra of $\Big[\overline{\mathcal{A}} \cup C(\{D_1,\dots,D_n\}),\overline{\mathcal{A}}\Big]_\mathpzc{M}$ isomorphic to $\mathscr{B}_n$. So, there are $X_1,\dots,X_k \subseteq I$ such that we can embed $\mathscr{B}_n / x$ as subalgebra of \[\Big[\overline{\mathcal{A}} \cup C(\{D_1,\dots,D_n\}), \alpha(X_1) \otimes \dots \otimes \alpha(X_k) \Big]_\mathpzc{M}.\] If we would additionally have that \begin{equation}\label{thm-main-eqn1} \Big(\overline{\mathcal{A}} \cup C(\{D_1,\dots,D_n\})\Big) \oplus \left\{i \conc f \mid f \geq_T A^{[i]}\right\} \equiv_\mathpzc{M} \alpha(X_1) \otimes \dots \otimes \alpha(X_k), \end{equation} then Lemma \ref{lem-intervals} tells us that \begin{align*} \mathrm{Th}\Big(\mathpzc{M} / \Big\{i \conc f &\mid f \geq_T A^{[i]}\Big\}\Big)\\ &\subseteq \mathrm{Th}\Big(\Big[\overline{\mathcal{A}} \cup C(\{D_1,\dots,D_n\}),\alpha(X_1) \otimes \dots \otimes \alpha(X_k)\Big]_\mathpzc{M}\Big)\\ &\subseteq \mathrm{Th}(\mathscr{B}_n / x). \end{align*} Now, if we would be able to do this for arbitrary $n \in \omega$ and $x \in \mathscr{B}_n$, then Proposition \ref{prop-free-n-ipc} tells us that \[\mathrm{Th}\left(\mathpzc{M} / \left\{i \conc f \mid f \geq_T A^{[i]}\right\}\right) = \mathrm{IPC},\] so then we would be done. \bigskip Thus, it suffices to show that for all $n \in \omega$ and all $X_1,\dots,X_k \subseteq \{1,\dots,n\}$ there exists a downwards closed $\mathcal{A}$ and an antichain $D_1,\dots,D_n \in \mathcal{A}$ such that \eqref{thm-main-eqn1} holds. Fix a $B$ for $A$ as in Lemma \ref{lem-extend}. Let $\mathcal{A} = \omega^\omega \setminus C(\{(A \oplus B)'\})$. For every $1 \leq i \leq n$ fix a $D_i \geq_T \left(\bigoplus_{1 \leq j \leq k, i \in X_j} A^{[j]}\right) \oplus B^{[i]}$ such that $D_i' \leq_T (A \oplus B)'$, such that $D_i \oplus A^{[j]} \geq_T (A \oplus B)'$ for every $j \in \{1 \leq j \leq k \mid i \not\in X_j\} \cup \{k+1,k+2,\dots\}$ and such that $D_i \oplus B^{[j]} \geq_T (A \oplus B)'$ for every $j \not= i$, which exists by Theorem \ref{thm-splitting-ext}. We claim: $\{D_1,\dots,D_n\}$ is an antichain in $\mathcal{A}$. Clearly, $D_1,\dots,D_n \in \mathcal{A}$. Next, let $1 \leq i < j \leq n$. Then: \[D_i \oplus D_j \geq_T B^{[i]} \oplus D_j \geq_T (A \oplus B)',\] so $D_i \oplus D_j \not\in \mathcal{A}$. Thus, we need to show that \eqref{thm-main-eqn1} holds. First, let $g \in \overline{\mathcal{A}} \cup C(\{D_1,\dots,D_n\})$ and let $f \geq_T A^{[j]}$. If $j > k$, then either $g \geq_T D_i$ for some $1 \leq i \leq n$ and $f \oplus g \geq_T A^{[j]} \oplus D_i \geq_T (A \oplus B)'$, or $g \geq_T (A \oplus B)'$ and then also $f \oplus g \geq_T (A \oplus B)'$. In both cases we see that $f \oplus g \in \overline{\mathcal{A}} \subseteq \alpha(X_1)$. Thus, we may assume that $j \leq k$. We claim: $f \oplus g \in \alpha(X_j)$. Indeed, if $g \geq_T D_i$ for some $i \in X_j$, then $f \oplus g \geq_T D_i$ and $C(D_i) \subseteq \alpha(X_j)$, while if $g \geq_T D_i$ for some $i \not\in X_j$, then $f \oplus g \geq_T A^{[j]} \oplus D_i \geq_T (A \oplus B)'$, and finally, if $g \geq_T (A \oplus B)'$ then clearly $f \oplus g \geq_T (A \oplus B)'$. Thus, we see that $f \oplus g$ computes an element of $\alpha(X_1) \otimes \dots \otimes \alpha(X_k)$, and that this computation is in fact uniform in $(j \conc f) \oplus g$. For the other direction, note that for fixed $1 \leq i \leq k$ we have that \[C(X_i) \subseteq C(\{D_1,\dots,D_n\})\] and also that \[C(X_i) \subseteq \left\{f \mid f \geq_T A^{[i]}\right\}\] because for every $j \in X_i$ we have that $D_j \geq_T A^{[i]}$. \end{proof} \section{Relativising the construction} We will next show that Skvortsova's construction can be performed below every mass problem $\mathcal{B} >_\mathpzc{M} 0'$. This also implies that for every $\mathcal{B} >_\mathpzc{M} 0'$ we have that $\mathrm{Th}(\mathpzc{M} / \mathcal{B}) \subseteq \mathrm{Jan}$, answering a question by Sorbi and Terwijn; see Corollary \ref{cor-jan} below. First, note that for every $\mathcal{B} > 0'$ we can find a countable mass problem $\mathcal{E} \subseteq 0'$ such that $\mathcal{E} \not\geq_\mathpzc{M} \mathcal{B}$ (e.g.\ by taking one function for every $n \in \omega$ witnessing that $\Phi_n(0') \not\subseteq \mathcal{B}$). Then the set $\{A \mid \forall f \in \mathcal{E} (A \not\geq_T f)\}$ has measure $1$ (by Sack's result that upper cones in the Turing degrees have measure $0$, see e.g.\ Downey and Hirschfeldt \cite[Corollary 8.12.2]{downey-hirschfeldt-2010}), so it contains a 1-random set; in particular it contains a computably independent set $A$. In this section we will show that we can use such sets to obtain factors with theory $\mathrm{IPC}$ below $\mathcal{B}$, by relativising Theorem \ref{thm-main}. However, we first show that we can relativise Theorem \ref{thm-pow-embed} below $\mathcal{B}$. \begin{thm}\label{thm-pow-embed-rel} Let $\mathcal{B}$ be a mass problem, let $\mathcal{E}$ be a mass problem such that $\mathcal{E} \not\geq_\mathpzc{M} \mathcal{B}$ and let $\mathcal{D} = \mathcal{E} \to_\mathpzc{M} \mathcal{B}$. Let $\mathcal{A} \subseteq \omega^\omega$ be a mass problem which is downwards closed under Turing reducibility such that $\mathcal{E} \subseteq \overline{\mathcal{A}}$. Let $(f_i)_{i \in I}$ be an antichain in $\mathcal{A}$. Then the map $\beta$ given by $\beta(X) = (\overline{\mathcal{A}} \cup \{g \mid \exists i \in X (g \geq_T f_i)\}) \otimes \mathcal{D}$ is an upper implicative semilattice embedding of $(\mathcal{P}(I),\supseteq)$ into $[\beta(I),\beta(\emptyset)]_\mathpzc{M}$ with range canonical in $[\beta(I),\beta(\emptyset)]_\mathpzc{M}$. \end{thm} \begin{proof} First, note that $\mathcal{E} \not\geq_\mathpzc{M} \mathcal{D}$, since if $\mathcal{E} \geq_\mathpzc{M} \mathcal{D}$ then \[\mathcal{E} \equiv_\mathpzc{M} \mathcal{E} \oplus \mathcal{D} = \mathcal{E} \oplus (\mathcal{E} \to \mathcal{B}) \geq_\mathpzc{M} \mathcal{B},\] a contradiction. As in the proof of Theorem \ref{thm-pow-embed}, if $X \subseteq I$ we will denote by $C(X)$ the set $C(\{f_i \mid i \in X\})$. By Theorem \ref{thm-pow-embed}, the function $\alpha: \mathcal{P}(I) \to \mathrm{\mathpzc{M} / \overline{\mathcal{A}}}$ given by $\alpha(X) = \overline{\mathcal{A}} \cup C(X)$ is an upper implicative semilattice embedding of $(\mathcal{P}(I),\supseteq)$ into $\Big[\overline{\mathcal{A}} \cup C(I),\overline{\mathcal{A}}\Big]_\mathpzc{M}$. Note that $\mathcal{E} \subseteq \overline{\mathcal{A}}$ and therefore $\mathcal{E} \subseteq \alpha(X)$ for every $X \subseteq I$. Now let $\beta: \mathcal{P}(I) \to \mathrm{\mathpzc{M} / \overline{\mathcal{A}}}$ be the function given by $\beta(X) = \alpha(X) \otimes \mathcal{D}$. Then the range of $\beta$ is certainly contained in $[\beta(I),\beta(\emptyset)]_\mathpzc{M}$. We prove that $\beta$ is in fact an upper implicative semilattice embedding into $[\beta(I),\beta(\emptyset)]_\mathpzc{M}$ with canonical range. \begin{itemize} \item $\beta$ is injective: assume $\beta(X) \leq_\mathpzc{M} \beta(Y)$. Thus, we have $\alpha(X) \otimes \mathcal{D} \leq_\mathpzc{M} \alpha(Y) \otimes \mathcal{D}$. In particular we have that $\alpha(X) \otimes \mathcal{D} \leq_\mathpzc{M} \alpha(Y)$, say via $\Phi_n$. We claim: $\Phi_n(\alpha(Y)) \subseteq 0 \conc \alpha(X)$. Namely, assume towards a contradiction that $\Phi_n(f) \in 1 \conc \mathcal{D}$ for some $f \in \alpha(Y)$. Determine $\sigma \subseteq f$ such that $\Phi_n(\sigma)(0) = 1$. As noted above we have that $\mathcal{E} \subseteq \alpha(Y)$, and since $\alpha(Y)$ is Muchnik we therefore see that $\sigma \conc \mathcal{E} \subseteq \alpha(Y)$. However, then we can reduce $\mathcal{E}$ to $1 \conc \mathcal{D}$ by sending $g \in \mathcal{E}$ to $\Phi_n(\sigma \conc g)$, and therefore $\mathcal{E} \geq_\mathpzc{M} \mathcal{D}$, a contradiction. Thus, $\alpha(X) \leq_\mathpzc{M} \alpha(Y)$, and since $\alpha$ is an upper implicative semilattice embedding this tells us that $X \supseteq Y$. \item $\beta$ preserves joins: we have \begin{align*} \beta(X \oplus Y) &= \alpha(X \oplus Y) \otimes \mathcal{D} \equiv_\mathpzc{M} (\alpha(X) \oplus \alpha(Y)) \otimes \mathcal{D}\\ &\equiv_\mathpzc{M} (\alpha(X) \otimes \mathcal{D}) \oplus (\alpha(Y) \otimes \mathcal{D}) = \beta(X) \oplus \beta(Y). \end{align*} \item $\beta$ preserves implications: we have \begin{align*} \beta(X) &\to_{[\beta(I),\beta(\emptyset)]_\mathpzc{M}} \beta(Y)\\ &= ((\alpha(X) \otimes \mathcal{D}) \to_\mathpzc{M} (\alpha(Y) \otimes \mathcal{D})) \oplus \beta(I)\\ &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} (\alpha(Y) \otimes \mathcal{D})) \oplus (\mathcal{D} \to_\mathpzc{M} (\alpha(Y) \otimes \mathcal{D}))) \oplus \beta(I)\\ &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} (\alpha(Y) \otimes \mathcal{D})) \oplus \omega^\omega) \oplus \beta(I)\\ &\equiv_\mathpzc{M} (\alpha(X) \to_\mathpzc{M} (\alpha(Y) \otimes \mathcal{D})) \oplus \beta(I).\\ \intertext{Next, using Proposition \ref{prop-muchnik-can} we see:} &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} \alpha(Y)) \otimes (\alpha(X) \to_\mathpzc{M} \mathcal{D})) \oplus \beta(I)\\ &= ((\alpha(X) \to_\mathpzc{M} \alpha(Y)) \otimes (\alpha(X) \to_\mathpzc{M} (\mathcal{E} \to_\mathpzc{M} \mathcal{B}))) \oplus \beta(I)\\ &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} \alpha(Y)) \otimes ((\alpha(X) \oplus \mathcal{E}) \to_\mathpzc{M} \mathcal{B})) \oplus \beta(I).\\ \intertext{As noted above, we have $\mathcal{E} \subseteq \alpha(X)$, and therefore:} &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} \alpha(Y)) \otimes (\mathcal{E} \to_\mathpzc{M} \mathcal{B})) \oplus \beta(I)\\ &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} \alpha(Y)) \otimes \mathcal{D}) \oplus (\alpha(I) \otimes \mathcal{D}).\\ &\equiv_\mathpzc{M} ((\alpha(X) \to_\mathpzc{M} \alpha(Y)) \oplus \alpha(I)) \otimes \mathcal{D}\\ &= \left(\alpha(X) \to_{[\alpha(I),\alpha(\emptyset)]_\mathpzc{M}} \alpha(Y)\right) \otimes \mathcal{D}\\ &= \alpha(X \to_{\mathcal{P}(I)} Y) \otimes \mathcal{D}\\ &= \beta(X \to_{\mathcal{P}(I)} Y). \end{align*} \item $\beta$ has canonical range: \begin{enumerate}[\rm (i)] \item Let $X \subseteq I$, we show that that $\beta(X)$ is meet-irreducible in $[\beta(I),\beta(\emptyset)]_\mathpzc{M}$. Indeed, let $\mathcal{C}_0,\mathcal{C}_1 \leq_\mathpzc{M} \beta(\emptyset)$ be such that $\mathcal{C}_0 \otimes \mathcal{C}_1 \leq_\mathpzc{M} \alpha(X) \otimes \mathcal{D}$. Then $\mathcal{C}_0 \otimes \mathcal{C}_1 \leq_\mathpzc{M} \alpha(X)$, and since $\alpha(X)$ is Muchnik, we see from Proposition \ref{prop-muchnik-can} that $\mathcal{C}_0 \leq_\mathpzc{M} \alpha(X)$ or $\mathcal{C}_1 \leq_\mathpzc{M} \alpha(X)$. Since $\mathcal{C}_0,\mathcal{C}_1 \leq_\mathpzc{M} \beta(\emptyset) \leq_\mathpzc{M} \mathcal{D}$ this shows that in fact $\mathcal{C}_0 \leq_\mathpzc{M} \beta(X)$ or $\mathcal{C}_1 \leq_M \beta(X)$. \item The range of $\beta$ is clearly closed under implication and joins. \item Let $X \subseteq \omega$ and let $\mathcal{C}_0,\mathcal{C}_1 \in [\beta(I),\beta(\emptyset)]_\mathpzc{M}$. Then we have: \begin{align*} \beta(X) &\to_{[\beta(I),\beta(\emptyset)]_\mathpzc{M}} (\mathcal{C}_0 \otimes \mathcal{C}_1)\\ &=(\alpha(X) \otimes \mathcal{D}) \to_{[\beta(I),\beta(\emptyset)]_\mathpzc{M}} (\mathcal{C}_0 \otimes \mathcal{C}_1)\\ &=((\alpha(X) \otimes \mathcal{D}) \to_\mathpzc{M} (\mathcal{C}_0 \otimes \mathcal{C}_1)) \oplus \beta(I)\\ &\equiv_\mathpzc{M} (\alpha(X) \to_\mathpzc{M} (\mathcal{C}_0 \otimes \mathcal{C}_1)) \oplus \beta(I),\\ \intertext{because $\mathcal{C}_0$ and $\mathcal{C}_1$ are below $\beta(\emptyset)$ and hence below $\mathcal{D}$. Since $\alpha(X)$ is Muchnik, we now see from Proposition \ref{prop-muchnik-can}:} &= ((\alpha(X) \to_\mathpzc{M} \mathcal{C}_0) \otimes (\alpha(X) \to_\mathpzc{M} \mathcal{C}_1)) \oplus \beta(I)\\ &\equiv_\mathpzc{M} (\beta(X) \to_{[\beta(I),\beta(\emptyset)]_\mathpzc{M}} \mathcal{C}_0) \otimes (\beta(X) \to_{[\beta(I),\beta(\emptyset)]_\mathpzc{M}} \mathcal{C}_1).\qedhere \end{align*} \end{enumerate} \end{itemize} \end{proof} We can now prove there is a principal factor of the Medvedev lattice with theory IPC below a given $\mathcal{B} > 0'$. \begin{thm}\label{thm-factor-rel} Let $\mathcal{B}$ be a mass problem, let $\mathcal{E}$ be a countable mass problem such that $\mathcal{E} \not\geq_\mathpzc{M} \mathcal{B}$ and let $\mathcal{D} = \mathcal{E} \to \mathcal{B}$ (so, $\mathcal{D} \leq_\mathpzc{M} \mathcal{B}$). Let $A$ be a computably independent set such that for all $f \in \mathcal{E}$ we have $A \not\geq_T f$. Then \[\mathrm{Th}\left(\mathpzc{M} / \left(\left\{i \conc g \mid g \geq_T A^{[i]} \text{ or } g \in C(\mathcal{E})\right\} \otimes \mathcal{D}\right)\right) = \mathrm{IPC}.\] \end{thm} \begin{proof} The proof largely mirrors that of Theorem \ref{thm-main}. Let $\mathcal{E} = \{f_0,f_1,\dots\}$, let $E_i$ be the graph of $f_i$ and let $U$ be such that $U^{[0]} = A$ and $U^{[i+1]} = E_i$. Then $A,E_0,E_1,\dots$ is uniformly computable in $U$. We need to make a slight modification to Lemma \ref{lem-extend}: we not only want $A \oplus B$ to be computably independent, but we also need to make sure that $A \oplus B \not\geq_T f$ for every $f \in \mathcal{E}$. This modification is straightforward and we omit the details. The requirements on $D_i$ are slightly different: we now want for every $1 \leq i \leq k$ that $D_i \geq_T \bigoplus_{1 \leq j \leq k, i \in X_j} A^{[j]} \oplus B^{[i]}$, that $D_i' \leq_T (U \oplus B)'$, that $D_i \oplus A^{[j]} \geq_T (U \oplus B)'$ for every $j \in \{1 \leq j \leq k \mid i \not\in X_j\} \cup \{k+1,k+2,\dots\}$, that $D_i \oplus B^{[j]} \geq_T (U \oplus B)'$ for every $j \not= i$ and that $D_i \oplus E_j \geq_T (U \oplus B)'$ for all $j \in \omega$; this is still possible by Theorem \ref{thm-splitting-ext}. We change the definition of $\mathcal{A}$ into $\mathcal{A} = \omega^\omega \setminus C(\{(U \oplus B)'\} \cup \mathcal{E})$. Then we still have $D_i \in \mathcal{A}$, because $D_i \geq_T f_j$ would imply that $D_i \geq_T D_i \oplus E_j \geq_T (U \oplus B)'$, a contradiction. Finally, replace $\alpha$ with the $\beta$ of Theorem \ref{thm-pow-embed-rel} and change \eqref{thm-main-eqn1} into \begin{align*} \left(\Big(\overline{\mathcal{A}} \cup C(\{D_1,\dots,D_n\})\Big) \otimes \mathcal{D}\right) &\oplus \left(\left\{i \conc g \mid g \geq_T A^{[i]} \text{ or } g \in C(\mathcal{E})\right\} \otimes \mathcal{D}\right)\\ &\equiv_\mathpzc{M} \beta(X_1) \otimes \dots \otimes \beta(X_k). \end{align*} Then the whole proof of Theorem \ref{thm-main} goes through. \end{proof} In particular, this allows us to give a positive answer to the question mentioned at the beginning of this section. \begin{cor}\label{cor-jan} Let $\mathcal{B} >_\mathpzc{M} 0'$. Then $\mathrm{Th}(\mathpzc{M} / \mathcal{B}) \subseteq \mathrm{Jan}$. \end{cor} \begin{proof} Since an intermediate logic is contained in $\mathrm{Jan}$ if and only if its positive fragment coincides with $\mathrm{IPC}$ (see Jankov \cite{jankov-1968}), we need to show that, denoting the positive fragment by $^{+}$, we have that $\mathrm{Th}^{+}(\mathpzc{M} / \mathcal{B}) \subseteq \mathrm{IPC}^{+}$. By Theorem \ref{thm-factor-rel} there exists a $\mathcal{C} \leq_\mathpzc{M} \mathcal{B}$ such that $\mathrm{Th}(\mathpzc{M} / \mathcal{C}) = \mathrm{IPC}$. Then $\mathpzc{M} / \mathcal{C}$ is a subalgebra of $\mathpzc{M} / \mathcal{B}$, except for the fact that the top element is not necessarily preserved. However, it can be directly verified that for any two Brouwer algebras $\mathscr{C}$ and $\mathscr{B}$ for which $\mathscr{C}$ is a $(\oplus,\otimes,\to,0)$-subalgebra of $\mathscr{B}$ we have for all positive formulas $\phi(x_1,\dots,x_n)$ and all elements $b_1,\dots,b_n \in \mathscr{B}$ that the interpretation of $\phi$ at $b_1,\dots,b_n$ is the same in both $\mathscr{C}$ and $\mathscr{B}$. Since we can refute every positive formula $\phi$ which is not in $\mathrm{IPC}^+$ in $\mathpzc{M} / \mathcal{C}$, we can therefore refute it in $\mathpzc{M} / \mathcal{B}$ using the same valuation. In other words, $\mathrm{Th}^{+}(\mathpzc{M} / \mathcal{B}) \subseteq \mathrm{Th}^{+}(\mathpzc{M} / \mathcal{C}) = \mathrm{IPC}^{+}$, as desired. \end{proof} \section*{Acknowledgements} The author thanks Sebastiaan Terwijn for helpful discussions on the subject. Furthermore, the author thanks the anonymous referees for their many useful comments. \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2]{% \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
\section{Introduction} Following the latest combination of mass and signal strengths of the Higgs boson by the ATLAS and CMS collaborations of the Large Hadron Collider (LHC) \cite{Higgs_ATLAS, Higgs_CMS}, the particle spectrum of the standard model (SM) is complete and it reigns supreme as an effective theory for weak scale physics. However, in spite of its astonishing success, the anomalous magnetic moment of the muon, namely $a_\mu \equiv (g-2)/2$, remains an enigma. When compared to the SM estimate \cite{gm2_hagiwara}, the latest experimental result \cite{gm2_exp} stands as \begin{equation} \Delta a_\mu \equiv (a_\mu)_{\rm exp} - (a_\mu)_{\rm SM} = (26.1 \pm 8.1) \times 10^{-10} \, . \label{eq:gm2} \end{equation} The deviation is above $3\sigma$ level (see also ~\cite{Davier:2010nc}), and it can be resolved if we invoke new physics at a scale $m_{\rm NP} = \mathcal{O}(100)$ GeV, which follows from $(\Delta a_\mu)_{\rm NP} \sim (g^2/16\pi^2) (m_{\mu}^2/m_{\rm NP}^2) = 20.7 \times 10^{-10} (120\,{\rm GeV}/m_{\rm NP})^2 (g/0.65)^2$, where $g$ is a coupling relevant to the new physics. In the minimal supersymmetric standard model (MSSM), a resolution of this deviation requires light superparticles, namely the smuons and chargino/neutralinos of $\mathcal{O}(100)$~GeV, which propagate in the loop. With $\tan\beta \equiv \left<H_u\right>/\left<H_d\right> \sim 10$, the size of $\Delta a_\mu$ can be as large as $\mathcal{O}(10^{-9})$ \cite{gm2susy}. On the other hand, the observed Higgs boson mass, $m_h \sim 125$ GeV, demands rather large radiative corrections, which are enhanced by heavy stops weighing $\mathcal{O}(10)$ TeV or substantial left-right mixing \cite{OYY}. Gauge mediated supersymmetry breaking (GMSB) models \cite{GMSB} start with an advantage in this context that at the supersymmetry breaking scale itself the squarks/guino are heavier than sleptons/gauginos, i.e. the splitting is in the right direction. However, in minimal conventional GMSB, which employs a ${\bf 5}$ and a ${\bar {\bf 5}}$ of ${\rm SU(5)}_{\rm GUT}$ as messengers, the heavy stop pulls up the slepton and weak gaugino soft masses to several hundred GeV to a TeV which are too high to explain the muon $(g-2)$. In a previous paper \cite{our_previous} (see also \cite{IMYY}), we proposed a GMSB model that naturally yielded light uncolored and heavy colored superpartners. To accomplish this, we employed weak SU(2) triplet and color SU(3) octet messenger multiplets instead of using the conventional SU(5) 5-plets. Even with these incomplete SU(5) multiplets, gauge couplings still unify, though at the string scale $M_{\rm str} \sim 10^{17}$ GeV which is somewhat higher than the grand unification theory (GUT) scale $M_G \sim 10^{16}$ GeV. In addition to satisfying the 125 GeV Higgs boson mass, we could explain the muon $(g-2)$ at 2$\sigma$ level, with the agreement getting better upon the addition of SU(5) 5-plet messengers. In the most favorable region the stau would be the next to lightest supersymmetric particle (NLSP) being lighter than the bino. For satisfying the cosmological and accelerators constraints, mild $R$-parity violation (RPV) had to be invoked which facilitated prompt stau decay. Recently, radiative corrections to the Higgs mass have been computed at 3-loop level \cite{H3m} (see also \cite{higgs_3loop}), and it has been observed that $m_h \sim 125$ GeV is consistent with stop mass as light as $3-5$ TeV even for minimal left-right scalar mixing \cite{Feng_3loop}. We show in the present paper that this reduction of the stop mass allows us to present an improved scenario which is more comfortable with experimental data. Through the discussion that follows, we show that a GMSB model with weak SU(2) triplet, color SU(3) octet and SU(5) 5-plet messengers not only satisfies $m_h \sim 125$ GeV, but also can explain the muon $(g-2)$ at 1$\sigma$ level. Gauge coupling unification is indeed nontrivially maintained. No less importantly, we can satisfy the (cosmological) gravitino problem and the LHC constraints without any need of introducing RPV operators (For an alternative approach, where sparticles of 1st/2nd generation are light and of the 3rd generation are heavy, see \cite{IYY}). \section{A practical GMSB model} We employ three types of messenger fields: $\Phi_5$($\Phi_{\bar 5}$) transforming as ${\bf 5}$(${\bar{\bf 5}}$) of ${\rm SU(5)}_{\rm GUT}$, weak SU(2) triplet $\Sigma_3 ({\bf 1, 3, Y=0})$, and color SU(3) octet $\Sigma_8 ({\bf 8, 1, Y=0})$. The superpotential can be written as \begin{equation} W = (M_5 + \lambda_5 F \theta^2) \Phi_5 \Phi_{\bar 5} + (M_8 + \lambda_8 F \theta^2) {\rm Tr} (\Sigma_8^2 ) + (M_3 + \lambda_3 F \theta^2) {\rm Tr} (\Sigma_3^2 ), \end{equation} where $F$ characterizes the supersymmetry breaking scale. The leading contributions to the gaugino and sfermion masses arising from the messenger loops are given by \begin{eqnarray} m_{\tilde{B}} &\simeq& \frac{\alpha_1}{4\pi} \Lambda_5, \ m_{\tilde{W}} \simeq \frac{\alpha_2}{4\pi} (2 \Lambda_3 + \Lambda_5), \ m_{\tilde{g}} \simeq \frac{\alpha_3}{4\pi} (3 \Lambda_8 + \Lambda_5) \, ; \nonumber \\ m_{\tilde{Q}}^2 &\simeq& \frac{1}{8\pi^2} \left[ \frac{4}{3} \alpha_3^2 \left(3 \Lambda_8^2 + \Lambda_5^2 \right) + \frac{3}{4} \alpha_2^2 ( 2 \Lambda_3^2 + \Lambda_5^2) + \frac{1}{60} \alpha_1^2 \Lambda_5^2 \right] \, , \nonumber \\ m_{\tilde{\bar U}}^2 &\simeq& \frac{1}{8\pi^2} \left[ \frac{4}{3} \alpha_3^2 (3 \Lambda_8^2 + \Lambda_5^2) + \frac{4}{15} \alpha_1^2 \Lambda_5^2 \right] \, , \nonumber \\ m_{\tilde{\bar D}}^2 &\simeq& \frac{1}{8\pi^2} \left[ \frac{4}{3} \alpha_3^2 (3 \Lambda_8^2 + \Lambda_5^2) + \frac{1}{15} \alpha_1^2 \Lambda_5^2 \right] \, , \\ m_{\tilde{L}}^2&=&m_{H_u}^2=m_{H_d}^2 \simeq \frac{1}{8\pi^2} \left[ \frac{3}{4} \alpha_2^2 (2 \Lambda_3^2 + \Lambda_5^2) + \frac{3}{20} \alpha_1^2 \Lambda_5^2 \right] \, , \nonumber \\ m_{\tilde{\bar E}}^2 &\simeq& \frac{1}{8\pi^2} \left[ \frac{3}{5} \alpha_1^2 \Lambda_5^2 \right] \, ; \nonumber \label{eq:squark} \end{eqnarray} where \begin{equation} \alpha_i \equiv \frac{g_i^2}{4\pi}, ~~ \Lambda_8 \equiv \frac{\lambda_8 F}{M_8}, ~~ \Lambda_3 \equiv \frac{\lambda_3 F}{M_3}, ~~ \Lambda_5 \equiv \frac{\lambda_5 F}{M_5} \, . \end{equation} \begin{figure}[t] \begin{center} \includegraphics[scale=0.9]{gut_2e12.eps} ~~~~ \includegraphics[scale=0.9]{gut_5e13.eps} \caption{\small {\sf Two-loop evolution of the gauge couplings with $\Sigma_3$ and $\Sigma_8$ as a function of the renormalization scale (GeV). We take $\alpha_s(M_Z)=0.1184$ and the supersymmetry breaking scale $m_{\rm SUSY} \sim m_{\rm stop} \simeq 3.6$ TeV.}} \label{fig:gut} \end{center} \end{figure} \paragraph{Gauge coupling unification:} Even with incomplete GUT multiplets, i.e. with $\Sigma_3$ and $\Sigma_8$ only as messengers, gauge couplings do unify with $M_{\rm mess} \equiv M_8 \sim M_3$ \cite{gut_adjoint}. Solving the coupling evolution equations explicitly, as done in our previous paper \cite{our_previous}, it follows that lower the messenger scale $M_{\rm mess}$ below the GUT scale $M_G$, higher the actual unification scale $M_{\rm str}$ above $M_G$. Pushing $M_{\rm str}$ closer to the Planck scale $M_{\rm Pl} \simeq 2.4 \times 10^{18}$ GeV sets a lower limit $M_8 \gtrsim 10^{11}$ GeV. The presence of the 5-plets does not change the evolution slopes, so the above discussion holds in the present scenario. In Fig.~\ref{fig:gut}, we exhibit the evolutions of the gauge couplings with $M_8$ and $M_3$ around $(10^{11}-10^{13})$ GeV scale. The calculation has been performed using the renormalization group equations (RGE) at 2-loop level \cite{rge_2loop}. The gauge couplings are unified at $M_{\rm str} \sim 10^{18}$ GeV. This allows us to take $M_8$ closer to its lower limit. Since $F/M_8$ sets the squark mass, and $F/M_{\rm Pl}$ determines the gravitino mass, a lower value of $M_8$ implies a lighter gravitino, which helps us solve the cosmological problem (see later). \begin{figure}[htbp] \begin{center} \includegraphics[scale=0.9]{mhiggs1e11.eps} \caption{\small {\sf Contours of the Higgs boson mass including $\mathcal{O}(y_t \alpha_s^2)$ corrections. Here, $m_t=173.2$ GeV and $\alpha_s(M_Z)=0.1184$. In the gray region, the stau mass is below the LEP2 limit of 90 GeV.}} \label{fig:mhiggs} \end{center} \end{figure} \paragraph{The Higgs boson mass:} The observed $m_h \sim 125$ GeV sets the scale of the stop mass, which in turn fixes $\Lambda_8$. In Fig.~\ref{fig:mhiggs}, we show the contours of the Higgs boson mass in the $\Lambda_8-(\Lambda_5/\Lambda_8)$ plane. The Higgs boson mass has been evaluated using H3m-v1.2 package~\cite{H3m}, which includes $\mathcal{O}(y_t \alpha_s^2)$ corrections, where $y_t$ is the top quark Yukawa coupling. We take note of the fact that $m_h\sim$ 125 GeV can be explained with $\Lambda_8 \simeq 200-300$ TeV. The corresponding stop masses are in the $(3.6-5.1)$ TeV range. \paragraph{Muon $g-2$:} A rough estimate of the Higgsino mixing parameter is \begin{equation} \mu^2 \sim (-m_{H_u}^2) \sim \frac{3}{4\pi^2} y_t^2 (m_{\rm stop}^2) \frac{M_{\rm mess}}{m_{\rm stop}}, \end{equation} where $m_{\rm stop}\equiv (m_{\tilde{Q}_3} m_{\tilde{\bar{U}}_3})^{1/2}$ is the (geometric) average stop mass scale. For illustration, we have neglected the soft mass of $H_u$ generated at the messenger scale, and considered only the radiative mass generation. Putting $m_{\rm stop}=3$ TeV and $M_{\rm mess}=10^{11}$ GeV, one obtains $\mu \sim 2.7$ TeV. The value of $\mu$ is still too large to make the chargino induced contributions to $(g-2)$ numerically relevant. This contribution is dominated by the bino-slepton loop, which is given by \begin{equation} (\Delta {a_{\mu}})_{\rm SUSY} \simeq \frac{3}{5}\frac{g_1^2}{8\pi^2}\frac{m_\mu^2 \mu\tan\beta}{M_1^3} F_b\left(\frac{m_{\tilde{L}}^2}{M_1^2}, \frac{m_{\tilde{\bar E}}^2}{M_1^2}\right), \end{equation} where $m_{\mu}$ is the muon mass. The contribution is proportional to the left-right smuon mixing term which contains the $(\mu \tan\beta)$ factor. The loop function $F_b$ is defined and explicitly displayed in Ref.~\cite{hagi_gm2_susy} (for a rough guide, $F_b(1,1) = 1/6$). In order to explain the muon $g-2$, the bino has to be necessarily light as $\mathcal{O}(100)$ GeV, and the smuon not much heavier. \begin{figure}[t] \begin{center} \includegraphics[scale=0.9]{200e3_1e11.eps} ~~~~ \includegraphics[scale=0.9]{300e3_1e11.eps} \caption{\small {\sf In the orange (yellow) region, the muon $g-2$ is explained at $1(2)\sigma$ level. The neutralino (stau) is NLSP above (below) the blue solid lines. In the gray region, the stau mass is smaller than 90~GeV. The contours of the chargino mass (red solid lines) and the soft mass of the left-handed sleptons (green dashed lines) are shown in units of GeV.}} \label{fig:gm2_SUSY} \end{center} \end{figure} In Fig.~\ref{fig:gm2_SUSY} we display to what extent we can explain the muon $(g-2)$. We take $M_{\rm mess}=M_8=M_3$ for simplicity. The supersymmetric mass spectrum as well as the RGE running of various parameters have been performed using {\tt SuSpect} \cite{suspect}. The supersymmetric contributions to the muon $g-2$ has been evaluated by {\tt FeynHiggs2.9.5} \cite{FeynHiggs}. To include the threshold corrections to slepton masses from the Higgsino and heavy Higgs boson \cite{BPMZ, tata}, we have modified the {\tt SuSpect} package appropriately. The contours of different chargino masses have been shown by red solid lines. In the orange (yellow) region, the muon $g-2$ is explained at 1$\sigma$ (2$\sigma$) level, at the same time keeping consistency with $m_h \sim 125$ GeV. In the region above the blue solid line, the neutralino (dominantly the bino, since $\mu$ is large) is the NLSP, while in the region below the line, the stau is the NLSP. When the stau is the NLSP, even though it eventually decays to gravitino, it is stable inside the detector. In this case, the stau mass of less than 340 GeV is excluded by the LHC data \cite{stau_stable}. But we need the stau to weigh in the (100-250) GeV ballpark so that the smuon acquires an appropriate mass to explain the muon $g-2$ at (1-2)$\sigma$ level. Hence, viable regions are only above the blue solid line, where the lightest neutralino (dominantly, the bino) is the NLSP. Because of the $\Lambda_5$ induced contributions in Eq.~(3), the bino mass is generated in a way which is completely uncorrelated to the gravitino mass generation. The gravitino mass is estimated as \begin{equation} m_{3/2} \simeq 0.01\,{\rm GeV} \left(\frac{\Lambda_8}{200 {\rm TeV}}\right) \left(\frac{(\Lambda_3/\Lambda_8)}{0.2}\right) \left(\frac{M_8}{10^{11} {\rm GeV}}\right) \left(\frac{(M_3/M_8)}{10}\right). \end{equation} With this gravitino mass of $\mathcal{O}(10^{-2})$ GeV, the life-time of the neutralino is $(1-10)$ second, giving a constraint on the primordial neutralino abundance. However, this abundance is very small, as a result of which the successful prediction of the big bang nucleosynthesis (BBN) is maintained \cite{gravitino_problem}, thus avoiding the gravitino problem. In Table~\ref{table:mass}, we have presented two sets of reference points, displaying the mass spectra and the prediction for $(g-2)$, that pass all constraints. In this context, two types of constraints deserve special mention: \\ \noindent ($i$)~ When the left-right stau mixing term proportional to ($m_{\tau} \mu \tan\beta$) is large, the charge breaking global minimum can appear. The life-time of the electroweak vacuum restricts the size of the $\mu\tan\beta$, which depends of course on the stau soft mass parameters \cite{stau_vac, stau_Hisano}. However, this constraint is not very decisive in our case. In fact, the LEP bound on the stau mass is stronger in the relevant region of the parameter space \cite{PDG}. \noindent ($ii$)~ LHC constraints on electroweak gauginos/sleptons also restrict the relevant parameter space. Searches for three leptons plus missing energy put a constraint on the wino mass~\cite{3lepton_LHC}. In the region consistent with the muon $g-2$ at $1\sigma$ level, the left handed sleptons are some what heavier than the wino. In this case, the final state leptons are the taus rather electrons/muons, giving the constraint $m_{\chi_1^\pm} \simeq m_{\chi_2^0} \gtrsim (300-350)$ GeV \cite{3lepton_LHC, three_taus}. Note that in some regions of the parameter space, the wino and the left-handed sleptons are nearly degenerate in mass. These regions are difficult to be constrained. Besides, separate (but, not so tight) constraints exist on the left-handed sleptons, namely, $m_{\tilde{\ell}_L} \gtrsim 300$ GeV \cite{sleptons_LHC}. The restrictions on the right-handed sleptons are, however, much less stringent. \begin{table}[t!] \begin{center} \begin{tabular}{ c | c } $\Lambda_3/\Lambda_8$ & 0.17 \\ $\Lambda_5/\Lambda_8$ & 0.41 \\ $\Lambda_8$ & 200 TeV \\ $M_{\rm mess}$ & $10^{11}$ GeV \\ $\tan\beta$ & 10 \\ \hline \hline $\mu$ & $2.4$\,TeV\\ $m_{\rm stop}$ & 3.6\,TeV \\ $\delta a_\mu$ & 20.3 $\times10^{-10}$ \\ \hline $m_{\rm gluino}$ & 4.4 TeV \\ $m_{\rm squark} $ & 4.1 TeV \\ $m_{\tilde{e}_L} (m_{\tilde{\mu}_L})$ & 379 GeV\\ $m_{\tilde{e}_R} (m_{\tilde{\mu}_R})$ & 181 GeV \\ $m_{\tilde{\tau}_1}$ & 123 GeV \\ $m_{\chi_1^0}$ & 100 GeV \\ $m_{\chi_1^{\pm}}/m_{\chi_2^0}$ & 375 GeV \\ \end{tabular} \begin{tabular}{ c | c } $\Lambda_3/\Lambda_8$ & 0.11 \\ $\Lambda_5/\Lambda_8$ & 0.35 \\ $\Lambda_8$ & 300 TeV \\ $M_{\rm mess}$ & $10^{11}$ GeV \\ $\tan\beta$ & 10 \\ \hline \hline $\mu$ & $3.5$\,TeV\\ $m_{\rm stop}$ & 5.1\,TeV \\ $\delta a_\mu$ & 18.6 $\times10^{-10}$ \\ \hline $m_{\rm gluino}$ & 6.3 TeV \\ $m_{\rm squark} $ & 5.8 TeV \\ $m_{\tilde{e}_L} (m_{\tilde{\mu}_L})$ & 425 GeV\\ $m_{\tilde{e}_R} (m_{\tilde{\mu}_R})$ & 218 GeV \\ $m_{\tilde{\tau}_1}$ & 133 GeV \\ $m_{\chi_1^0}$ & 128 GeV \\ $m_{\chi_1^{\pm}}/m_{\chi_2^0}$ & 411 GeV \\ \end{tabular} \caption{Mass spectra and $(\Delta {a_\mu})_{\rm SUSY}$ for two reference points.} \label{table:mass} \end{center} \end{table} \section{Conclusions} Reconciling the observed Higgs boson mass and the measurement of the muon $(g-2)$ poses a big challenge to supersymmetric model building. In this paper we have presented a realistic GMSB model that can address both these issues satisfying all other constraints. From the model-building perspective, the situation has considerably improved since we constructed the scenario of Ref.~\cite{our_previous}. We summarize below the salient features behind this improvement. The recent 3-loop radiative corrections to the Higgs boson mass imply that the stop squark is perhaps not as heavy as order $\mathcal{O} (10)$ TeV. A stop mass of mere (3-5) TeV can explain the observed 125 GeV mass of the Higgs boson even for small stop mixing. A lighter stop means a relatively smaller value of $\mu$ (but still $\sim$ 3 TeV). This has two implications. First, the lightest neutralino weighing $\sim$ 100 GeV is bino dominated because $\mu$ is still quite large ($\sim$ 3 TeV). Second, the left-right mixing in the slepton sector, which is proportional to $\mu\tan\beta$, is relatively smaller because the value of $\mu$ has come down from 6 TeV to 3 TeV thanks to the smaller stop masses. Since a smaller left-right mixing implies a smaller splitting between the two slepton mass eigenvalues of the same flavor, for a wide region of the interesting parameter space the lightest neutralino can remain lighter than stau. Note that we could have generated light uncolored and heavy colored superpartners just with $\Sigma_3$ and $\Sigma_8$, still nontrivially satisfying gauge coupling unification. The key area where we really improved with respect to Ref.~\cite{our_previous}, thanks to the presently known three-loop corrections to the Higgs boson mass and our assumption of a somewhat {\em late} unification, is that we can now take bino light enough to explain the muon $(g-2)$ within 1$\sigma$, satisfying at the same time the BBN constraint and LHC data {\em without} introducing RPV operators. The interesting region of parameter space can be probed at the 14 TeV LHC, and precision measurements of the lighter stau can be performed at the future International Linear Collider (ILC). \section*{Acknowledgements} The work of N.Y. is supported in part by JSPS Research Fellowships for Young Scientists. This work is also supported by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. \small
\section{Introduction}\SectionName{intro} Dimensionality reduction is a ubiquitous tool across a wide array of disciplines: machine learning \cite{WDLSA09}, high-dimensional computational geometry \cite{Indyk01}, privacy \cite{BlockiBDS12}, compressed sensing \cite{CT05}, spectral graph theory \cite{SpielmanS11}, interior point methods for linear programming \cite{LeeS13a}, numerical linear algebra \cite{Sarlos06}, computational learning theory \cite{BalcanB05,BalcanBV06}, manifold learning \cite{BHW07,Clarkson08}, motif-finding in computational biology \cite{BuhlerT02}, astronomy \cite{ContrerasM12}, and several others. Across all these disciplines one is typically faced with data that is not only massive, but each data item itself is represented as a very high-dimensional vector. For example, when learning spam classifiers a data point is an email, and it is represented as a high-dimensional vector indexed by dictionary words \cite{WDLSA09}. In astronomy a data point could be a star, represented as a vector of light intensities measured over various points sampled in time \cite{KovacsZM02,VanderburgJ14}. Dimensionality reduction techniques in such applications provide the following benefits: \begin{itemize} \item Smaller storage consumption. \item Speedup during data analysis. \item Cheaper signal acquisition. \item Cheaper transmission of data across computing clusters. \end{itemize} The technical guarantees required from a dimensionality reduction routine are application-specific, but typically such methods must reduce dimension while still preserving point geometry, e.g.\ inter-point distances and angles. That is, one has some point set $X\subset\mathbb{R}^n$ with $n$ very large, and we would like a dimensionality-reducing map $f:X\rightarrow\mathbb{R}^m$, $m\ll n$, such that \begin{equation} \forall x,y\in X,\ (1-\varepsilon)\|x-y\| \le \|f(x)-f(y)\| \le (1+\varepsilon)\|x-y\| \EquationName{jlf} \end{equation} for some norm $\|\cdot\|$. Note also that for unit vectors $x,y$, $\cos(\angle (x,y)) = (1/2)(\|x\|_2^2 + \|y\|_2^2 - \|x-y\|_2^2)$, and thus $f$ also preserves angles with additive error if it preserves Euclidean norms of points in $X\cup(X-X)$. A powerful tool for achieving \Equation{jlf}, used in nearly all the applications cited above, is the {\em Johnson-Lindenstrauss (JL) lemma} \cite{JL84}. \begin{theorem}[JL lemma] For any subset $X$ of Euclidean space and $0<\varepsilon<1/2$, there exists $f:X\rightarrow\ell_2^m$ with $m = O(\varepsilon^{-2}\log |X|)$ providing \Equation{jlf} for $\|\cdot\| = \|\cdot\|_2$. \end{theorem} This bound on $m$ is nearly tight: for any $n\ge 1$ Alon exhibited a point set $X\subset\ell_2^n$, $|X|=n+1$, such that any such JL map $f$ must have $m = \Omega(\varepsilon^{-2}(\log n)/\log(1/\varepsilon))$ \cite{Alon03}. In fact, all known proofs of the JL lemma provide linear $f$, and the JL lemma is tight up to a constant factor in $m$ when $f$ must be linear \cite{LarsenN14}. Unfortunately, for actual applications such {\em worst-case} understanding is unsatisfying. Rather we could ask: if given a distortion parameter $\varepsilon$ and point set $X$ as input (or a succinct description of it if $X$ is large or even infinite, as in some applications), what is the best target dimension $m = m(X,\varepsilon)$ such that a JL map exists for $X$ with this particular $\varepsilon$? That is, in practice we are more interested in moving beyond worst case analysis and being as efficient as possible {\em for our particular data} $X$. Unfortunately the previous question seems fairly difficult. For the related question of computing the optimal distortion for embedding $X$ into a line (i.e.\ $m=1$), it is computationally hard to approximate the optimal distortion even up to a multiplicative factor polynomial in $|X|$ \cite{BadoiuCIS05}. In practice, however, typically $f$ cannot be chosen arbitrarily as a function of $X$ anyway. For example, when employing certain learning algorithms such as stochastic gradient descent on dimensionality-reduced data, it is at least required that $f$ is differentiable on $\mathbb{R}^n$ (where $X\subset \mathbb{R}^n$) \cite{WDLSA09}. For several applications it is also crucial that $f$ is linear, e.g.\ in numerical linear algebra \cite{Sarlos06} and compressed sensing \cite{CT05,Donoho04}. In one-pass streaming applications \cite{ClarksonW09} and data structural problems such as nearest neighbor search \cite{HarPeledIM12}, it is further required that $f$ is chosen randomly without knowing $X$. For any particular $X$, a random $f$ drawn from some distribution must satisfy the JL guarantee with good probability. In streaming applications this is because $X$ is not fully known up front, but is gradually observed in a stream. In data structure applications this is because $f$ must preserve distances to some future query points, which are not known at the time the data structure is constructed. Due to the considerations discussed, in practice typically $f$ is chosen as a random linear map drawn from some distribution with a small number of parameters (in some cases simply the parameter $m$). For example, popular choices of $f$ include a random matrix with independent Gaussian \cite{HarPeledIM12} or Rademacher \cite{Achlioptas03} entries. While worst case bounds inform us how to set parameters to obtain the JL guarantee for worst case $X$, we typically can obtain better parameters by exploiting prior knowledge about $X$. Henceforth we only discuss linear $f$, so we write $f(x) = \Phi x$ for $\Phi\in\mathbb{R}^{m\times n}$. Furthermore by linearity, rather than preserving Euclidean distances in $X$ it is equivalent to discuss preserving norms of all vectors in $T = \{(x-y)/\|x-y\|_2 : x,y\in X\}\subset S^{n-1}$, the set of all normalized difference vectors amongst points in $X$. Thus up to changing $\varepsilon$ by roughly a factor of $2$, \Equation{jlf} is equivalent to \begin{equation} \sup_{x\in T} \Big|\|\Phi x\|^2 - 1\Big| < \varepsilon . \EquationName{jl-condition} \end{equation} Furthermore, since we consider $\Phi$ chosen at random, we more specifically want \begin{equation} \mathbb{E}_\Phi \sup_{x\in T} \Big|\|\Phi x\|^2 - 1\Big| < \varepsilon .\EquationName{rjl-condition} \end{equation} Instance-wise understanding for achieving \Equation{rjl-condition} was first provided by Gordon \cite{Gordon88}, who proved that a random $\Phi$ with i.i.d.\ Gaussian entries satisfies \Equation{rjl-condition} as long as $m \gtrsim (g^2(T)+1)/\varepsilon^2$, where we write $A\gtrsim B$ if $A \ge CB$ for a universal constant $C>0$. Denoting by $g$ a standard $n$-dimensional Gaussian vector, the parameter $g(T)$ is defined as the {\em Gaussian mean width} $$g(T)\mathbin{\stackrel{\rm def}{=}} \mathbb{E}_g\sup_{x\in T}\langle g,x\rangle .$$ One can think of $g(T)$ as describing the $\ell_2$-geometric complexity of $T$. It is always true that $g^2(T) \lesssim \log |T|$, and thus Gordon's theorem implies the JL lemma. In fact for all $T$ we know from applications, such as for the restricted isometry property from compressed sensing \cite{CT05} or subspace embeddings from numerical linear algebra \cite{Sarlos06}, the best bound on $m$ is a corollary of Gordon's theorem. Later works extended Gordon's theorem to other distributions for $\Phi$, such as $\Phi$ having independent subgaussian entries (e.g.\ Rademachers) \cite{KM05,MPT07,Dirksen14}. Although Gordon's theorem gives a good understanding for $m$ in most scenarios, it suffers from the fact that it analyzes a {\em dense} random $\Phi$, which means that performing the dimensionality reduction $x\mapsto \Phi x$ is dense matrix-vector multiplication, and is thus slow. For example, in some numerical linear algebra applications (such as least squares regression \cite{Sarlos06}), multiplying a dense unstructured $\Phi$ with the input turns out to be slower than obtaining the solution of the original, high-dimensional problem! In compressed sensing, certain iterative recovery algorithms such as CoSamp \cite{NT09} and Iterative Hard Thresholding \cite{BD08} involve repeated multiplications by $\Phi$ and $\Phi^*$, the conjugate transpose of $\Phi$, and thus $\Phi$ supporting fast matrix-vector multiply are desirable in such applications as well. The first work to provide $\Phi$ with small $m$ supporting faster multiplication is the Fast Johnson-Lindenstrauss Transform (FJLT) of \cite{AC09} for finite $T$. The value of $m$ was still $O(\varepsilon^{-2}\log|T|)$, with the time to multiply $\Phi x$ being $O(n\log n + m^3)$. \cite{AL09} later gave an improved construction with time $O(n\log n + m^{2+\gamma})$ for any small constant $\gamma>0$. Most recently several works gave nearly linear embedding time in $n$, independent of $m$, at the expense of increasing $m$ by a $(\log n)^c$ factor \cite{AL13,KW11,NPW14}. In all these works $\Phi$ is the product of some number of very sparse matrices and Fourier matrices, with the speed coming from the Fast Fourier Transform (FFT) \cite{CooleyT65}. It is also known that this FFT-based approach can be used to obtain fast RIP matrices for compressed sensing \cite{CT06,RV08,CGV13} and fast oblivious subspace embeddings for numerical linear algebra applications \cite{Sarlos06} (see also \cite{Tropp11,LDFU13} for refined analyses in the latter case). Another line of work, initiated in \cite{Achlioptas03} and greatly advanced in \cite{DKS10}, sought fast embedding time by making $\Phi$ sparse. If $\Phi$ is drawn from a distribution over matrices having at most $s$ non-zeroes per column, then $\Phi x$ can be computed in time $s\cdot \|x\|_0$. After some initial improvements \cite{KN10,BOR10}, the best known achievable value of $s$ to date for the JL lemma while still maintaining $m \lesssim \varepsilon^{-2}\log |T|$ is the sparse Johnson-Lindenstrauss Transform (SJLT) of \cite{KN14}, achieving $s \lesssim \varepsilon^{-1}\log |T| \lesssim \varepsilon m$. Furthermore, an example of a set $T$ exists which requires this bound on $s$ up to $O(\log(1/\varepsilon))$ for any linear JL map \cite{NN13a}. Note however that, again, this is an understanding of the {\em worst-case} parameter settings over all $T$. In summary, while Gordon's theorem gives us a good understanding of instance-wise bounds on $T$ for achieving good dimensionality reduction, it only does so for dense, slow $\Phi$. Meanwhile, our understanding for efficient $\Phi$, such as the SJLT with small $s$, has not moved beyond the worst case. In some very specific examples of $T$ we do have good bounds for settings of $s, m$ that suffice, such as $T$ the unit norm vectors in a $d$-dimensional subspace \cite{CW13,MM13,NN13b}, or all elements of $T$ having small $\ell_\infty$ norm \cite{Matousek08,DKS10,KN10,BOR10}. However, our understanding for general $T$ is non-existent. This brings us to the main question addressed in this work, where $S^{n-1}$ denotes the $\ell_2$-unit sphere $\{x\in\mathbb{R}^n : \|x\|_2 = 1\}$. \begin{question}\QuestionName{main} \textup{Let $T\subseteq S^{n-1}$ and $\Phi$ be the SJLT. What relationship must $s,m$ satisfy, in terms of the geometry of $T$, to ensure \Eqsub{rjl-condition}?} \end{question} We also note that while the FFT-based and sparse $\Phi$ approaches seem orthogonal at first glance, the two are actually connected, as pointed out before \cite{AC09,Matousek08,NPW14}. The FJLT sets $\Phi = SP$ where $P$ is some random preconditioning matrix that makes $T$ ``nice'' with high probability, and $S$ is a random sparse matrix. We point out that although the SJLT is not the same as the matrix $S$ typically used in the FFT-based literature, one could replace $S$ with the SJLT and hope for similar algorithmic outcome if $s$ is small: nearly linear embedding time. The answer to the analog of \Question{main} for a standard Gaussian matrix depends only on the $\ell_2$-metric structure of $T$. Indeed, since both $\ell_2$-distances and Gaussian matrices are invariant under orthogonal transformations, so is \Eqsub{rjl-condition} in this case. This is reflected in Gordon's theorem, where the embedding dimension $m$ is governed by the Gaussian width, which is invariant under orthogonal transformations. We stress that in sharp contrast, a resolution of \Question{main} cannot solely depend on the $\ell_2$-metric structure of $T$. Indeed, we require that $\Phi$ be sparse {\em in a particular basis} and is therefore not invariant under orthogonal transformations. Thus an answer to \Question{main} must be more nuanced (see our main theorem, \Theorem{main}). \medskip \paragraph{\textbf{Our Main Contribution:}} We provide a general theorem which answers \Question{main}. Specifically, for every $T\subseteq S^{n-1}$ analyzed in previous work that we apply our general theorem to here, we either (1) qualitatively recover or improve the previous best known result, or (2) prove the first non-trivial result for dimensionality reduction with sparse $\Phi$. We say ``qualitatively'' since applying our general theorem to these applications loses a factor of $\log^c(n/\varepsilon)$ in $m$ and $\log^c(n/\varepsilon)/\varepsilon$ in $s$. In particular for (2), our work is the first to imply that non-trivially sparse dimensionality reducing linear maps can be applied for gain in model-based compressed sensing \cite{BCDH10}, manifold learning \cite{TSL00,DonohoG03}, and constrained least squares problems such as the popular Lasso \cite{Tibshirani96}. \medskip Specifically, we prove the following theorem. \begin{theorem}[Main Theorem]\TheoremName{main} Let $T\subset S^{n-1}$ and $\Phi$ be an SJLT with column sparsity $s$. Define the complexity parameter $$\kappa(T)\mathbin{\stackrel{\rm def}{=}} \kappa_{s,m}(T) = \max_{q\le \frac ms \log s} \Big\{\frac 1{\sqrt{qs}} \Big(\mathbb{E}_\eta \Big(\mathbb{E}_g \sup_{x\in T} |\sum_{j=1}^n \eta_j g_j x_j|\Big)^q\Big)^{1/q}\Big\},$$ where $(g_j)$ are i.i.d.\ standard Gaussian and $(\eta_j)$ i.i.d.\ Bernoulli with mean $qs/(m\log s)$. If \begin{align*} m &\gtrsim (\log m)^3(\log n)^5 \cdot\frac{(g^2(T) + 1)}{\varepsilon^2}\\ s &\gtrsim (\log m)^6 (\log n)^4 \cdot\frac{1}{\varepsilon^2} . \end{align*} Then \Eqsub{rjl-condition} holds as long as $s,m$ furthermore satisfy the condition \begin{equation} (\log m)^2(\log n)^{5/2}\kappa(T) < \varepsilon \EquationName{kappa-suffices} . \end{equation} \end{theorem} The complexity parameter $\kappa(T)$ may seem daunting at first, but \Section{applications} shows it can be controlled quite easily for all the $T$ we have come across in applications. \subsection{Applications} Here we describe various $T$ and their importance in certain applications. Then we state the consequences that arise from our theorem. In order to highlight the qualitative understanding arising from our work, we introduce the notation $A\mathbin{\substack{<\\ *}}\xspace B$ if $A \le B \cdot \varepsilon^{-2}(\log n)^c$. A summary of our bounds is in \Figure{fig}. \medskip \paragraph{\textbf{Finite $|T|$:}} This is the setting $|T|<\infty$, for which the SJLT satisfies \Equation{rjl-condition} with $s\lesssim \varepsilon^{-1}\log|T|$, $m\lesssim \varepsilon^{-2}\log|T|$ \cite{KN14}. If also $T\subset B_{\ell_\infty^n}(\alpha)$, i.e.\ $\|x\|_\infty \le \alpha$ for all $x\in T$, \cite{Matousek08} showed it is possible to achieve $m \lesssim \varepsilon^{-2}\log|T|$ with a $\Phi$ that has an {\em expected} $O(\varepsilon^{-2}(\alpha\log|T|)^2)$ non-zeroes per column. Our theorem implies $s,m\mathbin{\substack{<\\ *}}\xspace \log|T|$ suffices in general, and $s\mathbin{\substack{<\\ *}}\xspace 1 + (\alpha \log |T|)^2, m \mathbin{\substack{<\\ *}}\xspace \log|T|$ in the latter case, qualitatively matching the above. \medskip \paragraph{\textbf{Linear subspace:}} Here $T = \{x \in E : \|x\|_2 = 1\}$ for some $d$-dimensional linear subspace $E\subset \mathbb{R}^n$, for which achieving \Equation{rjl-condition} with $m\lesssim d/\varepsilon^2$ is possible \cite{AHK06,CW13}. A distribution satisfying \Equation{rjl-condition} for any $d$-dimensional subspace $E$ is known as an {\em oblivious subspace embedding (OSE)}. The paper \cite{Sarlos06} pioneered the use of OSE's for speeding up approximate algorithms for numerical linear algebra problems such as low-rank approximation and least-squares regression. More applications have since been found to approximating leverage scores \cite{DMMW12}, $k$-means clustering \cite{BZMD11,CohenEMMP14}, canonical correlation analysis \cite{ABTZ13}, support vector machines \cite{PBMD13}, $\ell_p$-regression \cite{CDM+13,WZ13}, ridge regression \cite{LDFU13}, streaming approximation of eigenvalues \cite{AN13}, and speeding up interior point methods for linear programming \cite{LeeS13a}. In many of these applications there is some input $A\in\mathbb{R}^{n\times d}$, $n\gg d$, and the subspace $E$ is for example the column space of $A$. Often an exact solution requires computing the singular value decomposition (SVD) of $A$, but using OSE's the running time is reduced to that for computing $\Phi A$, plus computing the SVD of the smaller matrix $\Phi A$. The work \cite{CW13} showed $s=1$ with small $m$ is sufficient, yielding algorithms for least squares regression and low-rank approximation whose running times are linear in the number of non-zero entries in $A$ for sufficiently lopsided rectangular matrices. Our theorem implies that $s \mathbin{\substack{<\\ *}}\xspace 1$ and $m \mathbin{\substack{<\\ *}}\xspace d$ suffices to satisfy \Equation{rjl-condition}, which is qualitatively correct. Furthermore, a subset of our techniques reveals that if the maximum incoherence $\max_{1\le i\le n} \|P_E e_i\|_2$ is at most $\mathop{poly}(\varepsilon/\log n)$, then $s=1$ suffices (Theorem~\ref{thm:JLsubspace}). This was not known in previous work. A random $d$-dimensional subspace has incoherence $\sqrt{d/n}$ w.h.p.\ for $d\gtrsim\log n$ by the JL lemma, and thus is very incoherent if $n\gg d$. \medskip \paragraph{\textbf{Closed convex cones:}} For $A\in\mathbb{R}^{n\times d}$, $b\in\mathbb{R}^n$, and $\mathcal{C}\subseteq \mathbb{R}^d$ a closed convex set, consider the constrained least squares problem of minimizing $\|Ax-b\|_2^2$ subject to $x\in\mathcal{C}$. A popular choice is the Lasso \cite{Tibshirani96}, in which the constraint set $\mathcal{C} = \{ x\in\mathbb{R}^d : \|x\|_1 \le R\}$ encourages sparsity of $x$. Let $x_*$ be an optimal solution, and let $T_{\mathcal{C}}(x)$ be the tangent cone of $\mathcal{C}$ at some point $x\in\mathbb{R}^d$ (see Appendix~\ref{sec:convex-tools} for a definition). Suppose we wish to accelerate approximately solving the constrained least squares problem by instead computing a minimizer $\hat{x}$ of $\|\Phi Ax - \Phi b\|_2^2$ subject to $x\in \mathcal{C}$. The work \cite{PiW14} showed that to guarantee $\|A\hat{x} - b\|_2^2 \le (1+\varepsilon)\|Ax_* - b\|_2^2$, it suffices that $\Phi$ satisfy two conditions, one of which is \Equation{jl-condition} for $T = A T_{\mathcal{C}}(x_*)\cap S^{n-1}$. The paper \cite{PiW14} then analyzed dense random matrices for sketching constrained least squares problems. For example, for the Lasso if we are promised that the optimal solution $x_*$ is $k$-sparse, \cite{PiW14} shows that it suffices to set $$ m\gtrsim \frac 1{\varepsilon^2}\max_{j=1,\ldots, d} \frac{\|A_j\|_2^2}{\sigma_{\min,k}^2}k\log d, $$ where $A_j$ is the $j$th column of $A$ and $\sigma_{\min,k}$ the smallest $\ell_1$-restricted eigenvalue of $A$: $$ \sigma_{\min,k} = \inf_{\|y\|_2=1, \ \|y\|_{1}\leq 2\sqrt{k}}\|Ay\|_2 . $$ Our work also applies to such $T$ (and we further show the SJLT with small $s, m$ satisfies the second condition required for approximate constrained least squares; see Theorem~\ref{thm:SJLmain}). For example for the Lasso, we show that again it suffices that $$ m\mathbin{\substack{>\\ *}}\xspace \max_{j=1,\ldots,d} \frac{\|A_j\|_2^2}{\sigma_{\min,k}^2}k\log d $$ but where we only require from $s$ that $$ s\mathbin{\substack{>\\ *}}\xspace \max_{\substack{1\le i\le n\\1\le j\le d}} \frac{A_{ij}^2}{\sigma_{\min,k}^2}k $$ That is, the sparsity of $\Phi$ need only depend on the largest entry in $A$ as opposed to the largest column norm in $A$. The former can be much smaller. \medskip \paragraph{\textbf{Unions of subspaces:}} Define $T = \cup_{\theta\in \Theta} E_\theta\cap S^{n-1}$, where $\Theta$ is some index set and each $E_\theta\subset\mathbb{R}^n$ is a $d$-dimensional linear subspace. A case of particular interest is when $\theta\in\Theta$ ranges over all $k$-subsets of $\{1,\ldots,n\}$, and $E_\theta$ is the subspace spanned by $\{e_j\}_{j\in \theta}$ (so $d=k$). Then $T$ is simply the set of all $k$-sparse unit vectors of unit Euclidean norm: $S_{n,k}\mathbin{\stackrel{\rm def}{=}} \{x \in \mathbb{R}^n : \|x\|_2 = 1, \|x\|_0 \le k\}$ for $\|\cdot\|_0$ denoting support size. $\Phi$ satisfying \Eqsub{jl-condition} is then referred to as having the {\em restricted isometry property (RIP) of order $k$} with restricted isometry constant $\varepsilon_k = \varepsilon$ \cite{CT05}. Such $\Phi$ are known to exist with $m\lesssim \varepsilon_k^{-2}k\log(n/k)$, and furthermore it is known that $\varepsilon_{2k} < \sqrt{2}-1$ implies that any (approximately) $k$-sparse $x\in\mathbb{R}^n$ can be (approximately) recovered from $\Phi x$ in polynomial time by solving a certain linear program \cite{CT05,Candes08}. Unfortunately it is known for $\varepsilon = \Theta(1)$ that {\em any} RIP matrix $\Phi$ with such small $m$ must have $s\gtrsim m$ \cite{NN13a}. Related is the case of vectors sparse in some other basis, i.e.\ $T = \{Dx \in \mathbb{R}^n : \|Dx\|_2 = 1,\|x\|_0 \le k\}$ for some so-called ``dictionary'' $D$ (i.e.\ the subspaces are acted on by $D$), or when $T$ only allows for some subset of all $\binom{n}{k}$ sparsity patterns in {\em model-based} compressed sensing \cite{BCDH10} (so that $|\Theta| < \binom{n}{k}$). Our main theorem also implies RIP matrices with $s, m \mathbin{\substack{<\\ *}}\xspace k\log(n/k)$. More generally when a dictionary $D$ is involved such that the subspaces $\mathrm{span}(\{D e_j\}_{j\in \theta})$ are all $\alpha$-incoherent (as defined above), the sparsity can be further improved to $s\mathbin{\substack{<\\ *}}\xspace 1 + (\alpha k\log(n/k))^2$. That is, to satisfy the RIP with dictionaries yielding incoherent subspaces, we can keep $m$ qualitatively the same while making $s$ much smaller. For the general problem of unions of $d$-dimensional subspaces, our theorem implies one can either set $m \mathbin{\substack{<\\ *}}\xspace d + \log|\Theta|, s \mathbin{\substack{<\\ *}}\xspace \log|\Theta|$ or $m \mathbin{\substack{<\\ *}}\xspace d + \log|\Theta|, s \mathbin{\substack{<\\ *}}\xspace 1 + (\alpha\log|\Theta|)^2$. Previous work required $m$ to depend on the {\em product} of $d$ and $(\log|\Theta|)^c$ instead of the {\em sum} \cite{NN13b}, including a nice recent improvement by Cohen \cite{Cohen14}, and is thus unsuitable for the application to the set of $k$-sparse vectors (RIP matrices with $\mathbin{\substack{<\\ *}}\xspace k^2$ rows are already attainable via simpler methods using incoherence; e.g.\ see \cite[Proposition 1]{BDFKK11}). Iterative recovery algorithms such as CoSamp can also be used in model-based sparse recovery \cite{BCDH10}, which again involves multiplications by $\Phi, \Phi^*$, and thus sparse $\Phi$ is relevant for faster recovery. Our theorem thus shows, for the first time, that the benefit of model-based sparse recovery is not just a smaller $m$, but rather that the measurement matrix $\Phi$ can be made much sparser if the model is simple (i.e.\ $|\Theta|$ is small). For example, in the {\em block-sparse model} one wants to (approximately) recover a signal $x\in\mathbb{R}^n$ based on $m$ linear measurements, where $x$ is (approximately) $k$-block-sparse. That is, the $n$ coordinates are partitioned into $n/b$ blocks of size $b$ each, and each block is either ``on'' (at least one coordinate in that block non-zero), or ``off'' (all non-zero). A $k$-block-sparse signal has at most $k/b$ blocks on (thus $\|x\|_0 \le k$). Thus $s\mathbin{\substack{<\\ *}}\xspace \log|\Theta| = \log (\binom{n/b}{k/b}) \lesssim (k/b)\log(n/k)$. Then as long as $b = \omega(\log(n/k))$, our results imply non-trivial column-sparsity $s \ll m$. Ours is the first result yielding non-trivial sparsity in a model-RIP $\Phi$ for any model with a number of measurements qualitatively matching the optimal bound (which is on the order of $m\lesssim k + (k/b)\log(n/k)$ \cite{ADR14}). We remark that for model-based RIP$_1$, where one wants to approximately preserve $\ell_1$-norms of $k$-block-sparse vectors, which is useful for $\ell_1/\ell_1$ recovery, \cite{IndykR13} have shown a much better sparsity bound of $O(\ceil{\log_b(n/k)})$ non-zeroes per column in their measurement matrix. However, they have also shown that any model-based RIP$_1$ matrix for block-sparse signals must satisfy the higher lower bound of $m\gtrsim k + (k/\log b)\log(n/k)$ (which is tight). Previous work also considered $T = H S_{n,k}$, where $H$ is any bounded orthonormal system, i.e.\ $H\in\mathbb{R}^{n\times n}$ is orthogonal and $\max_{i,j} |H_{ij}| = O(1/\sqrt{n})$ (e.g.\ the Fourier matrix). Work of \cite{CT06,RV08,CGV13} shows $\Phi$ can then be a sampling matrix (one non-zero per row) with $m\lesssim \varepsilon^{-2} k\log(n)(\log k)^3$. Since randomly flipping the signs of every column in an RIP matrix yields JL \cite{KW11}, this also gives a good implementation of an FJLT. Our theorem recovers a similar statement, but using the SJLT instead of a sampling matrix, where we show $m \mathbin{\substack{<\\ *}}\xspace k$ and $s \mathbin{\substack{<\\ *}}\xspace 1$ suffice for orthogonal $H$ satisfying the weaker requirement $\max_{i,j} |H_{ij}| = O(1/\sqrt{k})$. \medskip \paragraph{\textbf{Smooth manifolds:}} Suppose we are given several images of a human face, but with varying lighting and angle of rotation. Or perhaps the input is many sample handwritten images of letters. In these examples, although input data is high-dimensional ($n$ is the number of pixels), we imagine all possible inputs come from a set of low intrinsic dimension. That is, they lie on a $d$-dimensional manifold $\mathcal{M}\subset \mathbb{R}^n$ where $d\ll n$. The goal is then, given a large number of manifold examples, to learn the parameters of $\mathcal{M}$ to allow for nonlinear dimensionality reduction (reducing just to the few parameters of interest). This idea, and the first successful algorithm (ISOMAP) to learn a manifold from sampled points is due to \cite{TSL00}. Going back to the example of human faces, \cite{TSL00} shows that different images of a human face can be well-represented by a 3-dimensional manifold, where the parameters are brightness, left-right angle of rotation, and up-down angle of rotation. Since \cite{TSL00}, several more algorithms have been developed to handle more general classes of manifolds than ISOMAP \cite{RS00,DonohoG03}. Baraniuk and Wakin \cite{BW09} proposed using dimensionality-reducing maps to first map $\mathcal{M}$ to $\Phi \mathcal{M}$ and then learn the parameters of interest in the reduced space (for improved speed). Sharper analyses were later given in \cite{Clarkson08,EfW13,Dirksen14}. Of interest are both that (1) any $C^1$ curve in $\mathcal{M}$ should have length approximately preserved in $\Phi\mathcal{M}$, and (2) $\Phi$ should be a {\em manifold embedding}, in the sense that all $C^1$ curves $\gamma'\in\Phi\mathcal{M}$ should satisfy that the preimage of $\gamma'$ (in $\mathcal{M}$) is a $C^1$ curve in $\mathcal{M}$. Then by (1) and (2), {\em geodesic} distances are preserved by $\Phi$. To be concrete, let $\mathcal{M}\subset\mathbb{R}^n$ be a $d$-dimensional manifold obtained as the image $\mathcal{M} = F(B_{\ell_2^d})$, for smooth $F:B_{\ell_2^d}\rightarrow \mathbb{R}^n$ ($B_X$ is the unit ball of $X$). We assume $\|F(x) - F(y)\|_2 \simeq \|x - y\|_2$ (where $A\simeq B$ denotes that both $A\lesssim B$ and $A\gtrsim B$), and that the map sending $x\in\mathcal{M}$ to the tangent plane at $x$, $E_x$, is Lipschitz from $d_{\mathcal{M}}$ to $\rho_{\mathrm{Fin}}$. Here $\rho_{\mathcal{M}}$ is geodesic distance on $\mathcal{M}$, and $\rho_{\mathrm{Fin}}(E_x, E_y) = \|P_{E_x} - P_{E_y}\|_{\ell_2^n\rightarrow\ell_2^n}$ is the {\em Finsler distance}, where $P_E$ is the orthogonal projection onto $E$. We want $\Phi$ satisfying $(1-\varepsilon)|\gamma| \le |\Phi(\gamma)| \le (1+\varepsilon)|\gamma|$ for all $C^1$ curves $\gamma\subset\mathcal{M}$. Here $|\cdot|$ is curve length. To obtain this, it suffices that $\Phi$ satisfy \Equation{jl-condition} for $T = \bigcup_{x\in \mathcal{M}} E_x \cap S^{n-1}$ \cite{Dirksen14}, an infinite union of subspaces. Using this observation, \cite{Dirksen14} showed this property is satisfied for $s=m \lesssim d/\varepsilon^2$ with a {\em dense} matrix of subgaussian entries. For $F$ as given above, preservation of geodesic distances is also satisfied for this $m$. Our main theorem implies that to preserve curve lengths one can set $m \mathbin{\substack{<\\ *}}\xspace d$ for $s \mathbin{\substack{<\\ *}}\xspace 1 + (\alpha d)^2$, where $\alpha$ is the largest incoherence of any tangent space $E_x$ for $x\in\mathcal{M}$. That is, non-trivial sparsity with $m\mathbin{\substack{<\\ *}}\xspace d$ is possible for any $\alpha \ll 1/\sqrt{d}$. Furthermore, we show that this is {\em optimal} by constructing a manifold with maximum incoherence of a tangent space approximately $1/\sqrt{d}$ such that {\em any} linear map $\Phi$ preserving curve lengths with $m\mathbin{\substack{>\\ *}}\xspace d$ must have $s\mathbin{\substack{>\\ *}}\xspace d$ (see \Remark{bad-manifold}). We also show that $\Phi$ is a manifold embedding with large probability if the weaker condition $m\mathbin{\substack{>\\ *}}\xspace d, s \mathbin{\substack{>\\ *}}\xspace 1$ is satisfied. Combining these observations, we see that for the SJLT to preserve geodesic distances it suffices to set $m \mathbin{\substack{<\\ *}}\xspace d$ and $s \mathbin{\substack{<\\ *}}\xspace 1 + (\alpha d)^2$. \vspace{.1in} \begin{figure} \begin{center} {\scriptsize \begin{tabular}{|c|c|c|c|c|c|} \hline \textbf{set $T$ to preserve} & \textbf{our $m$} & \textbf{our $s$} & \textbf{previous $m$} & \textbf{previous $s$} & \textbf{ref}\\ \hline $|T|<\infty$ & $\log |T|$ & $\log |T|$ & $\log |T|$ & $\log |T|$ & \cite{JL84}\\ \hline $|T|<\infty,\forall x\in T \|x\|_\infty \le \alpha$ & $\log |T|$ & $\ceil{\alpha\log|T|}^2$ & $\log |T|$ & $\ceil{\alpha\log|T|}^2$ & \cite{Matousek08}\\ \hline $E,\mathrm{dim}(E)\le d$ & $d$ & $1$ & $d$ & $1$ & \cite{NN13b} \\ \hline $S_{n,k}$ & $k\log(n/k)$ & $k\log(n/k)$ & $k\log(n/k)$ & $k\log(n/k)$ & \cite{CT05}\\ \hline $HS_{n,k}$ & $k\log(n/k)$ & $1$ & $k\log(n/k)$ & $1$ & \cite{RV08}\\ \hline tangent cone for Lasso & $\max_j\frac{\|A_j\|_2^2}{\sigma_{min,k}^2}k$ & $\max_{i,j}\frac{A_{ij}^2}{\sigma_{min,k}^2}k$ & same as here & $s=m$ & \cite{PiW14}\\ \hline $|\Theta|<\infty$ & $d + \log|\Theta|$ & $\log|\Theta|$ & $d \cdot (\log|\Theta|)^6$& $(\log|\Theta|)^3$ & \cite{NN13b}\\ $\forall E\in\Theta,\mathrm{dim}(E)\le d$ &&&&&\\ \hline $|\Theta|<\infty$ & $d + \log|\Theta|$ & $\ceil{\alpha \log|\Theta|}^2$ & --- & --- & --- \\ $\forall E\in\Theta,\mathrm{dim}(E)\le d$ &&&&&\\ $\max_{\substack{1\le j\le n\\ E\in\Theta}} \|P_E e_j\|_2 \le \alpha$&&&&&\\ \hline $|\Theta|$ infinite & see appendix & see appendix & similar to& $m$ & \cite{Dirksen14}\\ $\forall E\in\Theta,\mathrm{dim}(E)\le d$ & & (non-trivial) & this work & &\\ \hline $\mathcal{M}$ a smooth manifold & $d$& $1+(\alpha d)^2$&$d$ & $d$ & \cite{Dirksen14} \\ \hline \end{tabular} } \caption{The $m,s$ that suffice when using the SJLT with various $T$ as a consequence of our main theorem, compared with the best known bounds from previous work. All bounds shown hide $\mathrm{poly}(\varepsilon^{-1}\log n)$ factors. One row is blank in previous work due to no non-trivial results being previously known. For the case of the Lasso, we assume $k$ is the sparsity of the true optimal solution.}\FigureName{fig} \end{center \end{figure} As seen above, not only does our answer to \Question{main} qualitatively explain all known results, but it gives new results not known before with implications in numerical linear algebra, compressed sensing (model-based, and with incoherent dictionaries), constrained least squares, and manifold learning. We also believe it is possible for future work to sharpen our analyses to give asymptotically correct parameters for all the applications; see the discussion in \Section{discussion}. We now end the introduction with an outline for the remainder. \Section{prelim} defines the notation for the rest of the paper. \Section{overview} provides an overview of the proof of our main theorem, \Theorem{main}. \Section{linear} is a warmup that applies a subset of the proof ideas for \Theorem{main} to the special case where $T = E\cap S^{n-1}$, $E$ a linear subspace of dimension $d$. In fact the proof of our main theorem reduces the case of general $T$ to several linear subspaces of varying dimensions, and thus this case serves as a useful warmup. \Section{typetwo} applies a different subset of our proof ideas to the special case where the norm $\tnorm{\cdot}_T$ defined by $\tnorm{y}_T = \sup_{x\in T}|\inprod{x,y}|$ has a small type-$2$ constant, which is relevant for analyzing the FFT-based approaches to RIP matrices. \Section{cls} shows how similar ideas can be applied to constrained least squares problems, such as Lasso. \Section{general} states and proves our most general theorem for arbitrary $T$. \Section{applications} shows how to apply \Theorem{main} to obtain good bounds for various $T$, albeit losing a $\log^c(n/\varepsilon)$ factor in $m$ and a $\log^c(n/\varepsilon)/\varepsilon$ factor in $s$ as mentioned above. Finally, \Section{discussion} discusses avenues for future work. In the appendix, in Appendix~\ref{sec:probTools} for the benefit of the reader we review many probabilistic tools that are used throughout this work . Appendix~\ref{sec:convex-tools} reviews some introductory material related to convex analysis, which is helpful for understanding \Section{cls} on constrained least squares problems. In Appendix~\ref{sec:fjlt-cls} we also give a direct analysis for using the FJLT for sketching constrained least squares programs, providing quantitative benefits over some analyses in \cite{PiW14}. \section{Preliminaries}\SectionName{prelim} We fix some notation that we will be used throughout this paper. For a positive integer $t$, we set $[t] = \{1,\ldots,t\}$. For $a,b \in \mathbb{R}$, $a\lesssim b$ denotes $a \le C b$ for some universal constant $C > 0$, and $a\simeq b$ signifies that both $a\lesssim b$ and $b\lesssim a$ hold. For any $x\in \mathbb{R}^n$ and $1\leq p\leq\infty$, we let $\|x\|_p$ denote its $\ell_p$-norm. To any set $S\subset\mathbb{R}^n$ we associate a semi-norm $\tnorm{\cdot}_S$ defined by $\tnorm{z}_S = \sup_{x\in S}|\inprod{z,x}|$. Note that $\tnorm{z}_S = \tnorm{z}_{\mathrm{conv}(S)}$, where $\mathrm{conv}(S)$ is the closed convex hull of $S$, i.e., the closure of the set of all convex combinations of elements in $S$. We use $(e_i)_{1\leq i\leq n}$ and $(e_{ij})_{1\leq i\leq m,1\leq j\leq n}$ to denote the standard basis in $\mathbb{R}^n$ and $\mathbb{R}^{m\times n}$, respectively.\par If $\eta=(\eta_i)_{i\geq 1}$ is a sequence of random variables, we let $(\Omega_{\eta},\mathbb{P}_{\eta})$ denote the probability space on which it is defined. We use $\mathbb{E}_{\eta}$ and $L_{\eta}^p$ to denote the associated expected value and $L^p$-space, respectively. If $\zeta$ is another sequence of random variables, then $\|\cdot\|_{L^p_{\eta},L^q_{\zeta}}$ means that we first take the $L^p_{\eta}$-norm and afterwards the $L^q_{\zeta}$-norm. We reserve the symbol $g$ to denote a sequence $g=(g_i)_{i\geq 1}$ of i.i.d.\ standard Gaussian random variables; unless stated otherwise, the covariance matrix is the identity.\par If $A$ is an $m\times n$ matrix, then we use $\|A\|$ or $\|A\|_{\ell_2^n\rightarrow\ell_2^m}$ to denote its operator norm. Moreover we let $\mathrm{Tr}$ be the trace operator and use $\|A\|_F=(\mathrm{Tr}(A^*A))^{1/2}$ to denote the Frobenius norm. In the remainder, we reserve the letter $\rho$ to denote (semi-)metrics. If $\rho$ corresponds to a (semi-)norm $\|\cdot\|_X$, then we let $\rho_X(x,y)=\|x-y\|_X$ denote the associated (semi-)metric. Also, we use $d_{\rho}(S)=\sup_{x,y \in S}\rho(x,y)$ to denote the diameter of a set $S$ with respect to $\rho$ and write $d_X$ instead of $d_{\rho_X}$ for brevity. So, for example, $\rho_{\ell_2^n}$ is the Euclidean metric and $d_{\ell_2^n}(S)$ the $\ell_2$-diameter of $S$. From here on, $T$ is always a fixed subset of $S^{n-1} = \{x\in\mathbb{R}^n : \|x\| = 1\}$, and $\varepsilon\in (0,1/2)$ the parameter appearing in \Eqsub{rjl-condition}. We make use of chaining results in the remainder, so we define some relevant quantities. For a bounded set $S\subset \mathbb{R}^n$, $g(S) = \mathbb{E}_g \sup_{x\in S} \inprod{g, x}$ is the {\em Gaussian mean width} of $S$, where $g\in\mathbb{R}^n$ is a Gaussian vector with identity covariance matrix. For a (semi-)metric $\rho$ on $\mathbb{R}^n$, Talagrand's $\gamma_2$-functional is defined by \begin{equation} \gamma_2(S,\rho) = \inf_{\{S_r\}_{r=0}^\infty} \sup_{x\in S} \sum_{r=0}^\infty 2^{r/2} \cdot \rho(x, S_r) \end{equation} where $\rho(x, S_r)$ is the distance from $x$ to $S_r$, and the infimum is taken over all collections $\{S_r\}_{r=0}^\infty$, $S_0 \subset S_1 \subset \ldots \subseteq S$, with $|S_0| = 1, |S_r| \le 2^{2^r}$. If $\rho$ corresponds to a (semi-)norm $\|\cdot\|_X$, then we usually write $\gamma_2(S, \|\cdot\|_X)$ instead of $\gamma_2(S,\rho_X)$. It is known that for any bounded $S\subset \mathbb{R}^n$, $g(S)$ and $\gamma_2(S, \|\cdot\|_2)$ differ multiplicatively by at most a universal constant \cite{Fernique75,Talagrand05}. Whenever $\gamma_2$ appears without a specified norm, we imply use of $\ell_2$ or $\ell_2\rightarrow\ell_2$ operator norm. We frequently use the entropy integral estimate (see \cite{Talagrand05}) \begin{equation} \gamma_2(S, \rho) \lesssim \int_0^{\infty} \log^{1/2} \mathcal{N}(S,\rho, u) du .\EquationName{Dudley} \end{equation} Here $\mathcal{N}(S,\rho, u)$ denotes the minimum number of $\rho$-balls of radius $u$ centered at points in $S$ required to cover $S$. If $\rho$ corresponds to a (semi-)norm $\|\cdot\|_X$, then we write $\mathcal{N}(S,\|\cdot\|_X, u)$ instead of $\mathcal{N}(S,\rho_X, u)$. Let us now introduce the sparse Johnson-Lindenstrauss transform in detail. Let $\sigma_{ij}:\Omega_{\sigma}\rightarrow \{-1,1\}$ be independent Rademacher random variables, i.e., $\mathbb{P}(\sigma_{ij}=1)=\mathbb{P}(\sigma_{ij}=-1)=1/2$. We consider random variables $\delta_{ij}:\Omega_{\delta}\rightarrow\{0,1\}$ with the following properties: \begin{itemize} \item For fixed $j$ the $\delta_{ij}$ are negatively correlated, i.e. \begin{equation} \forall 1\leq i_1<i_2<\ldots<i_k\leq m,\ \mathbb{E}\Big(\prod_{t=1}^k\delta_{i_t,j}\Big)\leq \prod_{t=1}^k\mathbb{E} \delta_{i_t,j} = \Big(\frac sm\Big)^k;\EquationName{neg-cor} \end{equation} \item For any fixed $j$ there are exactly $s$ nonzero $\delta_{ij}$, i.e., $\sum_{i=1}^m \delta_{ij}=s$; \item The vectors $(\delta_{ij})_{i=1}^m$ are independent across different $1\le j\le n$. \end{itemize} We emphasize that the $\sigma_{ij}$ and $\delta_{ij}$ are independent, as they are defined on different probability spaces. The \emph{sparse Johnson-Lindenstrauss transform $\Phi$ with column sparsity $s$}, or \emph{SJLT} for short, is defined by \begin{equation} \Phi_{ij}=\frac{1}{\sqrt{s}}\sigma_{ij}\delta_{ij}.\EquationName{1.1} \end{equation} The work \cite{KN14} gives two implementations of such a $\Phi$ satisfying the above conditions. In one example, the columns are independent, and in each column we choose exactly $s$ locations uniformly at random, without replacement, to specify the $\delta_{ij}$. The other example is essentially the CountSketch of \cite{CCF04}. In this implementation, the rows of $\Phi$ are partitioned arbitrarily into $s$ groups of size $m/s$ each. Then each column of $\Phi$ is chosen independently, where in each column and for each block we pick a random row in that block to be non-zero for that column (the rest are zero). We say that $\Phi$ is an \emph{$\varepsilon$-restricted isometry} on $T$ if \Equation{jl-condition} holds. We define $$\varepsilon_{\Phi,T} = \sup_{x\in T} |\|\Phi x\|_2^2 - 1|$$ as the \emph{restricted isometry constant} of $\Phi$ on $T$. In the following we will be interested in estimating $\varepsilon_{\Phi,T}$. For this purpose, we use the following $L^p$-bound for the supremum of a second order Rademacher chaos from \cite{KMR14} (see \cite[Theorem 6.5]{Dir13} for the refinement stated here). \begin{theorem} Let $\mathcal{A}\subset \mathbb{R}^{m\times n}$ and let $\sigma_1,\ldots,\sigma_n$ be independent Rademacher random variables. For any $1\leq p<\infty$ \begin{align} \label{eqn:supChaosImproved} \Big(\mathbb{E}_{\sigma}\sup_{A\in\mathcal{A}}\Big|\|A\sigma\|_2^2 - \mathbb{E}\|A\sigma\|_2^2\Big|^p\Big)^{1/p} & \lesssim \gamma_{2}^2(\mathcal{A},\|\cdot\|_{\ell_2\rightarrow\ell_2}) + d_F(\mathcal{A})\gamma_{2}(\mathcal{A},\|\cdot\|_{\ell_2\rightarrow\ell_2}) \nonumber \\ & \qquad + \sqrt{p}d_{F}(\mathcal{A})d_{\ell_2\rightarrow\ell_2}(\mathcal{A}) + pd_{\ell_2\rightarrow\ell_2}^2(\mathcal{A}). \end{align} \end{theorem} For $u \in \mathbb{R}^n$ and $v\in \mathbb{R}^m$, let $u\otimes v:\mathbb{R}^m\rightarrow \mathbb{R}^n$ be the operator $(u\otimes v)w=\langle v,w\rangle u$. Then, for any $x\in \mathbb{R}^n$ we can write $\Phi x=A_{\delta,x}\sigma$, where \begin{equation} \label{eqn:AdelxDef} A_{\delta,x} := \frac{1}{\sqrt{s}}\sum_{i=1}^m\sum_{j=1}^n \delta_{ij}x_j e_i\otimes e_{ij} = \frac 1{\sqrt{s}}\begin{bmatrix} - x^{(\delta_{1,\cdot})} - & 0 & \cdots & 0\\ 0 & - x^{(\delta_{2,\cdot})} - & \cdots & 0\\ \vdots &\vdots & &\vdots\\ 0&0&\cdots& - x^{(\delta_{m,\cdot})} - \end{bmatrix} . \end{equation} for $x^{(\delta_{i,\cdot})}_j = \delta_{ij}x_j$. Note that $\mathbb{E}\|\Phi x\|_2^2 = \|x\|_2^2$ for all $x\in \mathbb{R}^n$ and therefore $$\sup_{x\in T}|\|\Phi x\|_2^2 - \|x\|_2^2| = \sup_{x\in T} |\|A_{\delta,x}\sigma\|_2^2 - \mathbb{E}\|A_{\delta,x}\sigma\|_2^2|.$$ Associated with $\delta=(\delta_{ij})$ we define a random norm on $\mathbb{R}^n$ by \begin{equation} \|x\|_{\delta} = \frac{1}{\sqrt{s}}\max_{1\leq i\leq m}\Big(\sum_{j=1}^n \delta_{ij}x_j^2\Big)^{1/2}. \EquationName{defDelNorm} \end{equation} With this definition, we have for any $x,y\in T$, \begin{equation} \label{eqn:AdelxNorm} \|A_{\delta,x} - A_{\delta,y}\| = \|x-y\|_{\delta}, \qquad \|A_{\delta,x} - A_{\delta,y}\|_F = \|x-y\|_2. \end{equation} Therefore, (\ref{eqn:supChaosImproved}) implies that \begin{align} & \Big(\mathbb{E}_{\sigma}\sup_{x \in T} \Big|\|\Phi x\|_2^2 - \|x\|_2^2\Big|^p\Big)^{1/p} \lesssim \gamma_2^{2}(T,\|\cdot\|_{\delta}) + \gamma_2(T,\|\cdot\|_{\delta}) + \sqrt{p}d_{\delta}(T) + pd_{\delta}^2(T).\EquationName{RIPExpect} \end{align} Taking $L_p(\Omega_{\delta})$-norms on both sides yields \begin{align} \nonumber \Big(\mathbb{E}_{\delta,\sigma}\sup_{x \in T} \Big|\|\Phi x\|_2^2 - \|x\|_2^2\Big|^p\Big)^{1/p} & \lesssim (\mathbb{E}_{\delta}\gamma_2^{2p}(T,\|\cdot\|_{\delta}))^{1/p} + (\mathbb{E}_{\delta}\gamma_2^{p}(T,\|\cdot\|_{\delta}))^{1/p} \\ & \qquad \qquad + \sqrt{p}(\mathbb{E} d_{\delta}^p(T))^{1/p} + p(\mathbb{E} d_{\delta}^{2p}(T))^{1/p}. \EquationName{RIPLpEst} \end{align} Thus, to find a bound for the expected value of the restricted isometry constant of $\Phi$ on $T$, it suffices to estimate $\mathbb{E}_{\delta}\gamma_2^2(T,\|\cdot\|_{\delta})$ and $\mathbb{E}_{\delta}d_{\delta}^2(T)$. If we can find good bounds on $(\mathbb{E}_{\delta}\gamma_2^{p}(T,\|\cdot\|_{\delta}))^{1/p}$ and $(\mathbb{E} d_{\delta}^p(T))^{1/p}$ for all $p\geq 1$, then we in addition obtain a high probability bound for the restricted isometry constant.\par Unless explicitly stated otherwise, from here on $\Phi$ always denotes the SJLT with $s$ non-zeroes per column. \section{Overview of proof of main theorem}\SectionName{overview} Here we give an overview of the proof of \Theorem{main}. To illustrate the ideas going into our proof, it is simplest to first consider the case where $T$ is the set of all unit vectors in a $d$-dimensional linear subspace $E\subset\mathbb{R}^n$. By \Equation{RIPLpEst} for $p=1$ we have to bound for example $\mathbb{E}_\delta \gamma_2(T,\|\cdot\|_\delta)$. Standard estimates give, up to $\log d$, \begin{equation} \gamma_2(T, \|\cdot\|_\delta) \le \gamma_2(B_E, \|\cdot\|_\delta) \ll \sup_{t>0} t [\log\mathcal{N}(B_E, \|\cdot\|_\delta, t)]^{1/2} . \EquationName{overview} \end{equation} for $B_E$ the unit ball of $E$. Let $U\in\mathbb{R}^{n\times d}$ have columns forming an orthonormal basis for $E$. Then dual Sudakov minoration \cite[Proposition 4.2]{BLM89}, \cite{PTJ86} states that \begin{equation} \sup_{t>0} t [\log\mathcal{N}(B_E, \|\cdot\|_\delta, t)]^{1/2} \le \mathbb{E}_g \|Ug\|_\delta \EquationName{sudakov} \end{equation} for a Gaussian vector $g$. From this point onward one can arrive at a result using the non-commutative Khintchine inequality \cite{LP86,LPP91} and other standard Gaussian concentration arguments (see appendix). Unfortunately, \Equation{sudakov} is very specific to unit balls of linear subspaces and has no analog for a general set $T$. At this point we use a statement about the duality of entropy numbers \cite[Proposition 4]{BPST89}. This is the principle that for two symmetric convex bodies $K$ and $D$, $\mathcal{N}(K, D)$ and $\mathcal{N}(D^\circ, a K^\circ)$ are roughly comparable for some constant $a$ ($\mathcal{N}(K,D)$ is the number of translates of $D$ needed to cover $K$). Although it has been an open conjecture for over 40 years as to whether this holds in the general case \cite[p.\ 38]{P72}, the work \cite[Proposition 4]{BPST89} shows that these quantities are comparable up to logarithmic factors as well as a factor depending on the type-$2$ constant of the norm defined by $D$ (i.e.\ the norm whose unit vectors are those on the boundary of $D$). In our case, this lets us relate $\log \mathcal{N}(\tilde{T}, \|\cdot\|_\delta, t)$ with $\log\mathcal{N}(\mathrm{conv}(\cup_{i=1}^m B_{J_i}), \tnorm{\cdot}_T, \sqrt{s}t/8)$, losing small factors. Here $\tilde{T}$ is the convex hull of $T\cup -T$, and $B_{J_i}$ is the unit ball of $\mathrm{span}\{e_j : \delta_{ij} = 1\}$. We next use Maurey's lemma, which is a tool for bounding covering numbers of the set of convex combinations of vectors in various spaces. This lets us relate $\log\mathcal{N}(\mathrm{conv}(\cup_{i=1}^m B_{J_i}), \tnorm{\cdot}_T, \epsilon)$ to quantities of the form $\log\mathcal{N}(\frac 1k\sum_{i\in A}B_{J_i}, \tnorm{\cdot}_T, \epsilon)$, where $A\subset [m]$ has size $k\lesssim 1/\epsilon^2$ (see \Lemma{maurey}). For a fixed $A$, we bucket $j\in [n]$ according to $\sum_{i\in A} \delta_{ij}$ and define $U_\alpha = \{j\in [n] : \sum_{i\in A} \delta_{ij} \simeq 2^\alpha\}$. Abusing notation, we also let $U_\alpha$ denote the coordinate subspace spanned by $j\in U_\alpha$. This leads to (see \Equation{5.4}) \begin{equation} \log\mathcal{N}\Big(\frac 1k\sum_{i\in A}B_{J_i}, \tnorm{\cdot}_T, \epsilon\Big) \lesssim \sum_\alpha \log \mathcal{N}\Big(B_{U_\alpha}, \tnorm{\cdot}_T, \sqrt{\frac{k}{2^\alpha}} \frac{\epsilon}{\log m}\Big) \EquationName{finally-sudakov} \end{equation} Finally we are in a position to apply dual Sudakov minoration to the right hand side of \Equation{finally-sudakov}, after which point we apply various concentration arguments to yield our main theorem. \section{The case of a linear subspace}\SectionName{linear} Let $E\subset\mathbb{R}^n$ be a $d$-dimensional linear subspace, $T = E\cap S^{n-1}$, $B_E$ the unit $\ell_2$-ball of $E$. We use $P_E$ to denote the orthogonal projection onto $E$. The values $\|P_E e_j\|_2$, $j=1,\ldots,n$, are typically referred to as the {\em leverage scores} of $E$ in the numerical linear algebra literature. We denote the maximum leverage score by $$\mu(E) = \max_{1\leq j\leq n}\|P_E e_j\|_2,$$ which is sometimes referred to as the {\em incoherence} $\mu(E)$ of $E$. In our proof we make use of the well-known Gaussian concentration of Lipschitz functions \cite[Corollary 2.3]{Pi86}: if $g$ is an $n$-dimensional standard Gaussian vector and $\phi:\mathbb{R}^n\rightarrow \mathbb{R}$ is $L$-Lipschitz from $\ell_2^n$ to $\mathbb{R}$, then for all $p\ge 1$ \begin{equation} \|\phi(g) - \mathbb{E}\phi(g)\|_{L^p_g} \lesssim L \sqrt{p}. \EquationName{GaussLip} \end{equation} \begin{theorem} \label{thm:JLsubspace} For any $p\geq \log m$ and any $0<\epsilon<1$, \begin{equation} (\mathbb{E}\gamma_2^{2p}(T,\|\cdot\|_{\delta}))^{1/p} \lesssim \epsilon^2 + \frac{(d+\log m)\log^2(d/\epsilon)}{m} + \frac{p\log^2(d/\epsilon)\log m}{s}\mu(E)^2 \EquationName{ga2Subs} \end{equation} and \begin{equation} (\mathbb{E} d^{2p}_{\delta}(T))^{1/p} \lesssim \frac{d}{m} + \frac{p}{s}\mu(E)^2. \EquationName{diamSubs} \end{equation} As a consequence, if $\eta\leq 1/m$ and \begin{align} \nonumber m &\gtrsim ((d+\log m)\min\{\log^2(d/\varepsilon),\log^2(m)\} + d\log(\eta^{-1}))/\varepsilon^2\\ s &\gtrsim (\log(m)\log(\eta^{-1})\min\{\log^2(d/\varepsilon),\log^2(m)\}+\log^2(\eta^{-1}))\mu(E)^2/\varepsilon^2\EquationName{two-ten} \end{align} then \Equation{jl-condition} holds with probability at least $1-\eta$. \end{theorem} \begin{proof} By dual Sudakov minoration (Lemma~\ref{lem:dualSudakov} in Appendix~\ref{sec:probTools}) \begin{equation*} \log \mathcal{N}(B_E, \|\cdot\|_{\delta}, t) \lesssim \frac{\mathbb{E}_g \|U g\|_{\delta}}t, \qquad \mathrm{for \ all} \ t>0, \EquationName{2.1} \end{equation*} with $U\in\mathbb{R}^{n\times d}$ having columns $\{f_k\}_{k=1}^d$ forming an orthonormal basis for $E$ and $g$ a random Gaussian vector. Let $U^{(i)}$ be $U$ but where each row $j$ is multiplied by $\delta_{ij}$. Then by \Equation{defDelNorm} and taking $\ell = \log m$, \begin{align} \nonumber \mathbb{E}_g \|U g\|_{\delta} &= \frac 1{\sqrt{s}} \mathbb{E}_g\max_{1\le i \le m} \Big[\sum_{j=1}^n \delta_{ij} \Big|\sum_{k=1}^d g_k \inprod{f_k, e_j} \Big|^2\Big]^{1/2}\\ \nonumber {}&= \frac 1{\sqrt{s}} \mathbb{E}_g \max_{1\le i \le m} \|U^{(i)} g\|_2\\ \nonumber {}&\le \frac 1{\sqrt{s}}\Big(\max_{1\le i\le m} \mathbb{E}_g \|U^{(i)} g\|_2 + \mathbb{E}_g \max_{1\le i\le m} \Big|\|U^{(i)} g\|_2 - \mathbb{E}_g \|U^{(i)} g\|_2\Big|\Big)\\ \nonumber {}&\le \frac 1{\sqrt{s}}\Big(\max_{1\le i\le m} \|U^{(i)}\|_F + \Big(\sum_{i=1}^m \mathbb{E}_g \Big|\|U^{(i)} g\|_2 - \mathbb{E}_g \|U^{(i)} g\|_2\Big|^\ell\Big)^{1/\ell}\Big)\\ {}&\lesssim \frac 1{\sqrt{s}}\Big(\max_{1\le i\le m} \|U^{(i)}\|_F + \sqrt{\ell} \cdot \max_{1\le i\le m} \|U^{(i)}\|_{\ell_2^d\rightarrow\ell_2^n}\Big) \EquationName{2.2} \end{align} where \Equation{2.2} used Gaussian concentration of Lipschitz functions (cf.\ \Equation{GaussLip}), noting that $g\mapsto \|U^{(i)} g\|_2$ is $\|U^{(i)}\|$-Lipschitz. Since $\|\cdot\|_{\delta} \le (1/\sqrt{s})\|\cdot\|_2$, we can use for small $t$ the bound (cf.\ \Equation{volComp}) \begin{align} \nonumber \log \mathcal{N}(T, \|\cdot\|_{\delta}, t) &\le \log \mathcal{N}(B_E, \|\cdot\|_2, t\sqrt{s})\\ &< d \cdot \log\Big(1 + \frac{2}{t\sqrt{s}}\Big) \EquationName{2.3} \end{align} Using \Equation{2.3} for small $t$ and noting $\|U^{(i)}\|_F = (\sum_{j=1}^n \delta_{ij}\|P_E e_j\|_2^2)^{1/2}$ for $P_E$ the orthogonal projection onto $E$, \Equation{Dudley} yields for $t^* = (\epsilon/d)/\log(d/\epsilon)$ \begin{align} & \gamma_2(T,\|\cdot\|_{\delta}) \nonumber \\ & \qquad \lesssim \int_0^{t^*} \sqrt{d}\cdot \Big[\log\Big(2 + \frac{1}{t\sqrt{s}}\Big)\Big]^{1/2} dt + \int_{t^*}^{1/\sqrt{s}} \frac{\mathbb{E}_g \|Ug\|_\delta}t dt \nonumber \\ & \qquad \lesssim \sqrt{d}t_*\Big[\log\Big(\frac{1}{t_*\sqrt{s}}\Big)\Big]^{1/2} + \mathbb{E}_g \|Ug\|_\delta \log\Big(\frac{1}{t_*\sqrt{s}}\Big) \nonumber \\ {}& \qquad \lesssim \epsilon + \frac{\log(d/\epsilon)}{\sqrt{s}}\cdot \Big[\max_{1\le i\le m} \Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2 \Big]^{1/2} + \sqrt{\log m}\max_{1\le i\le m} \|U^{(i)}\|_{\ell_2^d\rightarrow\ell_2^n}\Big] \EquationName{splittingArg} \end{align} As a consequence, \begin{align} (\mathbb{E}_{\delta}\gamma_2^{2p}(T,\|\cdot\|_\delta))^{1/p} & \lesssim \epsilon^2 + \frac{\log^2(d/\epsilon)}{s} \Big[\Big(\mathbb{E}_{\delta}\max_{1\le i\le m} \Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2 \Big]^{p}\Big)^{1/p} \nonumber \\ & \qquad \qquad \qquad \qquad + \log(m)\Big(\mathbb{E}_{\delta}\max_{1\le i\le m} \|U^{(i)}\|_{\ell_2^d\rightarrow\ell_2^n}^{2p}\Big)^{1/p}\Big]\EquationName{ga2SubsIntermed} \end{align} We estimate the first non-trivial term on the right hand side. Since $p\geq \log m$, \begin{align*} \Big(\mathbb{E}_{\delta}\max_{1\le i\le m} \Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2 \Big]^{p}\Big)^{1/p} & \le \Big(\sum_{i=1}^m\mathbb{E}_\delta\Big|\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2\Big|^p\Big)^{1/p} \\ & \leq e\cdot\max_{1\leq i\leq m} \Big(\mathbb{E}_{\delta}\Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2 \Big]^{p}\Big)^{1/p} \end{align*} Note that for fixed $i$, the $(\delta_{ij})_{1\le j\le n}$ are i.i.d.\ variables of mean $s/m$ and therefore \begin{equation*} \mathbb{E}\sum_{j=1}^n \delta_{ij}\|P_E e_j\|_2^2 = \frac{s}{m} \sum_{j=1}^n \langle P_E e_j,e_j\rangle = \frac{s}{m}\mathrm{Tr}(P_E) = \frac{sd}{m}. \end{equation*} Let $(r_j)_{1\le j\le n}$ be a vector of independent Rademachers. By symmetrization (\Equation{symmetrization}) and Khintchine's inequality (\Equation{KI}), \begin{align*} \Big\|\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2\Big\|_{L^p_{\delta}} &\le \frac{ds}m + \Big\|\sum_{j=1}^n (\delta_{ij}-\mathbb{E}\delta_{ij}) \|P_E e_j\|_2^2\Big\|_{L^p_{\delta}} \\ {}&\le \frac{ds}m + 2\Big\|\sum_{j=1}^n r_j\delta_{ij}\|P_E e_j\|_2^2\Big\|_{L^p_{\delta,r}}\\ {}&\lesssim \frac{ds}m + \sqrt{p}\Big\|\Big(\sum_{j=1}^n\delta_{ij}\|P_E e_j\|_2^4\Big)^{1/2}\Big\|_{L^p_{\delta}}\\ {}&\le \frac{ds}m + \sqrt{p}\cdot \max_{1\le j\le n}\|P_E e_j\|_2\cdot \Big\|\sum_{j=1}^n\delta_{ij}\|P_E e_j\|_2^2\Big\|_{L^p_{\delta}}^{1/2} \end{align*} Solving this quadratic inequality yields \begin{equation} \Big\|\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2\Big\|_{L^p_{\delta}} \lesssim \frac{ds}m + p\cdot \max_{1\le j\le n}\|P_E e_j\|_2^2 \end{equation} We conclude that \begin{equation} \Big(\mathbb{E}_{\delta}\max_{1\le i\le m} \Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2 \Big]^{p}\Big)^{1/p} \lesssim \frac{ds}{m} + p\max_j \|P_E e_j\|_2^2\EquationName{sqrt-trick} \end{equation} We now estimate the last term on the right hand side of \Equation{ga2SubsIntermed}. As $p\geq \log m$, \begin{align*} \Big(\mathbb{E}_\delta \max_{1\le i\le m} \|U^{(i)}\|_{\ell_2^d\rightarrow\ell_2^n}^{2p}\Big)^{1/p} {}&\lesssim \max_{1\leq i\leq m}\Big(\mathbb{E} \|U^{(i)}\|^{2p}_{\ell_2^d\rightarrow\ell_2^n}\Big)^{1/p}\\ {}& = \max_{1\leq i\leq m}\Big(\mathbb{E} \|(U^{(i)})^*U^{(i)}\|^{p}_{\ell_2^d\rightarrow\ell_2^d}\Big)^{1/p}. \end{align*} If $u_j$ denotes the $j$th row of $U$, then $$ (U^{(i)})^*U^{(i)} = \sum_{j=1}^n \delta_{ij} u_j u_j^*. $$ By symmetrization and the non-commutative Khintchine inequality (cf.\ \Equation{NCKI}), \allowdisplaybreaks \begin{align*} \Big\|\sum_{j=1}^n \delta_{ij} u_j u_j^*\Big\|_{L^p_{\delta}} &\le \frac sm + \Big\|\sum_{j=1}^n \delta_{ij} u_j u_j^* - \frac sm I\Big\|_{L^p_{\delta}}\\ &\le \frac sm + 2\Big\|\sum_{j=1}^n r_j\delta_{ij} u_j u_j^*\Big\|_{L^p_{\delta,r}}\\ &\lesssim \frac sm + \sqrt{p} \Big\|\sum_{j=1}^n \delta_{ij} \|u_j\|_2^2 u_j u_j^*\Big\|_{L^{p/2}_{\delta}}^{1/2}\\ &\le \frac sm + \sqrt{p} \max_{1\le j\le n} \|P_E e_j\|_2\cdot \Big\|\sum_{j=1}^n \delta_{ij} u_j u_j^*\Big\|_{L^{p}_{\delta}}^{1/2} \end{align*} Solving this quadratic inequality, we find \begin{equation*} \Big\|\sum_{j=1}^n \delta_{ij} u_j u_j^*\Big\|_{L^p_{\delta}} \lesssim \frac s m + p \cdot \max_{1\le j\le n} \|P_E e_j\|_2^2 \end{equation*} and as a consequence, \begin{equation*} \Big(\mathbb{E}_\delta \max_{1\le i\le m} \|U^{(i)}\|_{\ell_2^d\rightarrow\ell_2^n}^{2p}\Big)^{1/p} \lesssim \frac s m + p \cdot \max_{1\le j\le n} \|P_E e_j\|_2^2 \end{equation*} Applying this estimate together with \Equation{sqrt-trick} in \Equation{ga2SubsIntermed} yields the asserted estimate in \Equation{ga2Subs}. For the second assertion, note by Cauchy-Schwarz that \begin{align*} d_{\delta}(T) & = \frac{1}{\sqrt{s}} \max_{1\le i\le m} \sup_{x\in T} \Big[\sum_{j=1}^n \delta_{ij} x_j^2\Big]^{1/2} \\ & = \frac{1}{\sqrt{s}} \max_{1\le i\le m} \sup_{x\in T} \Big[\sum_{j=1}^n \delta_{ij} \Big(\sum_{k=1}^d \inprod{x, f_k}\inprod{f_k, e_j}\Big)^2\Big]^{1/2}\\ & \le \frac{1}{\sqrt{s}} \max_{1\le i\le m} \Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2\Big]^{1/2}. \end{align*} Therefore, $$(\mathbb{E}_{\delta}d^{2p}_{\delta}(T))^{1/p} \leq \frac{1}{s}\Big(\mathbb{E}_{\delta}\max_{1\le i\le m} \Big[\sum_{j=1}^n \delta_{ij} \|P_E e_j\|_2^2 \Big]^{p}\Big)^{1/p}$$ and \Equation{diamSubs} immediately follows from \Equation{sqrt-trick}.\par To prove the final assertion, we combine \Eqsub{RIPLpEst}, \Equation{ga2Subs} and \Equation{diamSubs} to obtain \begin{align*} & \Big(\mathbb{E}_{\delta,\sigma}\sup_{x \in T} \Big|\|\Phi x\|^2 - \|x\|^2\Big|^p\Big)^{1/p} \\ & \qquad \lesssim \epsilon^2 + \frac{(d+\log m)\log^2(d/\epsilon)}{m} + \frac{p\log^2(d/\epsilon)\log m}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2 \\ & \qquad \qquad \ + \Big[\epsilon^2 + \frac{(d+\log m)\log^2(d/\epsilon)}{m} + \frac{p\log^2(d/\epsilon)\log m}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2\Big]^{1/2} \\ & \qquad \qquad \ + \frac{pd}{m} + \frac{p^2}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2 + \Big[\frac{pd}{m} + \frac{p^2}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2\Big]^{1/2}. \end{align*} Now apply Lemma~\ref{lem:MomentsToTails} of Appendix~\ref{sec:probTools} to obtain, for any $w\geq \log m$, \begin{align*} \mathbb{P}_{\delta,\sigma}\Big(\varepsilon_{\Phi,T} & \gtrsim \epsilon^2 + \frac{(d+\log m)\log^2(d/\epsilon)}{m} + \frac{w\log^2(d/\epsilon)\log m}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2 \\ & \ \ \ + \Big[\epsilon^2 + \frac{(d+\log m)\log^2(d/\epsilon)}{m} + \frac{w\log^2(d/\epsilon)\log m}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2\Big]^{1/2} \\ & \ \ \ + \frac{wd}{m} + \frac{w^2}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2 + \Big[\frac{wd}{m} + \frac{w^2}{s}\max_{1\leq j\leq n} \|P_E e_j\|_2^2\Big]^{1/2}\Big) \leq e^{-w}. \end{align*} Now set $w=\log(\eta^{-1})$, choose $\epsilon=\varepsilon/C$ (with $C$ a large enough absolute constant) and $\epsilon=d/m$, and use the assumptions on $m$ and $s$ to conclude $\mathbb{P}(\varepsilon_{\Phi,T}\geq \varepsilon)\leq \eta$. \end{proof} Theorem~\ref{thm:JLsubspace} recovers a similar result in \cite{NN13b} but via a different method, less logarithmic factors in the setting of $m$, and the revelation that $s$ can be taken smaller if all the leverage scores $\|P_E e_j\|_2$ are small (note if $\|P_E e_j\|_2 \ll (\log d\cdot \log m)^{-1}$ for all $j$, we may take $s=1$, though this is not true in general). Our dependence on $1/\varepsilon$ in $s$ is quadratic instead of the linear dependence in \cite{NN13b}, but as stated earlier, in most applications of OSE's $\varepsilon$ is taken a constant. \begin{remark} \label{rem:coherence} \textup{ As stated, one consequence of the above is we can set $s=1$, and $m$ to nearly linear in $d$ as long as $\mu$ is at most inverse polylogarithmic in $d$. It is worth noting that if $d \gtrsim \log n$, then a random $d$-dimensional subspace $E$ has $\mu(E) \lesssim \sqrt{d/n}$ by the JL lemma, and thus most subspaces do have even much lower coherence. Note that the latter bound is sharp. Indeed, let $E$ be any $d$-dimensional subspace and let $U\in\mathbb{R}^{n\times d}$ have orthonormal columns such that $E$ is the column space of $U$. Then $$\max_{v\in E, \|v\|_2=1} \max_{1\le j\le n}|\inprod{v,e_j}|$$ is the maximum $\ell_2$ norm of a row $u_j$ of $U$. Since $U$ has $n$ rows and $\|U\|_F^2 = d$, there exists a $j$ such that $\|u_j\|_2\ge \sqrt{d/n}$ and in particular $\mu(E)\geq\sqrt{d/n}$. } \textup{ It is also worth noting that \cite[Theorem 3.3]{AMT10} observed that if $H\in\mathbb{R}^{n\times n}$ is a bounded orthonormal system (i.e.\ $H$ is orthogonal and $\max_{i,j} |H_{ij}| = O(1/\sqrt{n})$, and $D$ is a diagonal matrix with independent Rademacher entries, then $HDE$ has incoherence $\mu\lesssim \sqrt{d(\log n)/n}$ with probability $1 - 1/\mathrm{poly}(n)$. They then showed that incoherence $\mu$ implies that a sampling matrix $S\in\mathbb{R}^{m\times n}$ suffices to achieve \Equation{rjl-condition} with $m\lesssim n\mu^2\log m / \varepsilon^2$. Putting these two facts together implies $SHD$ gives \Equation{rjl-condition} with $m\lesssim d(\log n)(\log m)/\varepsilon^2$. \Equation{two-ten} combined with \cite[Theorem 3.3]{AMT10} implies a statement of a similar flavor but with an arguably stronger implication for applications: $\Phi HD$ with $s=1$ satisfies \Equation{rjl-condition} for $T = E\cap S^{n-1}$ without increasing $m$ as long as $n \gtrsim d(\log(d/\varepsilon))^c/\varepsilon^2$. } \end{remark} \section{The type-$2$ case}\SectionName{typetwo} Recall that the \emph{type-$2$ constant} $T_2(\|\cdot\|)$ of a semi-norm $\|\cdot \|$ is defined as the best constant (if it exists) such that the inequality \begin{equation} \mathbb{E}_g\Big\|\sum_\alpha g_\alpha x_\alpha\Big\| \le T_2\Big(\sum_\alpha \|x_\alpha\|^2\Big)^{1/2} \EquationName{3.1} \end{equation} holds, for all finite systems of vectors $\{x_{\alpha}\}$ and i.i.d.\ standard Gaussian $(g_\alpha)$. Given $T\subset S^{n-1}$, define the semi-norm $$ \tnorm{x}_T = \sup_{y\in T} |\inprod{x,y}| $$ It will be convenient to use the following notation. For $i=1,\ldots,m$ we let $J_i$ be the set of `active' indices, i.e., $J_i = \{j=1,\ldots,n : \delta_{ij} = 1\}$. We can then write \begin{equation} \vertiii{x} = \frac 1{\sqrt{s}} \max_{1\le i\le m} \|P_{J_i} x\|_2, \EquationName{3.2} \end{equation} where we abuse notation and let $P_{J_i}$ denote orthogonal projection onto the coordinate subspace specified by $J_i$. \begin{lemma} \label{lem:BoN410} Let $\|\cdot\|$ be any semi-norm satisfying $\tnorm{x}_T\leq \|x\|$ for all $x\in \mathbb{R}^n$. Then, \begin{align*} \gamma_2(T,\|\cdot\|_{\delta}) & \lesssim \frac{1}{\sqrt{s}} T_2(\|\cdot\|)(\log n)^3(\log m)^{1/2} \\ & \qquad \qquad \qquad \times \Big(\max_{1\leq i\leq m} \mathbb{E}_g\Big\|\sum_{j=1}^n g_j\delta_{ij} e_j\Big\| + (\log m)^{1/2}d_{\|\cdot\|}\Big(\bigcup_{i=1}^m B_{J_i}\Big)\Big), \end{align*} where $T_2(\|\cdot\|)$ is the type-$2$ constant and for $1\leq i\leq m$, $$B_{J_i} = \Big\{\sum_{j\in J_i} x_je_j : \sum_{j\in J_i} x_j^2 \le 1\Big\}.$$ \end{lemma} Note also that $d_{\delta}(T)$ can be bounded as \begin{align} \EquationName{delTermBd} d_{\delta}(T) & = \frac{1}{\sqrt{s}} \max_{1\leq i\leq m}\sup_{x\in T} \Big(\sum_{j=1}^n \delta_{ij}\langle x,e_j\rangle^2\Big)^{1/2} \nonumber \\ & = \frac{1}{\sqrt{s}} \max_{1\leq i\leq m}\sup_{x\in T} \Big(\mathbb{E}_g\Big|\sum_{j=1}^n \delta_{ij}g_j\langle x,e_j\rangle\Big|^2\Big)^{1/2} \nonumber \\ & \lesssim \frac{1}{\sqrt{s}} \max_{1\leq i\leq m}\sup_{x\in T} \mathbb{E}_g\Big|\Big\langle x,\sum_{j=1}^n \delta_{ij}g_je_j\Big\rangle\Big| \leq \frac{1}{\sqrt{s}} \max_{1\leq i\leq m}\mathbb{E}_g\Big\|\sum_{j=1}^n \delta_{ij}g_je_j\Big\|. \end{align} \begin{remark} \textup{ Replacing $\tnorm{\cdot}_T$ by a stronger norm $\|\cdot\| \ge \tnorm{\cdot}_T$ in this lemma may reduce the type-$2$ constant. The application to $k$-sparse vectors will illustrate this. } \end{remark} \begin{proof}[Proof of Lemma~\ref{lem:BoN410}] By \Equation{Dudley}, to bound $\gamma_2(T,\|\cdot\|_{\delta})$ it suffices to evaluate the covering numbers $\mathcal{N}(T, \|\cdot\|_{\delta}, t)$ for $t<d_{\delta}^*(T)/\sqrt{s}$, where we set $$d_{\delta}^*(T) = \sqrt{s}d_{\delta}(T) = \sup_{x\in T}\max_{1\le i\le m} \|P_{J_i} x\|_2.$$ For large values of $t$ we use duality for covering numbers. It will be convenient to work with $\tilde{T} = \mathrm{conv}(T\cup (-T))$, which is closed, bounded, convex and symmetric. Clearly, $$\mathcal{N}(T, \|\cdot\|_{\delta}, t)\leq \mathcal{N}(\tilde{T}, \|\cdot\|_{\delta}, t)$$ and in the notation of Appendix~\ref{sec:probTools}, $$\tnorm{x}_T = \sup_{y\in T\cup\{-T\}}\inprod{x,y} = \sup_{y\in \tilde{T}}\inprod{x,y} = \|x\|_{\tilde{T}^{\circ}}.$$ Here $\tilde{T}^{\circ}$ denotes the polar (see Appendix~\ref{sec:convex-tools} for a definition). Lemma~\ref{lem:BPST}, applied with $U=\tilde{T}$, $V=\cap_{i=1}^m \{x\in \mathbb{R}^n \ : \ s^{-1/2}\|P_{J_i}x\|_2\leq 1\}$, $\varepsilon=t$ and $\theta=d_{\delta}^*(T)/\sqrt{s}$, implies $$ \log \mathcal{N}(\tilde{T}, \|\cdot\|_{\delta}, t) \le \Big(\log\frac {d_{\delta}^*(T)}{t\sqrt{s}}\Big) A \log \mathcal{N}\Big(\frac 1{\sqrt{s}}\mathrm{conv}\Big(\bigcup_{i=1}^m B_{J_i}\Big), \tnorm{\cdot}_T, \frac t8\Big) $$ where $A \lesssim T_2(\|\cdot\|_{\delta})^2 \lesssim \log m$ (by \Eqsub{3.2}) and we used that $$B_{J_i} = \Big\{\sum_{j\in J_i} x_je_j : \sum_{j\in J_i} x_j^2 \le 1\Big\} = \{x\in \mathbb{R}^n \ : \ \|P_{J_i}x\|_2\leq 1\}^{\circ}.$$ Hence, \begin{equation} \log\mathcal{N}(\tilde{T}, \|\cdot\|_{\delta}, t) \lesssim \Big(\log\frac {d_{\delta}^*(T)}{t\sqrt{s}}\Big)(\log m)\log\mathcal{N}\Big(\mathrm{conv}\Big(\bigcup_{i=1}^m B_{J_i}\Big), \|\cdot\|, \frac 18\sqrt{s} t\Big) \EquationName{3.3} \end{equation} To estimate the right hand side we use the following consequence of Maurey's lemma (Lemma~\ref{lem:Maurey} of Appendix~\ref{sec:probTools}). \begin{claim}\ClaimName{claim} Fix $0<R<\infty$. Let $\|\cdot\|$ be a semi-norm on $\mathbb{R}^n$ and $\Omega\subset \{x\in \mathbb{R}^n : \|x\| \le R\}$. Then for $0 < \epsilon < R$, \begin{equation} \epsilon [\log \mathcal{N}(\mathrm{conv}(\Omega), \|\cdot\|, \epsilon)]^{1/2} \lesssim \Big(\log \frac {R}{\epsilon}\Big)^{3/2} T_2(\|\cdot\|) \max_{\epsilon\leq \epsilon'\leq R} \epsilon'[\log\mathcal{N}(\Omega, \|\cdot\|, \epsilon')]^{1/2} \EquationName{3.4} \end{equation} \end{claim} \begin{proof} It suffices to prove the result for $R=1$, the general case then follows by considering $R^{-1}\Omega$. For each positive integer $k$, let $\Omega_k\subset \Omega$ be a finite set satisfying $$ |\Omega_k| \le \mathcal{N}(\Omega, \|\cdot\|, 2^{-k}) $$ and $$ \inf_{y\in \Omega_k} \|x - y\| < 2^{-k}\hbox{ for all } x\in\Omega $$ Given $x\in\Omega$, denote $x_k\in\Omega_k$ a sequence s.t.\ $\|x - x_k\| < 2^{-k}$. Then, setting $y_0 = x_0$, $y_k = x_k - x_{k-1}$, we obtain \begin{equation} \Big\|x - \sum_{\epsilon/2<2^{-k}\leq 1} y_k\Big\| < \epsilon \EquationName{3.5} \end{equation} \begin{equation} y_k \in \Omega_k - \Omega_{k-1}\hbox{ and } \|y_k\| < 2^{-k} + 2^{-k + 1} = 3\cdot 2^{-k} \EquationName{3.6} \end{equation} It follows from \Equation{3.5} that $$ \mathrm{conv}(\Omega) \subset \sum_{\epsilon/2<2^{-k}\leq 1} \mathrm{conv}(\tilde{\Omega}_k) + B_{\|\cdot\|}(0, \epsilon) $$ with \begin{equation} \tilde{\Omega}_k = \Big\{y\in \Omega_k - \Omega_{k-1} : \|y\| < 3\cdot 2^{-k} \Big\} \EquationName{3.7} \end{equation} By Maurey's lemma, given $z\in\mathrm{conv}(\tilde{\Omega}_k)$ and a positive integer $\ell_k$, there are points $y_1,\ldots,y_\ell\in\tilde{\Omega}_k$ such that \begin{equation} \Big\|z - \frac 1{\ell_k}(y_1 + \ldots + y_{\ell_k})\Big\| \lesssim T_2(\|\cdot\|)\frac 1{\sqrt{\ell_k}} \cdot 3\cdot 2^{-k} \EquationName{3.8} \end{equation} Taking $\ell_k\simeq T_2(\|\cdot\|)^2 4^{-k} \epsilon^{-2} \log^2(1/\epsilon)$, we obtain a point $z_k\in \Omega_k' = \frac 1{\ell_k}(\tilde{\Omega}_k + \ldots + \tilde{\Omega}_k)$ such that $$ \|z - z_k\| < \frac{\epsilon}{2\log(1/\epsilon)} $$ Moreover $$ \log|\Omega_k'| \le \ell_k \log|\tilde{\Omega}_k| \lesssim T_2(\|\cdot\|)^2 4^{-k} \epsilon^{-2}\log^2(1/\epsilon) \log \mathcal{N}(\Omega, \|\cdot\|, 2^{-k}) $$ Since $|\{k \ : \ \epsilon/2<2^{-k}\leq 1\}|\leq \log(2 /\epsilon)$, we obtain from the preceding, $$ \log\mathcal{N}(\mathrm{conv}(\Omega), \|\cdot\|, \epsilon) \lesssim \log^2(1/\epsilon) T_2(\|\cdot\|)^2 \epsilon^{-2} \sum_{\epsilon/2<2^{-k}\leq 1} 4^{-k} \log\mathcal{N}(\Omega, \|\cdot\|, 2^{-k}) $$ and \Equation{3.4} follows. This proves the claim. \end{proof} Observe that \begin{align*} d_{\delta}^*(T) & = \sup_{x\in T} \max_{1\leq i\leq m} \|P_{J_i}x\|_2 \\ & = \sup_{x\in T}\sup_{y\in B_{\ell_2^n}} \max_{1\leq i\leq m} \langle x,P_{J_i}y\rangle = d_{\tnorm{\cdot}_T}\Big(\bigcup_{i=1}^m B_{J_i}\Big) \leq d_{\|\cdot\|}\Big(\bigcup_{i=1}^m B_{J_i}\Big) \end{align*} Write $d_{\|\cdot\|}:=d_{\|\cdot\|}(\cup_{i=1}^m B_{J_i})$ for brevity. Applying \Claim{claim} with $\Omega = \cup_{i=1}^m B_{J_i}$, $\epsilon=t\sqrt{s}$, and the dominating semi-norm $\|\cdot\|\geq\tnorm{\cdot}_T$ to the right hand side of \Equation{3.3}, we find \begin{align} \nonumber t\sqrt{s}\Big[\log\mathcal{N}(\tilde{T}, \|\cdot\|_{\delta}, t)\Big]^{1/2} \le &\Big(\log \frac {d_{\|\cdot\|}}{t\sqrt{s}}\Big)^2(\log m)^{1/2} T_2(\|\cdot\|)\cdot\\ {}&\Big[\max_{t\sqrt{s}\leq t'\leq d_{\|\cdot\|}} t'\Big[\log \mathcal{N}\Big(\bigcup_{i=1}^m B_{J_i}, \|\cdot\|, t'\Big)\Big]^{1/2}\Big]. \EquationName{3.9} \end{align} Using the dual Sudakov inequality (Lemma~\ref{lem:dualSudakov} of Appendix~\ref{sec:probTools}) we arrive at \begin{align} \nonumber \Big[\log\mathcal{N}(\tilde{T}, \|\cdot\|_{\delta}, t)\Big]^{1/2} \le &\frac {1}{t\sqrt{s}}\Big(\log \frac {d_{\|\cdot\|}}{t\sqrt{s}}\Big)^2(\log m)^{1/2} T_2(\|\cdot\|)\cdot\\ {}&\Big[\max_{1\le i\le m} \mathbb{E}_g\Big\|\sum_{j\in J_i} g_j e_j\Big\| + d_{\|\cdot\|}(\log m)^{1/2}\Big]. \EquationName{dualsud} \end{align} We now apply \Equation{dualsud} for $(sn)^{-1/2}d_{\|\cdot\|}\leq t<s^{-1/2}d_{\|\cdot\|}$ and the estimate (cf.\ \Equation{volComp}) $$\mathcal{N}(T,\|\cdot\|_{\delta},t)\leq \mathcal{N}(d_{\delta}(T)B_{\|\cdot\|_{\delta}},\|\cdot\|_{\delta},t)\leq \Big(1+\frac{2d_{\delta}^*(T)}{t\sqrt{s}}\Big)^n \leq \Big(1+\frac{2d_{\|\cdot\|}}{t\sqrt{s}}\Big)^n$$ for $0<t<(sn)^{-1/2}d_{\|\cdot\|}$, respectively, in \Equation{Dudley}. A straightforward computation, similar to \Equation{splittingArg}, yields the result. \end{proof} \subsection{Application to $k$-sparse vectors}\SectionName{ksparse} Let $\|x\|_0$ denote the number of non-zero entries of $x$ and set $$ S_{n,k} = \{x\in\mathbb{R}^n : \|x\|_2 = 1, \|x\|_0 \le k \}. $$ \begin{theorem} Let $A\in\mathbb{R}^{n\times n}$ be orthogonal and denote $$ T = A(S_{n,k}) $$Let $\eta\leq 1/m$. If \begin{equation} \max |A_{ij}| < (k\log n)^{-1/2} \EquationName{4.8} \end{equation} and \begin{align} \nonumber m &\gtrsim k(\log n)^8(\log m) / \varepsilon^2 \\ s &\gtrsim (\log n)^7 (\log m)(\log \eta^{-1}) /\varepsilon^2, \EquationName{4.7} \end{align} then with probability at least $1-\eta$ we have $$(1-\varepsilon)\|x\|_2^2 \leq \|\Phi x\|_2^2 \leq (1+\varepsilon)\|x\|_2^2 \qquad \mathrm{for \ all} \ x\in T.$$ \end{theorem} \begin{proof} We apply \Equation{RIPLpEst} and estimate $(\mathbb{E}_{\delta}\gamma_2^{2p}(T,\|\cdot\|_{\delta}))^{1/p}$ and $\mathbb{E}_{\delta} d_{\delta}^{2p}(T)$ for $p\geq \log m$. We trivially estimate $\mathbb{E}_{\delta} d_{\delta}^{2p}(T)\leq 1/s$. To bound the $\gamma_2$-functional, note that $T\subset K$, where \begin{equation} K = B_{\ell_2^n} \cap \sqrt{k} A(B_{\ell_1^n}). \EquationName{4.1} \end{equation} The polar body of $K$ is given by $$ K^\circ = B_{\ell_2^n} + \frac 1{\sqrt{k}} A(B_{\ell_{\infty}^n}) $$ and one can readily calculate that $T_2(\tnorm{\cdot}_K) \lesssim \sqrt{\log n}$ and $d_{\tnorm{\cdot}_K}(\cup_{i=1}^m B_{J_i})\leq 1$. We apply Lemma~\ref{lem:BoN410} with $\|\cdot\|=\tnorm{\cdot}_K$. Note that for every $1\leq i\leq m$, $$\mathbb{E}_g\btnorm{\sum_j g_j\delta_{ij}e_k}_K \leq \sqrt{k}\mathbb{E}_g\max_{1\leq \ell\leq n}\Big|\sum_j g_j\delta_{ij} A_{j\ell}\Big| \leq \sqrt{k\log n} \max_{1\leq\ell\leq n}\Big(\sum_j \delta_{ij}A_{j\ell}^2\Big)^{1/2}$$ and therefore the lemma implies that \begin{align*} & \gamma_2(T,\|\cdot\|_{\delta}) \\ & \qquad \leq \frac{1}{\sqrt{s}}(\log n)^{7/2}\log m + \sqrt{\frac{k}{s}}(\log n)^4(\log m)^{1/2} \max_{1\leq i\leq m,1\leq \ell\leq n} \Big(\sum_j \delta_{ij}A_{j\ell}^2\Big)^{1/2}. \end{align*} As a consequence, as $p\geq \log m$, \begin{align} \EquationName{ga2pSparse} & (\mathbb{E}_{\delta}\gamma_2^{2p}(T,\|\cdot\|_{\delta}))^{1/p} \nonumber\\ & \ \ \ \leq \frac{1}{s}(\log n)^{7}(\log m)^2 + \frac{k}{s}(\log n)^8(\log m) \max_{1\leq i\leq m}\Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq n} \Big(\sum_j \delta_{ij}A_{j\ell}^2\Big)^{p}\Big)^{1/p}. \end{align} Since $\mathbb{E}\delta_{ij}=s/m$ and $A$ is orthogonal, we obtain using symmetrization and Khintchine's inequality, \begin{align*} & \Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq n} \Big(\sum_j \delta_{ij}A_{j\ell}^2\Big)^{p}\Big)^{1/p} \\ & \qquad \leq \frac{s}{m} + \Big(\mathbb{E}_r\mathbb{E}_{\delta}\max_{1\leq \ell\leq n} \Big(\sum_j r_j\delta_{ij}A_{j\ell}^2\Big)^{p}\Big)^{1/p} \\ & \qquad \leq \frac{s}{m} + \sqrt{p}\Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq n} \Big(\sum_j \delta_{ij}A_{j\ell}^4\Big)^{p/2}\Big)^{1/p} \\ & \qquad \leq \frac{s}{m} + \sqrt{p}\max_{j,\ell}|A_{j\ell}|\Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq n} \Big(\sum_j \delta_{ij}A_{j\ell}^2\Big)^{p}\Big)^{1/(2p)}. \end{align*} By solving this quadratic inequality, we find $$\Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq n} \Big(\sum_j \delta_{ij}A_{j\ell}^2\Big)^{p}\Big)^{1/p} \lesssim \frac{s}{m} + p\max_{j,\ell}A_{j\ell}^2 \leq \frac{s}{m} + \frac{p}{k\log n}.$$ Combine this bound with \Equation{ga2pSparse} and \Equation{RIPLpEst} to arrive at \begin{align*} & \Big(\mathbb{E}_{\delta,\sigma}\sup_{x \in T} \Big|\|\Phi x\|_2^2 - \|x\|_2^2\Big|^p\Big)^{1/p} \\ & \ \ \lesssim \frac{1}{s}(\log n)^{7}(\log m)^2 + \frac{k}{m}(\log n)^8(\log m) + \frac{p}{s}(\log n)^7(\log m) \\ & \ \ \ \ \ \ + \Big(\frac{1}{s}(\log n)^{7}(\log m)^2 + \frac{k}{m}(\log n)^8(\log m) + \frac{p}{s}(\log n)^7(\log m)\Big)^{1/2} \\ & \ \ \ \ \ \ \ \ \qquad \qquad + \sqrt{\frac{p}{s}} + \frac{p}{s}. \end{align*} Since $p\geq \log m$ was arbitrary, the result now follows from Lemma~\ref{lem:MomentsToTails}. \end{proof} \section{Sketching constrained least squares programs}\SectionName{cls} Let us now apply the previous results to constrained least squares minimization. We refer to Appendix~\ref{sec:convex-tools} for any unexplained terminology. Consider $A\in \mathbb{R}^{n\times d}$ and a sketching matrix $\Phi\in \mathbb{R}^{m\times n}$. Define $f(x) = \|Ax - b\|_2^2$ and $g(x) = \|\Phi Ax - \Phi b\|_2^2$. Let $\mathcal{C}\subset \mathbb{R}^d$ be any closed convex set. Let $x_*$ be a minimizer of the constrained least squares program \begin{equation} \label{eqn:CLS} \min f(x) \qquad \mathrm{subject \ to} \qquad x\in \mathcal{C} \end{equation} and let $\hat{x}$ be a minimizer of the associated sketched program \begin{equation} \label{eqn:SCLS} \min g(x) \qquad \mathrm{subject \ to} \qquad x\in \mathcal{C}. \end{equation} Before giving our analysis, we define two quantities introduced in \cite{PiW14}. Given $\mathcal{K}\subset\mathbb{R}^d$ and $u\in S^{n-1}$ we set \begin{align*} Z_1(A,\Phi,\mathcal{K}) & = \inf_{v\in A\mathcal{K}\cap S^{n-1}} \|\Phi v\|_2^2 \\ Z_2(A,\Phi,\mathcal{K},u) & = \sup_{v \in A\mathcal{K}\cap S^{n-1}} |\langle \Phi u, \Phi v\rangle - \langle u,v\rangle|. \end{align*} We denote the tangent cone of $\mathcal{C}$ at a point $x$ by $T_{\mathcal{C}}(x)$ (see Appendix~\ref{sec:convex-tools} for a definition). The first statement in the following lemma is \cite[Lemma 1]{PiW14}. For the convenience of the reader we provide a proof, which is in essence the same as in \cite{PiW14}, with the second case (when $x_*$ is a global minimizer) being a slight modification of the proof there. In what follows we assume $Ax_*\neq b$, since otherwise $f(\hat{x}) = f(x_*) = 0$. \begin{lemma} \label{lem:compare} Define $u=(Ax_*-b)/\|Ax_*-b\|_2$ and set $Z_1=Z_1(A,\Phi,T_{\mathcal{C}}(x_*))$ and $Z_2=Z_2(A,\Phi,T_{\mathcal{C}}(x_*),u)$. Then, $$f(\hat{x})\leq \Big(1+\frac{Z_2}{Z_1}\Big)^2f(x_*).$$ If $x_*$ is a global minimizer of $f$, then $$f(\hat{x})\leq \Big(1+\frac{Z_2^2}{Z_1^2}\Big)f(x_*).$$ \end{lemma} \begin{proof} Set $e=\hat{x}-x_* \in T_{\mathcal{C}}(x_*)$ and $w=b-Ax_*$. By optimality of $x_*$, for any $x\in \mathcal{C}$, $$\langle\nabla f(x_*),x-x_*\rangle = \langle Ax_*-b,A(x-x_*)\rangle\geq 0.$$ In particular, taking $x=\hat{x}$ gives $\langle w,Ae\rangle\leq 0$. As a consequence, \begin{equation} \label{eqn:compareZ2} \Big\langle \frac{w}{\|w\|_2},\frac{\Phi^*\Phi Ae}{\|Ae\|_2}\Big\rangle \leq \Big\langle \frac{w}{\|w\|_2},\frac{\Phi^*\Phi Ae}{\|Ae\|_2}\Big\rangle - \Big\langle \frac{w}{\|w\|_2},\frac{Ae}{\|Ae\|_2}\Big\rangle \leq Z_2. \end{equation} By optimality of $\hat{x}$, for any $x\in \mathcal{C}$, $$\langle\nabla g(\hat{x}),x-\hat{x}\rangle = \langle \Phi A\hat{x}-\Phi b,\Phi A(x-\hat{x})\rangle\geq 0.$$ Taking $x=x_*$ yields $$0\geq \langle \Phi A\hat{x} - \Phi b,\Phi Ae\rangle = \langle -\Phi w,\Phi Ae\rangle + \|\Phi Ae\|_2^2.$$ Therefore, by (\ref{eqn:compareZ2}), $$Z_1\|Ae\|_2^2 \leq \|\Phi Ae\|_2^2 \leq \langle \Phi w,\Phi Ae\rangle \leq Z_2\|w\|_2\|Ae\|_2.$$ In other words, $$\|Ae\|_2\leq \frac{Z_2}{Z_1}\|w\|_2$$ and by Cauchy-Schwarz, $$\|A\hat{x}-b\|_2^2 = \|w\|_2^2 + \|Ae\|_2^2 - 2\langle w,Ae\rangle \leq \|w\|_2^2\Big(1+\frac{Z_2}{Z_1}\Big)^2.$$ If $x_*$ is a global minimizer of $f$, then $\nabla f(x_*)=0$ and therefore $\langle w,Ae\rangle=0$ (a similar observation to \cite[Theorem 12]{Sarlos06}). This yields the second assertion. \end{proof} Clearly, if $\Phi$ satisfies \Eqsub{jl-condition} for $T=A T_{\mathcal{C}}(x_*)\cap S^{n-1}$ then $Z_1\geq 1-\varepsilon$. We do not immediately obtain an upper bound for $Z_2$, however, as $u$ is in general not in $A T_{\mathcal{C}}(x_*)\cap S^{n-1}$. We mend this using the observation in Lemma~\ref{lem:Z2boundGen}. For its proof we recall a chaining result from \cite{Dir13}. Let $(T,\rho)$ be a semi-metric space. Recall that a real-valued process $(X_t)_{t\in T}$ is called subgaussian if for all $s,t\in T$, \begin{equation*} \mathbb{P}(|X_t - X_s|\geq w\rho(t,s)) \leq 2\exp(-w^2) \qquad (w\geq 0). \end{equation*} The following result is a special case of \cite[Theorem 3.2]{Dir13}. \begin{lemma} \label{lem:supSubg} If $(X_t)_{t\in T}$ is subgaussian, then for any $t_0\in T$ and $1\leq p<\infty$, \begin{equation*} \Big(\mathbb{E}\sup_{t\in T}|X_t - X_{t_0}|^p\Big)^{1/p} \lesssim \gamma_2(T,\rho) + \sqrt{p}d_{\rho}(T). \end{equation*} \end{lemma} In particular, if $T$ contains $n$ elements, then \begin{equation*} \Big(\mathbb{E}\sup_{t\in T}|X_t - X_{t_0}|^p\Big)^{1/p} \lesssim (\sqrt{p} + (\log n)^{1/2})d_{\rho}(T). \end{equation*} \begin{lemma} \label{lem:Z2boundGen} Fix $u\in B_{\ell_2^n}$, $T\subset \mathbb{R}^n$ and let $\Phi$ be the SJLT. Set $$Z = \sup_{v\in T}|\langle \Phi u,\Phi v\rangle - \langle u,v\rangle|.$$ For any $p\geq 1$, $$(\mathbb{E}_{\delta,\sigma}Z^p)^{1/p} \lesssim \Big(\sqrt{\frac{p}{s}} + 1\Big)\Big((\mathbb{E}_{\delta}\gamma_2^p(T,\|\cdot\|_{\delta}))^{1/p} + \sqrt{p}(\mathbb{E}_{\delta}d_{\delta}^p(T))^{1/p}\Big).$$ \end{lemma} \begin{proof} With the notation (\ref{eqn:AdelxDef}) we have $\Phi x=A_{\delta,x}\sigma$ for any $x\in \mathbb{R}^n$. Since $\mathbb{E}_{\sigma} \Phi^*\Phi=I$ we find \begin{align*} Z = \sup_{v\in T}|\langle u,(\Phi^*\Phi-I)v\rangle| = \sup_{v\in T} |\sigma^* A_{\delta,u}^*A_{\delta,v}\sigma - \mathbb{E}_{\sigma}(\sigma^* A_{\delta,u}^*A_{\delta,v}\sigma)|. \end{align*} Let $\sigma'$ be an independent copy of $\sigma$. By decoupling (see (\ref{eqn:decRad})), $$(\mathbb{E}_{\sigma}Z^p)^{1/p} \leq 4 (\mathbb{E}_{\sigma,\sigma'}\sup_{v\in T} |\sigma^* A_{\delta,u}^*A_{\delta,v}\sigma'|^p)^{1/p}.$$ For any $v_1,v_2\in T$, Hoeffding's inequality implies that for all $w\geq 0$ \begin{align*} & \mathbb{P}_{\sigma'}(|\sigma^* A_{\delta,u}^*A_{\delta,v_1}\sigma' - \sigma^* A_{\delta,u}^*A_{\delta,v_2}\sigma'| \\ & \qquad \qquad \qquad \qquad \geq \sqrt{w}\|\sigma^*A_{\delta,u}^*\|_2 \ \|A_{\delta,v_1}-A_{\delta,v_2}\|_{\ell_2\rightarrow\ell_2}) \leq 2e^{-w} \end{align*} and therefore $v\mapsto \sigma^* A_{\delta,u}^*A_{\delta,v}\sigma'$ is a subgaussian process. By Lemma~\ref{lem:supSubg} (and (\ref{eqn:AdelxNorm})), \begin{align*} & (\mathbb{E}_{\sigma'}\sup_{v\in T} |\sigma^* A_{\delta,u}^*A_{\delta,v}\sigma'|^p)^{1/p} \\ & \qquad \lesssim \|A_{\delta,u}\sigma\|_2 \gamma_2(T,\|\cdot\|_{\delta}) + \sqrt{p}\|A_{\delta,u}\sigma\|_2 d_{\delta}(T). \end{align*} Taking the $L_p(\Omega_{\sigma})$-norm on both sides and using Lemma~\ref{lem:Lpl2} of Appendix~\ref{sec:probTools} we find \begin{align*} (\mathbb{E}_{\sigma}Z^p)^{1/p} & \lesssim (\mathbb{E}_{\sigma}\|A_{\delta,u}\sigma\|_2^p)^{1/p}\Big(\gamma_2(T,\|\cdot\|_{\delta}) + \sqrt{p}d_{\delta}(T)\Big) \\ & \lesssim (\sqrt{p}\|A_{\delta,u}\| + \|A_{\delta,u}\|_F)\Big(\gamma_2(T,\|\cdot\|_{\delta}) + \sqrt{p}d_{\delta}(T)\Big) \\ & \leq \Big(\sqrt{\frac{p}{s}} + 1\Big)\Big(\gamma_2(T,\|\cdot\|_{\delta}) + \sqrt{p}d_{\delta}(T)\Big). \end{align*} Now take the $L_p(\Omega_{\delta})$-norm on both sides to obtain the result. \end{proof} \subsection{Unconstrained case} We first consider unconstrained least squares minimization, i.e., the constraint set is $\mathcal{C}=\mathbb{R}^d$. \begin{theorem} \label{thm:LSuc} Let $\mathrm{Col}(A)$ be the column space of $A$ and let $$\mu(A)=\max_{1\leq j\leq n} \|P_{\mathrm{Col}(A)}e_j\|_2$$ be its largest leverage score. Set $\mathcal{C}=\mathbb{R}^d$ and let $x_*$ and $\hat{x}$ be minimizers of (\ref{eqn:CLS}) and (\ref{eqn:SCLS}), respectively. Let $r(A)$ be the rank of $A$. Suppose that \begin{align*} m &\gtrsim \varepsilon^{-1}((\log \eta^{-1})(r(A) + \log m)(\log m)^2 + (\log \eta^{-1})^2 r(A)) \\ s &\gtrsim \varepsilon^{-1}\mu(A)^2((\log \eta^{-1})^2 + (\log \eta^{-1})(\log m)^3) \\ & \qquad \qquad + \varepsilon^{-1/2}\mu(A)((\log \eta^{-1})^{3/2} + (\log \eta^{-1})(\log m)^{3/2}) \end{align*} and $\eta\leq \frac{1}{m}$. Then, with probability at least $1-\eta$, \begin{equation} \label{eqn:LSuc} f(\hat{x})\leq (1-\varepsilon)^{-2}f(x_*). \end{equation} \end{theorem} \begin{proof} Set $Z_1=Z_1(A,\Phi,T_{\mathcal{C}}(x_*))$ and $Z_2=Z_2(A,\Phi,T_{\mathcal{C}}(x_*),u)$. Observe that $T_{\mathcal{C}}(x_*)=\mathbb{R}^d$. Applying Theorem~\ref{thm:JLsubspace} for the $r(A)$-dimensional subspace $E=\mathrm{Col}(A)$ immediately yields that $Z_1\geq 1-\sqrt{\varepsilon}$ with probability at least $1-\eta$. Setting $T=A T_{\mathcal{C}}(x_*)\cap S^{n-1}=\mathrm{Col}(A)\cap S^{n-1}$ in Lemma~\ref{lem:Z2boundGen} and subsequently using \Eqsub{ga2Subs} and \Eqsub{diamSubs} yields, for any $p\geq \log(m)$, \begin{align*} (\mathbb{E}_{\delta,\sigma} Z_2^p)^{1/p} & \lesssim \Big(\sqrt{\frac{p}{s}} + 1\Big)\Big((\mathbb{E}_{\delta}\gamma_2^p(T,\|\cdot\|_{\delta}))^{1/p} + \sqrt{p}(\mathbb{E}_{\delta}d_{\delta}^p(T))^{1/p}\Big) \\ & \lesssim \Big(\sqrt{\frac{p}{s}} + 1\Big)\Big[\frac{(\sqrt{r(A)}+\log^{1/2}(m))\log(m)}{\sqrt{m}} + \sqrt{\frac{pr(A)}{m}} \\ & \qquad \qquad \qquad \qquad \qquad \qquad + \frac{\sqrt{p}\log^{3/2}(m)}{\sqrt{s}}\mu(A) + \frac{p}{\sqrt{s}}\mu(A)\Big]. \end{align*} Applying Lemma~\ref{lem:MomentsToTails} (and setting $w=\log(\eta^{-1})$ in (\ref{eqn:MomentsToTails})) shows that $Z_2\leq \sqrt{\varepsilon}$ with probability at least $1-\eta$. The second statement in Lemma~\ref{lem:compare} now implies that with probability at least $1-2\eta$, $$f(\hat{x})\leq \Big(1+\frac{\varepsilon}{(1-\sqrt{\varepsilon})^2}\Big)f(x_*) \leq (1-\varepsilon)^{-2}f(x_*).$$ To see the last inequality, note that it is equivalent to $$-\varepsilon+4\varepsilon^{3/2}-3\varepsilon^2-2\varepsilon^{5/2}+2\varepsilon^3\leq 0.$$ Clearly this is satisfied for $\varepsilon$ small enough (since then $-\varepsilon$ is the leading term). In fact, one can verify by calculus that this holds for $\varepsilon\leq 1/10$, which we may assume without loss of generality. \end{proof} If $\Phi$ is a full random sign matrix (i.e., $s=m$) then it follows from our proof that (\ref{eqn:LSuc}) holds with probability at least $1-\eta$ if $$m\gtrsim \varepsilon^{-1}(r(A)+\log(\eta^{-1})).$$ This bound on $m$ is new, and was also recently shown in \cite{Woodruff14}. Previous works \cite{Sarlos06,KN14} allowed either $m \gtrsim \varepsilon^{-2}(r(A) + \log(\eta^{-1}))$ or $m\gtrsim \varepsilon^{-1} r(A)\log(\eta^{-1})$. Theorem~\ref{thm:LSuc} substantially improves $s$ while maintaining $m$ up to logarithmic factors. \subsection{$\ell_{2,1}$-constrained case} \label{sec:l21Constraint} Throughout this section, we set $d=bD$. For $x=(x_{B_1},\ldots,x_{B_b})\in \mathbb{R}^{d}$ consisting of $b$ blocks, each of dimension $D$, we define its $\ell_{2,1}$-norm by \begin{equation} \label{eqn:l21NormDef} \|x\|_{2,1}:= \|x\|_{\ell_1^b(\ell_2^D)} = \sum_{\ell=1}^b \|x_{B_{\ell}}\|_2. \end{equation} The corresponding dual norm is $$\|x\|_{2,\infty} := \|x\|_{\ell_{\infty}^b(\ell_2^D)} = \max_{1\leq \ell\leq b} \|x_{B_{\ell}}\|_2.$$ In this section we study the effect of sketching on the $\ell_{2,1}$-constrained least squares program $$\min \|Ax-b\|_2^2 \qquad \mathrm{subject \ to} \qquad \|x\|_{2,1}\leq R,$$ which is Eq.~(\ref{eqn:CLS}) for $\mathcal{C}=\{x\in \mathbb{R}^d \ : \ \|x\|_{2,1}\leq R\}$. In the statistics literature, this is called the \emph{group Lasso} (with non-overlapping groups of equal size). The $\ell_{2,1}$-constraint encourages a block sparse solution, i.e., a solution which has few blocks containing non-zero entries. We refer to e.g.\ \cite{BuG11} for more information. In the special case $D=1$ the program reduces to $$\min \|Ax-b\|_2^2 \qquad \mathrm{subject \ to} \qquad \|x\|_1\leq R,$$ which is the well-known \emph{Lasso} \cite{Tibshirani96}.\par To formulate our results we consider two norms on $\mathbb{R}^{n\times d}$, given by \begin{equation} \label{eqn:blockNormA} \tnorm{A} := \max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n\sum_{k\in B_{\ell}}|A_{jk}|^2\Big)^{1/2} \end{equation} and \begin{equation*} \|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}} = \max_{1\leq j\leq n}\max_{1\leq \ell\leq b}\|(A_{jk})_{k\in B_{\ell}}\|_2. \end{equation*} \begin{lemma} \label{lem:gamma2pEst} Let $A\in\mathbb{R}^{n\times d}$ and consider any set $\mathcal{K}\subset\mathbb{R}^d$. If $p\geq \log m$, then \begin{align*} & (\mathbb{E}_{\delta}\gamma_2^{2p}(A\mathcal{K}\cap S^{n-1},\|\cdot\|_{\delta}))^{1/p} \\ & \qquad \lesssim \alpha\Big[\sup_{x\in \mathcal{K} \ : \ \|Ax\|_2=1}\|x\|_{2,1}^2\Big] \Big(\frac{1}{s}(\log b + p)\|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}}^2 + \frac{1}{m} \tnorm{A}^2\Big). \end{align*} where $\alpha = (\log n)^6(\log m)^2(\log b)^2$. \end{lemma} \begin{proof} We deduce the assertion from Lemma~\ref{lem:BoN410}. By H\"{o}lder's inequality, $$\sup_{y\in A\mathcal{K}\cap S^{n-1}}\langle x,y\rangle = \sup_{v \in \mathcal{K} \ : \ \|Av\|_2=1} \langle A^*x,v\rangle\leq \|A^*x\|_{2,\infty} \sup_{v\in \mathcal{K} \ : \ \|Av\|_2=1}\|v\|_{2,1}.$$ We define $\|x\|$ to be the expression on the right hand side. Observe that for any sequence $(x_k)_{k\geq 1}$ in $\mathbb{R}^n$, \begin{align*} \mathbb{E}_g\Big\|\sum_{k\geq 1} g_k A^*x_k\Big\|_{2,\infty} & = \mathbb{E}_g\max_{1\leq \ell\leq b}\Big\|\sum_{k\geq 1} g_k(A^*x_k)_{B_{\ell}}\Big\|_2 \\ & \leq (\log b)^{1/2} \max_{1\leq \ell\leq b} \Big(\sum_{k\geq 1} \|(A^*x_k)_{B_{\ell}}\|_2^2\Big)^{1/2} \\ & \leq (\log b)^{1/2} \Big(\sum_{k\geq 1} \|A^*x_k\|_{2,\infty}^2\Big)^{1/2} \end{align*} and therefore $T_2(\|\cdot\|)\leq (\log b)^{1/2}$. Similarly, \begin{align} \EquationName{gamma2pFirst} \mathbb{E}_g\Big\|\sum_{j=1}^n g_j\delta_{ij} A^*e_j\Big\|_{2,\infty} & = \mathbb{E}_g\max_{1\leq \ell\leq b}\Big\|\Big(\sum_{j=1}^n g_j\delta_{ij}A_{jk}\Big)_{k\in B_{\ell}}\Big\|_2 \nonumber \\ & \leq (\log b)^{1/2} \max_{1\leq \ell\leq b} \Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^{1/2}. \end{align} Finally, \begin{equation*} d_{\|\cdot\|}\Big(\bigcup_{i=1}^m B_{J_i}\Big) = \sup_{v\in \mathcal{K} \ : \ \|Av\|_2=1}\|v\|_{2,1} \max_{1\leq i\leq m} \sup_{x\in B_{J_i}} \|A^*x\|_{2,\infty} \end{equation*} and for any $1\leq i\leq m$ and $x\in B_{J_i}$, \begin{align*} \|A^*x\|_{2,\infty} & = \max_{1\leq \ell\leq b}\Big\|\sum_{j=1}^n x_j\delta_{ij}(A_{jk})_{k\in B_{\ell}}\Big\|_2 \\ & \leq \max_{1\leq \ell\leq b}\sum_{j=1}^n |x_j| \ \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2 \leq \max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^{1/2}. \end{align*} We now apply Lemma~\ref{lem:BoN410} for $T=A\mathcal{K}\cap S^{n-1}$ and subsequently take $L_p(\Omega_{\delta})$-norms to conclude that for any $p\geq \log m$, \begin{align} \label{eqn:ga2pIntermediate} & (\mathbb{E}\gamma_2^{2p}(A\mathcal{K}\cap S^{n-1},\|\cdot\|_{\delta}))^{1/p} \nonumber\\ & \qquad \lesssim \frac{\alpha}{s}\Big[\sup_{x\in \mathcal{K} \ : \ \|Ax\|_2=1}\|x\|_{2,1}^2\Big] \max_{1\leq i\leq m} \Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^p\Big)^{1/p}. \end{align} Since $\mathbb{E}\delta_{ij} = \frac{s}{m}$ and $(\delta_{ij})_{j=1}^n$ is independent for any fixed $i$, we obtain by symmetrization (see \Equation{symmetrization}) for $(r_j)$ a vector of independent Rademachers \begin{align*} & \Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^p\Big)^{1/p} \\ & \qquad \leq \frac{s}{m}\max_{1\leq \ell\leq b} \sum_{j=1}^n\|(A_{jk})_{k\in B_{\ell}}\|_2^2 + 2\Big(\mathbb{E}_{\delta}\mathbb{E}_r\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n r_j\delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^p\Big)^{1/p}. \end{align*} By Khintchine's inequality, \begin{align*} & \Big(\mathbb{E}_{\delta}\mathbb{E}_r\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n r_j\delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^p\Big)^{1/p} \\ & \qquad \lesssim ((\log b)^{1/2} + p^{1/2}) \Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^4\Big)^{p/2}\Big)^{1/p} \\ & \qquad \leq ((\log b)^{1/2} + p^{1/2}) \Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^{p}\Big)^{1/(2p)} \\ & \qquad \qquad \qquad \qquad \qquad \qquad \qquad \times \max_{1\leq \ell\leq b}\max_{1\leq j\leq n} \|(A_{jk})_{k\in B_{\ell}}\|_2. \end{align*} Putting the last two displays together we find a quadratic inequality. By solving it, we arrive at \begin{align} \EquationName{gamma2pLast} & \Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^p\Big)^{1/p} \nonumber\\ & \qquad \lesssim \frac{s}{m} \max_{1\leq \ell\leq b} \sum_j \|(A_{jk})_{k\in B_{\ell}}\|_2^2 + (\log b + p) \max_{1\leq \ell\leq b}\max_{1\leq j\leq n}\|(A_{jk})_{k\in B_{\ell}}\|_2^2. \end{align} Combining this estimate with (\ref{eqn:ga2pIntermediate}) yields the assertion. \end{proof} \begin{theorem} \label{thm:SJLmain} Let $\mathcal{C}\subset\mathbb{R}^d$ be a closed convex set and let $\alpha=(\log n)^6(\log m)^2(\log b)^2$. Let $x_*$ and $\hat{x}$ be minimizers of (\ref{eqn:CLS}) and (\ref{eqn:SCLS}), respectively. Set $$d_{2,1} = \sup_{x\in T_{\mathcal{C}}(x_*) \ : \ \|Ax\|_{2,1}=1}\|x\|_{2,1}.$$ Suppose that \begin{align*} m & \gtrsim \varepsilon^{-2}\log(\eta^{-1})(\alpha + \log(\eta^{-1})\log b) \tnorm{A}^2 d_{2,1}^2,\\ s & \gtrsim \varepsilon^{-2}\log(\eta^{-1})(\log b + \log(\eta^{-1}))(\alpha + \log(\eta^{-1})\log b) \|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}}^2 d_{2,1}^2, \end{align*} and $\eta\leq \frac{1}{m}$. Then, with probability at least $1-\eta$, $$f(\hat{x})\leq (1-\varepsilon)^{-2}f(x_*).$$ \end{theorem} \begin{proof} Set $T=A T_{\mathcal{C}}(x_*)\cap S^{n-1}$, $Z_1=Z_1(A,\Phi,T_{\mathcal{C}}(x_*))$ and $Z_2=Z_2(A,\Phi,T_{\mathcal{C}}(x_*),u)$. We apply \Equation{delTermBd}, \Equation{gamma2pFirst} and \Equation{gamma2pLast} to find for any $p\geq \log m$, \begin{align} \EquationName{del2pl2l1} & (\mathbb{E} d_{\delta}^{2p}(T))^{1/p} \nonumber\\ & \qquad \leq \frac{1}{s}\Big[\sup_{x\in T_{\mathcal{C}}(x_*) \ : \ \|Ax\|_{2,1}=1}\|x\|_{2,1}^2\Big] \Big(\mathbb{E}_{\delta}\Big(\max_{1\leq i\leq m} \mathbb{E}_g\Big\|\sum_{j=1}^n g_j\delta_{ij} A^*e_j\Big\|_{2,\infty}\Big)^{2p}\Big)^{1/p} \nonumber \\ & \qquad \leq \frac{1}{s} (\log b)\Big[\sup_{x\in T_{\mathcal{C}}(x_*) \ : \ \|Ax\|_{2,1}=1}\|x\|_{2,1}^2\Big]\Big(\mathbb{E}_{\delta}\max_{1\leq \ell\leq b}\Big(\sum_{j=1}^n \delta_{ij}\|(A_{jk})_{k\in B_{\ell}}\|_2^2\Big)^p\Big)^{1/p} \nonumber \\ & \qquad \leq (\log b)\Big[\sup_{x\in T_{\mathcal{C}}(x_*) \ : \ \|Ax\|_{2,1}=1}\|x\|_{2,1}^2\Big] \Big(\frac{1}{m} \tnorm{A}^2 + \frac{1}{s}(\log b + p) \|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}}^2\Big). \end{align} By \Equation{RIPLpEst}, Lemma~\ref{lem:gamma2pEst} and \Equation{del2pl2l1} \begin{align*} & \Big(\mathbb{E}_{\delta,\sigma}\sup_{x \in T} \Big|\|\Phi x\|_2^2 - \|x\|_2^2\Big|^p\Big)^{1/p}\nonumber\\ & \ \ \ \lesssim (\mathbb{E}_{\delta}\gamma_2^{2p}(T,\|\cdot\|_{\delta}))^{1/p} + (\mathbb{E}_{\delta}\gamma_2^{p}(T,\|\cdot\|_{\delta}))^{1/p} + \sqrt{p}(\mathbb{E} d_{\delta}^p(T))^{1/p} + p(\mathbb{E} d_{\delta}^{2p}(T))^{1/p} \\ & \ \ \ \lesssim \frac{\alpha E}{s}(\log b + p) + \frac{\alpha F}{m} + \Big(\frac{\alpha E}{s}(\log b + p) + \frac{\alpha F}{m}\Big)^{1/2} \\ & \ \ \ \qquad + \frac{E p\log b}{s}(\log b + p) + \frac{F p\log b}{m} + \Big(\frac{E p\log b}{s}(\log b + p) + \frac{F p\log b}{m}\Big)^{1/2}, \end{align*} where we write \begin{equation*} \label{eqn:CDDef} E = \|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}}^2 d_{2,1}^2, \ F = \tnorm{A}^2 d_{2,1}^2. \end{equation*} Similarly, using Lemma~\ref{lem:Z2boundGen} (with $Z(u):=Z_2(A,\Phi,T_{\mathcal{C}}(x_*),u)$) we find \begin{align*} (\mathbb{E}_{\delta,\sigma}Z_2^p)^{1/p} & \lesssim \Big(\sqrt{\frac{p}{s}} + 1\Big)\Big((\mathbb{E}_{\delta}\gamma_2^p(T,\|\cdot\|_{\delta}))^{1/p} + \sqrt{p}(\mathbb{E}_{\delta}d_{\delta}^p(T))^{1/p}\Big) \\ & \lesssim \Big(\sqrt{\frac{p}{s}} + 1\Big)\Big[\Big(\frac{\alpha E}{s}(\log b + p) + \frac{\alpha F}{m}\Big)^{1/2} \\ & \qquad \qquad \qquad \qquad \qquad \qquad + \Big(\frac{E p\log b}{s}(\log b + p) + \frac{F p\log b}{m}\Big)^{1/2}\Big]. \end{align*} Now apply Lemma~\ref{lem:MomentsToTails} (with $w=\log(\eta^{-1})$) to conclude that with probability at least $1-\eta$ we have $Z_2\leq \varepsilon$ and $Z_1\geq 1-\varepsilon$ under the assumptions on $m$ and $s$. Lemma~\ref{lem:compare} now implies that $$f(\hat{x})\leq \Big(1+\frac{\varepsilon}{1-\varepsilon}\Big)^2f(x_*) = (1-\varepsilon)^{-2}f(x_*).$$ \end{proof} \begin{corollary} \label{cor:SJLl21} Set $$\mathcal{C}=\{x\in \mathbb{R}^d \ : \ \|x\|_{2,1}\leq R\}.$$ Suppose that $x_*$ is $k$-block sparse and $\|x_*\|_{2,1}=R$. Define $$\sigma_{\min,k} = \inf_{\|y\|_2=1, \ \|y\|_{2,1}\leq 2\sqrt{k}}\|Ay\|_2.$$ Set $\alpha = (\log n)^6(\log m)^2(\log b)^2$. Assume that \begin{align*} m & \gtrsim \varepsilon^{-2}\log(\eta^{-1})(\alpha + \log(\eta^{-1}) \log b) \tnorm{A}^2 k\sigma_{\min,k}^{-2},\\ s & \gtrsim \varepsilon^{-2}\log(\eta^{-1})(\log(\eta^{-1}) + \log b)(\alpha + \log(\eta^{-1}) \log b) \|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}}^2 k\sigma_{\min,k}^{-2}, \end{align*} and $\eta\leq \frac{1}{m}$. Then with probability at least $1-\eta$ $$f(\hat{x})\leq (1-\varepsilon)^{-2}f(x^*).$$ \end{corollary} \begin{proof} As calculated in Example~\ref{exa:TCl21} of Appendix~\ref{sec:convex-tools}, every $x\in T_{\mathcal{C}}(x_*)$ satisfies $\|x\|_{2,1}^2\leq 4k\|x\|_2^2$. If moreover $\|Ax\|_2=1$, then $$\sigma_{\min,k}^2\leq \|x\|_2^{-2}\leq 4k\|x\|_{2,1}^{-2}.$$ In other words, $$\sup_{x\in T_{\mathcal{C}}(x_*) \ : \ \|Ax\|_2=1}\|x\|_{2,1}^2 \leq 4k\sigma_{\min,k}^{-2}.$$ \end{proof} For a dense random sign matrix $\Phi$, it follows from \cite[Theorem 1]{PiW14} that the result in Corollary~\ref{cor:SJLl21} holds under the condition $$m\gtrsim \varepsilon^{-2}(\tnorm{A}^2 k\sigma_{\min,k}^{-2}+\log(\eta^{-1})).$$ Thus, our result shows that one can substantially increase the sparsity in the matrix $\Phi$, while maintaining the embedding dimension $m$ (up to logarithmic factors).\par Let us now specialize our results to the case $D=1$, which corresponds to the Lasso. In this case, if we let $A_k$ denote the columns of $A$, $$\tnorm{A} = \max_{1\leq k\leq d}\|A_k\|_2,$$ the maximum column of $A$. We can alternatively interpret this as the norm $\|A\|_{\ell_1\rightarrow\ell_2}$. Moreover, the norm $\|A\|_{\ell_{2,1}\rightarrow \ell_{\infty}}$ reduces to $$\|A\|_{\ell_{1}\rightarrow \ell_{\infty}} = \max_{1\leq j\leq n}\max_{1\leq k\leq d} |A_{jk}|,$$ the largest entry of the matrix. The first norm can be significantly larger than the second. For example, if $A$ is filled with $\pm 1$ entries, then $$\|A\|_{\ell_1\rightarrow\ell_2} = \sqrt{n}, \qquad \|A\|_{\ell_{1}\rightarrow \ell_{\infty}} = 1.$$ A similar result for the fast J-L transform, but with a worse dependence of $m$ on the matrix $A$, was obtained in \cite{PiW14}. For completeness, we will show in Appendix~\ref{sec:fjlt-cls} that the fast J-L transform in fact satisfies similar results as Theorem~\ref{thm:SJLmain} and Corollary~\ref{cor:SJLl21}, with the essentially the same dependence of $m$ on $A$. \section{Proof of the main theorem}\SectionName{general} After having seen instantiations of various subsets of our ideas for specific applications (linear subspaces, $\tnorm{\cdot}_T$ of small type-2 constant, and closed convex cones), we now prove the main theorem of this work, \Theorem{main}. Our starting point is the observation that inequality \Equation{3.3}, i.e., \begin{align} & \log\mathcal{N}(\tilde{T}, \vertiii{\cdot}, t) \nonumber\\ & \qquad \lesssim \Big(\log\frac {d_{\delta}(T)}{t}\Big)(\log m)\log\mathcal{N}\Big(\mathrm{conv}\Big(\bigcup_{i=1}^m B_{J_i}\Big), \tnorm{\cdot}_T, \frac 18\sqrt{s} t\Big) \EquationName{3.3Rep} \end{align} holds in full generality. We will need the following replacement of \Claim{claim}. \begin{lemma}\LemmaName{maurey} Let $\epsilon>0$. Then \begin{align} \nonumber & \log\mathcal{N}\Big(\mathrm{conv}\Big(\bigcup_{i=1}^m B_{J_i}\Big), \tnorm{\cdot}_T, \epsilon\Big) \\ & \qquad \lesssim \frac 1{\epsilon^2}\log m + (\log 1/\epsilon) \max_{k\lesssim \frac 1{\epsilon^2}} \max_{|A| = k} \log\mathcal{N}\Big(\frac 1k\sum_{i\in A}B_{J_i}, \tnorm{\cdot}_T, \epsilon\Big) \EquationName{5.2} \end{align} \end{lemma} \begin{proof} Set $\rho_T(x,y)=\tnorm{x-y}_T$ and let $\rho_{\ell_2}(x,y)=\|x-y\|_2$. By Maurey's lemma (Lemma~\ref{lem:Maurey} for $\|\cdot\|_2$), given $x\in\mathrm{conv}(B_{J_i} : 1\le i\le m)$, there is a $k\lesssim 1/\epsilon^2$ and a set $A\subset\{1,\ldots,m\}$, $|A| = k$, such that \begin{equation} \rho_T(x, \mathrm{conv}(B_{J_i}; i\in A)) \leq \rho_{\ell_2}(x, \mathrm{conv}(B_{J_i}; i\in A)) \le \epsilon \EquationName{5.1} \end{equation} Now, let us take an element $y\in\mathrm{conv}(B_{J_i} ; i\in A)$, $|A| \lesssim 1/\epsilon^2$. Thus $$ y = \sum_{i=1}^k \lambda_i y_i $$ with $y_i\in B_{J_i}$, $\lambda_i\ge 0$, $\sum_i \lambda_i = 1$, $k \lesssim 1/\epsilon^2$. Firstly, we may dismiss the coefficients $\lambda_i < \epsilon^3$. Indeed, let $S=\{i \ : \ \lambda_i<\epsilon^3\}$ and set $\hat{y}=\sum_{i\in S}\lambda_i y_i$. Then, $$\|\hat{y}\|_2 \leq \sum_{i\in S}\lambda_i\leq k\epsilon^3\lesssim \epsilon.$$ Consider now the $\lambda_i$ with $\epsilon^3 \le \lambda_i \le 1$. For $\ell=0,1,\ldots,\ell_*$, $\ell_* \simeq \log(1/\epsilon)$, denote $$ A_\ell = \Big\{1\leq i\leq k \ : \ \frac 1{2^\ell} \ge \lambda_i > \frac 1{2^{\ell+1}} \Big\} $$ and write \begin{equation} y = \sum_{\ell=0}^{\ell_*} 2^{-\ell} |A_\ell| \cdot\Big(\frac 1{|A_\ell|} \sum_{i\in A_{\ell}} y_i'\Big) + \hat{y} \EquationName{one} \end{equation} where $y_i' = \lambda_i 2^\ell y_i \in B_{J_i}$. Note that \begin{equation} \sum_{\ell=0}^{\ell_*} 2^{-\ell} |A_\ell| < 2\sum_i \lambda_i \le 2 \EquationName{two} \end{equation} and $$ \frac 1{|A_\ell|} \sum_{i\in A_{\ell}} y_i' \in \frac 1{k_\ell} \sum_{i\in A_{\ell}} B_{J_i} $$ where $k_\ell = |A_\ell| \lesssim 1/\epsilon^2$. Take finite sets $\xi_\ell \subset \frac 1{k_\ell} \sum_{i\in A_\ell} B_{J_i}$ such that \begin{align} &\rho_T(z, \xi_\ell) < \epsilon \text{ for all } z\in \frac 1{k_\ell}\sum_{i\in A_\ell} B_{J_i}\\ &|\xi_\ell| = \mathcal{N}\Big(\frac 1{k_\ell}\sum_{i\in A_\ell} B_{J_i},\tnorm{\cdot}_T,\epsilon\Big) \end{align} Let $z_\ell\in\xi_\ell$ satisfy $$ \btnorm{\frac 1{|A_\ell|} \sum_{i\in A_\ell} y_i' - z_\ell}_T < \epsilon $$ and set \begin{equation} z = \sum_{\ell=0}^{\ell_*} 2^{-\ell} |A_\ell| z_\ell. \EquationName{five} \end{equation} By \Eqsub{one}, \Eqsub{two} $$ \tnorm{y - z}_T \lesssim \sum_{\ell=0}^{\ell_*} 2^{-\ell} |A_\ell| \epsilon + \epsilon \lesssim \epsilon $$ To summarize, we can find an $\epsilon$-net for $\mathrm{conv}(\cup_{1\leq i\leq m}B_{J_i})$ with respect to $\tnorm{\cdot}_T$ as follows. For every $A\subset [m]$ with $|A|=k$, $k\lesssim 1/\epsilon^2$, we select $\xi_1^A,\ldots,\xi_{\ell_*}^A$ as above. Then, $$\bigcup_{k\in [m] \ : \ k\lesssim 1/\epsilon^2}\bigcup_{A\subset[m], |A|=k} \sum_{\ell=0}^{\ell_*}2^{-\ell}|A_{\ell}|\xi_{\ell}^A$$ is an $\epsilon$-net of cardinality at most \begin{align*} & \max_{k\lesssim 1/\epsilon^2} \frac 1{\epsilon^2} m {m\choose k} \prod_{\ell=0}^{\ell_*} \mathcal{N}\Big(\frac{1}{k_{\ell}}\sum_{i\in A_{\ell}} B_{J_i},\tnorm{\cdot}_T,\epsilon\Big) \\ & \qquad \lesssim \max_{k\lesssim 1/\epsilon^2}\frac 1{\epsilon^2} m \Big(\frac{em}{k}\Big)^k \Big[\max_{k\leq 1/\epsilon^2}\max_{|A|=k} \mathcal{N}\Big(\frac{1}{k}\sum_{i\in A} B_{J_i},\tnorm{\cdot}_T,\epsilon\Big)\Big]^{\log(1/\epsilon)}. \end{align*} This yields the result. \end{proof} Next, we analyze further the set $(1/k)\sum_{i\in A} B_{J_i}$ for some $k\lesssim 1/\epsilon^2$ ($\epsilon>0$ will be fixed later). The elements of $(1/k) \sum_{i\in A}B_{J_i}$ are of the form $$ y = \frac 1k\sum_{j=1}^n \Big(\sum_{i\in A} \delta_{ij}x_j^{(i)}\Big) e_j $$ with $$ \sum_{j\in J_i} |x_j^{(i)}|^2 \le 1, \qquad \mathrm{for \ all} \ i. $$ Therefore, by Cauchy-Schwarz, \begin{align*} \|y\|_2 &= \frac 1k \Big(\sum_{j=1}^n \Big|\sum_{i\in A} \delta_{ij} x_j^{(i)}\Big|^2\Big)^{1/2}\\ {}&\le \frac 1k \Big[\sum_{j=1}^n \Big(\sum_{i\in A}\delta_{ij}\Big)\sum_{i\in A} |x_j^{(i)}|^2 \Big]^{1/2} \end{align*} Define for $\alpha = 1,\ldots,\log(\min\{k, s\})$ the set \begin{equation} U_{\alpha} = U_{\alpha}(\delta) = \Big\{1\leq j\leq n : 2^{\alpha}\leq \sum_{i\in A} \delta_{ij}<2^{\alpha+1}\Big\} \EquationName{5.3} \end{equation} and define $$U_0 = U_0(\delta) = \Big\{1\leq j\leq n : \sum_{i\in A} \delta_{ij}<2\Big\}.$$ Estimate for fixed $j$ \begin{equation} \EquationName{IntensityEst} \tau_{k,\alpha} \mathbin{\stackrel{\rm def}{=}} \Pr_\delta\Big(2^\alpha\leq \sum_{i\in A} \delta_{ij} < 2^{\alpha+1}\Big) \le \begin{cases} 1,\ \hbox{if } 2^\alpha \le \frac {2esk} m\\ \min\Big\{2^{-\alpha} \frac {sk} m, 2^{-2^\alpha}\Big\},\ \hbox{if } 2^\alpha > \frac {2esk} m, \end{cases} \end{equation} where the first term in the $\min$ is a consequence of Markov's inequality, and the second term follows by using that $$\tau_{k,\alpha} \leq \mathbb{P}\Big(\sum_{j=1}^k\zeta_j\geq 2^{\alpha}\Big),$$ whenever $\zeta_1,\ldots,\zeta_k$ are i.i.d.\ Bernoulli random variables with mean $s/m$ (see \Eqsub{binom-calc} for bounding Bernoulli moments). Hence, $U_\alpha$ is a random set of intensity $\tau_{k,\alpha}$, i.e., $\Pr_\delta (j \in U_\alpha(\delta)) = \tau_{k,\alpha}$. Write according to the preceding $$ y = \sum_\alpha y_\alpha, \text{ with } y_\alpha = \sum_{j\in U_\alpha} y_j e_j $$ and $$ \|y_\alpha\|_2 \lesssim \frac 1{\sqrt{k}} 2^{\alpha/2} $$ Hence, denoting $B_{U_{\alpha}}:=\{\sum_{j\in U_{\alpha}}x_je_j \ : \ \sum_{j\in U_{\alpha}}|x_j|^2\leq 1\}$, $$ \frac 1k \sum_{i\in A} B_{J_i} \subset \sum_\alpha \frac 1{\sqrt{k}} 2^{\alpha/2} B_{U_\alpha}. $$ Therefore, we deduce that \begin{align} \nonumber \log\mathcal{N}\Big(\frac 1k\sum_{i\in A} B_{J_i}, \tnorm{\cdot}_T, \epsilon\Big) &\lesssim \sum_\alpha \log\mathcal{N}\Big(\frac {1}{\sqrt{k}} 2^{\alpha/2}B_{U_\alpha}, \tnorm{\cdot}_T, \frac{\epsilon}{\log m}\Big)\\ {}&= \sum_\alpha \log\mathcal{N}\Big(B_{U_\alpha}, \tnorm{\cdot}_T, \sqrt{\frac k{2^\alpha}} \frac{\epsilon}{\log m}\Big)\EquationName{5.4} \end{align} By the dual Sudakov inequality (Lemma~\ref{lem:dualSudakov} of Appendix~\ref{sec:probTools}), $$ t\Big[\log\mathcal{N}(B_{U_\alpha}, \tnorm{\cdot}_T, t)\Big]^{1/2} \lesssim \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T $$ It follows that $$ \Big[\log\mathcal{N}\Big(\frac 1k\sum_{i\in A} B_{J_i}, \tnorm{\cdot}_T, \epsilon\Big)\Big]^{1/2} \lesssim \sum_\alpha \Big(\frac{2^\alpha}k\Big)^{1/2} \Big(\frac{\log m}{\epsilon}\Big) \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T $$ and, by \Lemma{maurey} \begin{align} \nonumber \Big[\log& \mathcal{N}\Big(\mathrm{conv}\Big(\bigcup_{i=1}^m B_{J_i}\Big), \tnorm{\cdot}_T, \epsilon\Big)\Big]^{1/2}\\ {}&\le \frac 1{\epsilon}(\log m)^{1/2} + \frac{\log m}{\epsilon}\Big(\log\frac 1{\epsilon}\Big)^{1/2} \max_{\substack{k\lesssim \frac 1{\epsilon^2}\\|A| = k}}\Big[\sum_\alpha \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T\Big] \EquationName{5.5} \end{align} Applying \Equation{5.5} to \Equation{3.3Rep}, taking $\epsilon \simeq \sqrt{s} t$, gives \begin{align*} t\log^{1/2} \mathcal{N}(\tilde{T}, \vertiii{\cdot}, t) \lesssim& \frac 1{\sqrt{s}} \Big(\log \frac{1}{\sqrt{s}t}\Big)^{1/2}\log m\\ \nonumber {}& + \frac 1{\sqrt{s}}(\log m)^{3/2}\Big(\log \frac{1}{\sqrt{s}t}\Big)\cdot\\ {}& \hspace{.2in}\max_{\substack{|A| = k\\k\lesssim \frac 1{\epsilon^2}}}\Big\{\sum_\alpha \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T \Big\}. \end{align*} Using this bound for large $t$ and the elementary bound in \Equation{volComp} for small $t$ in \Equation{Dudley}, we obtain \begin{align} \nonumber \gamma_2(T,\|\cdot\|_{\delta}) \lesssim &\frac 1{\sqrt{s}}(\log n)^{3/2}\log m \\ \nonumber &{} + \frac 1{\sqrt{s}}(\log m)^{3/2}(\log n)^2\cdot\\ {}& \hspace{.2in}\max_{k\le m}\max_{|A| = k} \Big\{\sum_{\alpha,2^\alpha\le k} \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T \Big\}\EquationName{5.6} \end{align} We split the sum over $\alpha$ into three different parts. Firstly, \begin{align} \EquationName{MainBd1} \max_{k\le m}\max_{|A| = k} \sum_{\alpha,2^\alpha\leq 2esk/m} \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T & \lesssim \sqrt{\frac{s}{m}} \mathbb{E}_g\btnorm{\sum_{j=1}^n g_j e_j}_T \nonumber \\ & \lesssim \sqrt{\frac{s}{m}}\gamma_2(T,\|\cdot\|_2). \end{align} Next, by setting $$U_A=\Big\{j\in [n] \ : \ \frac{2esk}{m}\leq \sum_{i\in A} \delta_{ij}\Big\},$$ we obtain \begin{align*} & \max_{k\le m}\max_{|A| = k} \sum_{\alpha,2esk/m<2^{\alpha}\leq 10\log n} \sqrt{\frac{2^\alpha}{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T \\ & \qquad \lesssim (\log n)^{1/2} \max_{k\le m}\max_{|A| = k} \frac{1}{\sqrt{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T \\ & \qquad \leq (\log n)^{1/2} \Big[\max_{m/s\leq k\le m}\max_{|A| = k} \frac{1}{\sqrt{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T \\ & \qquad \qquad \qquad \qquad \qquad \qquad + \max_{k\leq m/s}\max_{|A| = k} \frac{1}{\sqrt{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T\Big]. \end{align*} Note that the first term on the right hand side is bounded by \begin{equation} \EquationName{MainBd2a} \max_{m/s\leq k\le m}\max_{|A| = k} \frac{1}{\sqrt{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T \lesssim \sqrt{\frac{s}{m}}\gamma_2(T,\|\cdot\|_2). \end{equation} To bound the second term, we take expectations with respect to $\delta$ and find for $p_k=k\log\Big(\frac{m}{sk}\Big)$ \begin{align} \EquationName{UAint} & \mathbb{E}_{\delta}\Big[\max_{k\leq m/s}\max_{|A| = k} \frac{1}{\sqrt{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T\Big] \nonumber\\ & \qquad \lesssim \max_{k\leq m/s}\max_{|A| = k} \frac{1}{\sqrt{k}} \Big(\mathbb{E}_{\delta}\Big(\mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T\Big)^{p_k}\Big)^{1/p_k}. \end{align} By \Equation{IntensityEst}, for any $A\subset [m]$ with $|A|=k$, $U_A$ is a random set of intensity at most $esk/m=sq/(m\log s)$, where we set $q=ek(\log s)$. If we now let $\eta_1,\ldots,\eta_n$ be i.i.d.\ $\{0,1\}$-valued random variables with expectation $qs/(m\log s)$, then the random set $U$ defined by $\mathbb{P}(j\in U)=\mathbb{P}(\eta_j=1)$ has a higher intensity than $U_A$ and therefore, \begin{align} \EquationName{MainBd2b} & \mathbb{E}_{\delta}\Big[\max_{k\leq m/s}\max_{|A| = k} \frac{1}{\sqrt{k}} \mathbb{E}_g\btnorm{\sum_{j\in U_A} g_j e_j}_T\Big] \nonumber\\ & \qquad \lesssim (\log s)^{1/2}\max_{q\leq \frac{m}{s}\log s} \frac{1}{\sqrt{q}} \Big(\mathbb{E}_{\eta}\Big(\mathbb{E}_g\btnorm{\sum_{j=1}^n \eta_j g_j e_j}_T\Big)^q\Big)^{1/q}. \end{align} Finally, consider the $\alpha$ with $2^{\alpha}>10\log n$. Since $\tnorm{\cdot}_T\leq \|\cdot\|_2$, \begin{align} \EquationName{appLambda} & \max_{k\le m}\max_{|A| = k} \sum_{\alpha,10\log n<2^{\alpha}\leq k} \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T \nonumber\\ & \qquad \leq \max_{k\le m}\max_{|A| = k} \sum_{\alpha,10\log n<2^{\alpha}\leq k} \sqrt{\frac{2^\alpha}k} |U_{\alpha}|^{1/2}. \end{align} By \Equation{IntensityEst}, the intensity of $U_{\alpha}$ is bounded by $2^{-2^{\alpha}}$ and therefore, \begin{align*} & \mathbb{E}_{\delta}\Big[\max_{k\le m}\max_{|A| = k} \sum_{\alpha,10\log n<2^{\alpha}\leq k} \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T \Big] \\ & \qquad \leq \max_{k\le m}\max_{|A| = k} \sum_{\alpha,10\log n<2^{\alpha}\leq k} \sqrt{\frac{2^\alpha}k} \Big(\mathbb{E}\Big|\sum_{j=1}^n \zeta_j\Big|^{q_k}\Big)^{1/(2q_k)}, \end{align*} where $q_k\simeq k\log(m/k)$ and the $\zeta_j$ are i.i.d.\ $\{0,1\}$-valued with mean $2^{-2^{\alpha}}$. Since $\sum_j \zeta_j$ is binomially distributed, we find using that $2^{\alpha}>10\log n$, \begin{equation} \Big\|\sum_j \zeta_j\Big\|_{L_{\zeta}^{q_k}} \leq \Big(\sum_{t=1}^n t^{q_k} e^{-\frac{9}{10}2^{\alpha}t}\Big)^{1/q_k} \lesssim n^{1/q_k} 2^{-\alpha} q_k \lesssim 2^{-\alpha} k(\log m)\EquationName{binom-calc} \end{equation} and therefore \begin{equation} \EquationName{MainBd3} \mathbb{E}_{\delta}\Big[\max_{k\le m}\max_{|A| = k} \sum_{\alpha,10\log n<2^{\alpha}\leq k} \sqrt{\frac{2^\alpha}k} \mathbb{E}_g\btnorm{\sum_{j\in U_\alpha} g_j e_j}_T\Big] \lesssim (\log m)^{3/2}. \end{equation} By combining \Equation{5.6}, \Equation{MainBd1}, \Equation{MainBd2a}, \Equation{MainBd2b}, and \Equation{MainBd3} we obtain \begin{align} \nonumber & \mathbb{E} \gamma_2(T,\|\cdot\|_{\delta}) \\ \nonumber & \qquad \lesssim \frac{(\log m)^{3/2}(\log n)^{5/2}\gamma_2(T,\|\cdot\|_2)}{\sqrt{m}} +\frac 1{\sqrt{s}}\Bigg\{(\log m)^3(\log n)^2 +\\ {}&\qquad \hspace{.6in} (\log m)^2(\log n)^{5/2}\max_{q\le \frac ms \log s} \frac 1{\sqrt{q}}\Big(\mathbb{E}_{\eta}\Big(\mathbb{E}_g\btnorm{\sum_{j=1}^n \eta_jg_j e_j}_T\Big)^q\Big)^{1/q}\Bigg\}, \EquationName{5.13} \end{align} with $\eta_j$ i.i.d.\ Bernoulli with mean $qs/(m\log s)$. From our proof it is clear that $\mathbb{E} \gamma_2^2(T,\|\cdot\|_{\delta})$ can be bounded by the square of the right hand side of \Equation{5.13}. Using \Equation{RIPLpEst} (for $p=1$), we conclude that $$\mathbb{E}\sup_{x\in T}\Big|\|\Phi x\|_2^2 - 1\Big|<\varepsilon$$ provided that $m,s$ satisfy \begin{align} m &\gtrsim (\log m)^3(\log n)^5 \frac{\gamma_2^2(T,\|\cdot\|_2)}{\varepsilon^2} \EquationName{5.14}\\ s &\gtrsim (\log m)^6 (\log n)^4 \frac 1{\varepsilon^2} \EquationName{5.15} \end{align} and \begin{equation} \Eqsub{5.16} = (\log m)^2(\log n)^{5/2}\max_{q\le \frac ms \log s} \Big\{\frac 1{\sqrt{qs}} \Big(\mathbb{E}_{\eta}\Big(\mathbb{E}_g\btnorm{\sum_{j=1}^n \eta_j g_j e_j}_T\Big)^q\Big)^{1/q}\Big\} < \varepsilon. \EquationName{5.16} \end{equation} This completes the proof. \section{Example applications of main theorem}\SectionName{applications} In this section we use our main theorem to give explicit conditions under which $$\mathbb{E}\sup_{x\in T}\Big|\|\Phi x\|_2^2 - 1\Big|<\varepsilon$$ for several interesting sets $T\subset S^{n-1}$. This amounts to computing an upper bound for the parameters $\gamma_2(T,\|\cdot\|_2)$ and \begin{equation} \EquationName{kappaTDefEx} \kappa(T) = \max_{q\le \frac ms \log s} \Big\{\frac 1{\sqrt{qs}} \Big(\mathbb{E}_{\eta}\Big(\mathbb{E}_g\btnorm{\sum_{j=1}^n \eta_j g_j e_j}_T\Big)^q\Big)^{1/q}\Big\}. \end{equation} We focus on the latter two and refer to \cite{Dirksen14} for details on how to estimate $\gamma_2(T,\|\cdot\|_2)$. Note, however, that $\gamma_2(T,\|\cdot\|_2)\lesssim (m \log s)^{1/2}\kappa(T)$. Indeed, take $q=\frac{m}{s}\log s$ in \Equation{kappaTDefEx} and note that the $\eta_j$ are identically equal to $1$ in this case. This gives $$\kappa(T)\geq (m \log s)^{-1/2} g(T) \simeq (m \log s)^{-1/2} \gamma_2(T,\|\cdot\|_2).$$ Thus, if we ignore logarithmic factors, it suffices to bound $\kappa(T)$. \subsection{Linear subspace} In the application from \Section{linear} with $T$ the unit ball of a $d$-dimensional linear subspace $E$ of $\ell_2^n$, we have $\gamma_2(T,\|\cdot\|_2) \simeq \sqrt{d}$ and \begin{equation} \Eqsub{5.16} \lesssim (\log m)^2(\log n)^{5/2}\frac 1{\sqrt{s}} \max_{q\le \frac ms \log s} \frac 1{\sqrt{q}} \Big\|\sum_j \|P_E e_j\|_2^2 \eta_j\Big\|_{L_\eta^q}^{1/2} \EquationName{5.17} \end{equation} with $\eta_j \in \{0,1\}$ i.i.d.\ of mean $qs/(m\log s)$. Using Khintchine's inequality, \begin{equation} \Big\|\sum_j \eta_j \|P_E e_j\|_2^2\Big\|_{L_\eta^q} < \frac{dqs}{m\log s} + q\max_j \|P_E e_j\|_2^2 \EquationName{5.18} \end{equation} and therefore \begin{equation} \Eqsub{5.16} \lesssim (\log m)^2(\log n)^{5/2}\Big(\sqrt{\frac dm} + \frac 1{\sqrt{s}}\max_j \|P_E e_j\|_2\Big) \EquationName{5.19} \end{equation} \Equation{5.14} and \Equation{5.15} then give conditions \begin{align} m &\gtrsim d(\log n)^5(\log m)^4 /\varepsilon^2 \EquationName{5.20}\\ s &\gtrsim (\log n)^4(\log m)^6/\varepsilon^2 + (\log n)^5(\log m)^4\max_j \|P_E e_j\|_2^2/\varepsilon^2 \EquationName{5.21} \end{align} \subsection{$k$-sparse vectors}\SectionName{ksparse2} Consider next the application from \Section{ksparse}, replacing $T$ by $K$ given by \Equation{4.1}. Thus $\gamma_2(K) \lesssim \sqrt{k\log n}$ and \Equation{5.16} is bounded by \begin{equation} \sqrt{k}(\log m)^2(\log n)^3 \frac 1{\sqrt{s}} \max_{q\le \frac ms \log s}\Big\{\frac 1{\sqrt{q}} \Big\|\max_i \Big(\sum_j \eta_j |A_{ij}|^2\Big) \Big\|_{L^q_{\eta}}^{1/2}\Big\} \EquationName{5.22} \end{equation} and \begin{equation} \Big\|\max_i (\sum_j \eta_j |A_{ij}|^2)\Big\|_{L^q_{\eta}} \lesssim \frac{qs}{m\log s} + q\max_{i,j}|A_{ij}|^2 \EquationName{5.23} \end{equation} Hence \begin{equation} \Eqsub{5.16} \lesssim (\log m)^2(\log n)^3\sqrt{\frac km} + (\log m)^2(\log n)^3\sqrt{\frac ks}\max_{i,j}|A_{ij}| \EquationName{5.24} \end{equation} We then arrive at the conditions \begin{align} m &\gtrsim k(\log n)^6(\log m)^4/\varepsilon^2 \EquationName{5.25}\\ s &\gtrsim (\log n)^4(\log m)^6/\varepsilon^2 \EquationName{5.26} \end{align} provided that \begin{equation} \max_{i,j} |A_{ij}| < k^{-1/2} (\log n)^{-1} \EquationName{5.27} \end{equation} (compare with \Equation{4.7}-\Eqsub{4.8}). \subsection{Flat vectors} Let $T\subseteq S^{n-1}$ be finite with $\|x\|_\infty\le \alpha$ for all $x\in T$. Then by a similar calculation as in \Eqsub{sqrt-trick}, \begin{align} \Big(\mathbb{E}_\eta \Big(\mathbb{E}_g \sup_{x\in T} \Big|\sum_{i=1}^n \eta_i g_i x_i\Big|\Big)^q\Big)^{1/q} &\lesssim \sqrt{\log|T|}\cdot \Big\|\sup_{x\in T}\sum_{i=1}^n \eta_i x_i^2\Big\|_{L^q_{\eta}}^{1/2}\\ {}&\lesssim \sqrt{\frac{qs\log|T|}{m}} + (\alpha\log |T|)\sqrt{q} \end{align} Since $\gamma_2(T,\|\cdot\|_2) \lesssim \sqrt{\log|T|}$ we find the conditions \begin{align} m&\gtrsim (\log |T|)(\log m)^4(\log n)^5/\varepsilon^2\\ s&\gtrsim (\alpha\log |T|)^2(\log m)^4(\log n)^5/\varepsilon^2 + (\log m)^6(\log n)^4/\varepsilon^2, \end{align} which is qualitatively similar to \cite[Theorem 4.1]{Matousek08}. \subsection{Finite collection of subspaces} Let $\Theta$ be a finite collection of $d$-dimensional subspaces $E\subset \mathbb{R}^n$ with $|\Theta| = N$. Define $$ T = \bigcup_{E\in \Theta} \{x\in E : \|x\|_2 = 1 \}. $$ In this case, $\gamma_2(T, \|\cdot\|_2) \lesssim \sqrt{d} + \sqrt{\log N}$. For the duration of the next two sections, we define $$\alpha= \sup_{E\in\Theta} \max_j \|P_E e_j\|_2.$$ Recall that $\max_j \|P_E e_j\|_2$ is referred to as the {\em incoherence} of $E$, and thus $\sqrt{d/n} \le \alpha\le 1$ (cf.\ Remark~\ref{rem:coherence}) is the maximum incoherence in $\Theta$. \subsubsection{Collection of incoherent subspaces} To estimate $\kappa(T)$, consider the collection $\mathcal{A}$ of operators $A = \sum_j (P_E e_j)\otimes e_j$, $E\in \Theta$. Fix $\eta$ and define $R_\eta x = \sum_j \eta_j x_j e_j$. Then applying \cite[Theorem 3.5]{KMR14} to $\mathcal{A} R_\eta$, \begin{align} \nonumber \Big\|\max_{E\in \Theta} \|\sum_j \eta_j g_j P_E e_j\| \Big\|_{L_g^1} &= \Big\|\max_{A\in\mathcal{A}} \| AR_\eta g\|\Big\|_{L_g^1}\\ {}&\lesssim d_F(\mathcal{A}R_\eta) + \gamma_2(\mathcal{A} R_\eta, \|\cdot\|) . \EquationName{twoc} \end{align} Clearly \begin{equation*} \|A\| \le 1,\ \|AR_\eta\|_F = \Big(\sum_j \eta_j \|P_E e_j\|_2^2\Big)^{1/2} . \end{equation*} By \cite{RV07} (see the formulation in \cite[Proposition 7]{Tropp08}), \begin{equation} \Big(\mathbb{E}_\eta \|A R_\eta\|^p\Big)^{1/p} \le 3\sqrt{p} \Big(\mathbb{E} \|AR_\eta\|_{1,2}^p\Big)^{1/p} + \sqrt{\varrho}\cdot \|A\| , \end{equation} where $\mathbb{E} \eta_j = \varrho$ and we assume $p \ge 2 \log n$. Here $\|\cdot\|_{1,2}$ is the $\ell_1^n \rightarrow \ell_n^2$ norm, hence $$\|AR_{\eta}\|_{1,2} = \max_{1\leq k\leq n}\|AR_{\eta}e_k\|_2 = \max_{1\leq k\leq n}\eta_k \|P_E e_k\|_2 \le \alpha.$$ Taking $\varrho = \frac{qs}{m\log s}$, it follows that \begin{equation} \Big(\mathbb{E}_\eta \|AR_\eta\|^p\Big)^{1/p} \lesssim \alpha\sqrt{p} + \Big(\frac{qs}{m\log s}\Big)^{1/2} \EquationName{fourc} \end{equation} To estimate $\kappa(T)$, we need to bound \begin{align} \|\Eqsub{twoc}\|_{L_\eta^q} &\le \underbrace{\Big\|\max_E \Big(\sum_j \eta_j \|P_E e_j\|_2^2\Big)\Big\|_{L_\eta^q}^{1/2}}_{\sqrt{\Eqsub{fivec}}} \EquationName{fivec}\\ {}&+ \underbrace{\|\gamma_2(\mathcal{A} R_\eta)\|_{L_{\eta}^q}}_{\Eqsub{sixc}} \EquationName{sixc} \end{align} First, by \Equation{5.18} and denoting $q_1 = q + \log N$, \begin{equation} \Eqsub{fivec} \lesssim \varrho d + q_1 \max_j \|P_E e_j\|_2^2 = \frac{dqs}{m\log s} + (q + \log N)\alpha^2 \EquationName{sevenc} \end{equation} Estimate trivially $$ \gamma_2(\mathcal{A}R_\eta) \le \sqrt{\log N} \cdot \max_{A\in\mathcal{A}} \|A R_\eta\| $$ and hence, applying \Equation{fourc} with $p = q + 2\log n + \log N$ \begin{align} \nonumber \Eqsub{sixc} &\lesssim \sqrt{\log N} \cdot \Big(\max_A \mathbb{E} \|AR_\eta\|^p\Big)^{1/p} \\ {}&\lesssim \sqrt{\log N} \cdot \Big\{(q + \log n + \log N)\alpha^2 + \frac{qs}{m\log s}\Big\}^{1/2} \EquationName{eightc} \end{align} Collecting our estimates we find \begin{align} \nonumber \kappa(T) & \lesssim \max_{q\leq\frac{m}{s}\log s}\frac{1}{\sqrt{qs}} \Bigg\{\Big(\frac{dqs}{m\log s}\Big)^{1/2} + \alpha\sqrt{q}\\ \nonumber&\hspace{.1in}{} \qquad + \sqrt{\log N}\cdot\Big[(\sqrt{q} + \sqrt{\log n} + \sqrt{\log N})\alpha + \Big(\frac{qs}{m\log s}\Big)^{1/2}\Big]\Bigg\}\\ {}&\lesssim \Big(\frac{d + \log N}{m\log s}\Big)^{1/2} + \frac{(\sqrt{\log n}{\sqrt{\log N}} + \log N)\alpha}{\sqrt{s}} \EquationName{ninec} \end{align} Hence in this application (assuming $\log N \ge \log n$), \begin{equation} \Eqsub{5.16} < (\log m)^2(\log n)^{5/2} \Big\{\Big(\frac{d + \log N}{m}\Big)^{1/2} + \frac{\alpha\log N}{\sqrt{s}}\Big\} \EquationName{10c} \end{equation} and the conditions on $m,s$ become \begin{align} m &\gtrsim (\log m)^4(\log n)^5(d + \log N)\varepsilon^{-2} \EquationName{11c}\\ s&\gtrsim ((\log m)^6(\log n)^4 + (\alpha \log N)^2(\log m)^4(\log n)^5)\varepsilon^{-2} \EquationName{12c} \end{align} Notice that $s$ depends \emph{only} on $|\Theta|$ and not on the dimension $d$. Thus this bound is of interest when $\log|\Theta|$ is small compared with $d$. \subsubsection{Collection of coherent subspaces} We can also obtain a bound on $s$ that does not improve for small $\alpha$, but has linear dependence on $\log N$. Here we will not rely on \cite{RV07}. As described in \Section{intro}, this setting has applications to model-based compressed sensing. For example, for approximate recovery of block-sparse signals using the notation of \Section{intro}, our bounds will show that a measurement matrix $\Phi$ may have $m\mathbin{\substack{<\\ *}}\xspace kb + k\log(n/k)$, $s \mathbin{\substack{<\\ *}}\xspace k\log(n/k)$ and allow for efficient recovery. This is non-trivial if the number of blocks satisfies $b \mathbin{\substack{>\\ *}}\xspace \log(n/k)$. One may indeed trivially bound $$ \gamma_2(\mathcal{A} R_\eta) \lesssim \sqrt{\log N} $$ since certainly $\|AR_\eta\| \le \|A\| \le 1$. Hence $$ \Eqsub{sixc} \lesssim \sqrt{\log N} $$ leading to the following bound on $\kappa(T)$ \begin{equation*} \max_{q\leq\frac{m}{s}\log s} \frac 1{\sqrt{qs}} \Big\{\sqrt{\frac{dqs}{m\log s}} + (\sqrt{q} + \sqrt{\log N})\alpha + \sqrt{\log N}\Big\} \lesssim \sqrt{\frac dm} + \sqrt{\frac {\log N}s} \end{equation*} Instead of \Eqsub{11c}, \Eqsub{12c}, one may impose the conditions \begin{align} m &\gtrsim (\log m)^4(\log n)^5 \frac{d}{\varepsilon^2} + (\log m)^3(\log n)^5\frac{\log N}{\varepsilon^2}\\ s &\gtrsim (\log m)^6(\log n)^4\frac 1{\varepsilon^2} + (\log m)^4(\log n)^5\frac{\log N}{\varepsilon^2} \end{align} We remark that previous work \cite{NN13b} which achieved $m \approx d/\varepsilon^2$ for small $s$ had worse dependence on $\log N$: in particular $s \gtrsim (\log N)^3, m \gtrsim (\log N)^6$. In fact, Conjecture 14 of \cite{NN13b} if true would imply that the correct dependence on $N$ in both $m$ and $s$ should be $\log N$ (which is optimal due to known lower bounds for the Johnson-Lindenstrauss lemma, i.e., the special case $d=1$). We thus have shown that this implication is indeed true. \subsection{Possibly infinite collection of subspaces}\SectionName{finsler} Assume next $\Theta$ is an arbitrary family of linear $d$-dimensional subspaces $E\subset \mathbb{R}^n$, equipped with the Finsler metric $$ \rho_{\mathrm{Fin}}(E, E') = \|P_E - P_E'\| . $$ Let $$ T = \bigcup_{E\in\Theta} \{ x\in E : \|x\|_2 = 1 \} $$ for which (cf.\ \cite{Dirksen14}) \begin{equation} \gamma_2(T, \|\cdot\|_2) \lesssim \sqrt{d} + \gamma_2(\Theta, \rho_{\mathrm{Fin}}) . \EquationName{13c} \end{equation} Fix some parameter $\varepsilon_0>0$ and let $\Theta_1\subset \Theta$ be a finite subset such that \begin{align} &|\Theta_1| \le \mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, \varepsilon_0) \EquationName{15c}\\ &\rho_{\mathrm{Fin}}(E, \Theta_1) \leq \varepsilon_0\text{ for any } E \in \Theta \EquationName{16c} \end{align} Let further $$ T_1 = \bigcup_{E'\in\Theta_1} \{x \in E' : \|x\|_2 = 1\} $$ Let $x\in T, x\in E, E\in\Theta$ and $\rho_{\mathrm{Fin}}(E,E') \leq \varepsilon_0$ for some $E' \in \Theta_1$. Hence \begin{equation} x = P_{E'} x + (P_Ex - P_{E'} x) = x_1 + x_2 \EquationName{17c} \end{equation} where $x_1\in T_1$ and $x_2\in B_E + B_{E'}$, $\|x_2\|_2 \leq \varepsilon_0$. Hence $x_2\in T_2$ with $$ T_2 = \bigcup_{E, F\in\Theta} \{x \in B_E + B_F : \|x\|_2 \leq \varepsilon_0 \} $$ For $t < \varepsilon_0$, we estimate $\mathcal{N}(T_2, \|\cdot\|_2, t)$. Let $\Theta_t \subset \Theta$ satisfy \begin{align} &|\Theta_t| \leq \mathcal{N}\Big(\Theta, \rho_{\mathrm{Fin}}, \frac t{4}\Big) \EquationName{18c}\\ &\rho_{\mathrm{Fin}}(E, \Theta_t) \leq \frac t{4} \text{ for all } E\in\Theta \EquationName{19c} \end{align} By \Equation{volComp}, for each $E'\in\Theta_t$ we can find $\xi_{E'} \subset B_{E'}$ such that \begin{align} &\log|\xi_{E'}| \lesssim d \log\frac 1t \EquationName{20c}\\ &\rho_{\ell_2}(x, \xi_{E'}) \leq t\text{ for all } x\in B_{E'} \EquationName{21c} \end{align} Denote $$ \xi_t = \bigcup_{E',F'\in\Theta_t} (\xi_{E'} - \xi_{F'}) $$ for which by construction $$ \log|\xi_t| \lesssim \log\mathcal{N}\Big(\Theta, \rho_{\mathrm{Fin}}, \frac t{4}\Big) + d\log \frac 1t $$ Also, for $x\in T_2$, $x = y + z \in B_E + B_F$ and $E',F'\in\Theta_t$ satisfying $$ \rho_{\mathrm{Fin}}(E, E') \leq \frac t{4}, \rho_{\mathrm{Fin}}(F, F') \leq \frac t{4} , $$ we have $$ \|x - (P_{E'} y + P_{F'} z)\|_2 \leq \frac{t}{2}, $$ while $$ \rho_{\ell_2}(P_{E'} y, \xi_{E'}) \leq \frac t{4}, \qquad \rho_{\ell_2}(P_{F'} z, \xi_{F'}) \leq \frac t{4} $$ Therefore $$ \rho_{\ell_2}(x, \xi_t) \leq t $$ and we get for $t < \varepsilon_0$ (otherwise $\Eqsub{22c} = 0$) \begin{equation} \log\mathcal{N}(T_2, \|\cdot\|_2, t) \lesssim \log\mathcal{N}\Big(\Theta, \rho_{\mathrm{Fin}}, \frac t{4}\Big) + d\log \frac 1t \EquationName{22c} \end{equation} Using the decomposition \Equation{17c} and the bound \Equation{10c} for the contribution of $T_1$, we find \begin{align} \kappa(T) & \lesssim \Big\{\Big(\frac{d + \log\mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, \varepsilon_0)}{m}\Big)^{1/2} + \frac{\alpha\log\mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, \varepsilon_0)}{\sqrt{s}}\Big\} \EquationName{23c}\\ & \qquad \ + \max_{q\le \frac ms\log s}\Big\{\frac 1{\sqrt{qs}} \Big(\mathbb{E}_{\eta}\Big(\mathbb{E}_g\sup_{x\in T_2} \Big|\sum_{j=1}^n \eta_j g_j x_j\Big|\Big)^q\Big)^{1/q}\Big\} \EquationName{24c} \end{align} with $(\eta_j)\in\{0,1\}$ i.i.d. of mean $\frac{qs}{m\log s}$. Estimate by the contraction principle \cite{Kahane68}, the Gaussian concentration for Lipschitz functions (\Equation{GaussLip}), and Dudley's inequality \Equation{Dudley}\cite{Dudley67} \begin{align} \nonumber \frac 1{\sqrt{q}} \Big\|\sup_{x\in T_2} \Big|\sum_{j=1}^n \eta_j g_j x_j\Big|\Big\|_{L_g^1,L_\eta^q} &\le \frac 1{\sqrt{q}} \Big\|\sup_{x\in T_2} \Big|\sum_{j=1}^n g_j x_j\Big|\Big\|_{L_g^q}\\ \nonumber {}&\lesssim \Big\|\sup_{x\in T_2} \Big|\sum_{j=1}^n g_j x_j\Big|\Big\|_{L_g^1}\\ \nonumber {}&\lesssim \int_0^{\varepsilon_0} [\log\mathcal{N}(T_2, \|\cdot\|_2, t)]^{1/2} dt \\ {}&\lesssim \int_0^{\varepsilon_0}[\log\mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, t)]^{1/2} dt + \sqrt{d} \varepsilon_0 \sqrt{\log\frac 1{\varepsilon_0}}, \EquationName{25c} \end{align} where in the final step we used \Equation{22c}. Summarizing, the conditions on $m$ and $s$ are as follows (for any $\varepsilon_0>0$) \begin{align} \nonumber m &\gtrsim \varepsilon^{-2}(\log m)^3(\log n)^5 \gamma_2^2(\Theta, \rho_{\mathrm{Fin}})\\ &\hspace{.2in}{}+\varepsilon^{-2}(\log m)^4(\log n)^5[d + \log\mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, \varepsilon_0)] \EquationName{26c}\\ \nonumber s&\gtrsim \varepsilon^{-2}(\log m)^6(\log n)^4 + \varepsilon^{-2}(\log m)^4(\log n)^5[\alpha \log \mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, \varepsilon_0)]^2\\ \nonumber &\hspace{.2in}{}+\varepsilon^{-2}(\log m)^4(\log n)^5\varepsilon_0^2\Big(\log\frac 1{\varepsilon_0}\Big)d\\ &\hspace{.2in}{} + \varepsilon^{-2}(\log m)^4(\log n)^5\Big[\int_0^{\varepsilon_0}[\log\mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, t)]^{1/2} dt\Big]^2 \EquationName{27c} \end{align} If $|\Theta| = N < \infty$, then $\log\mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, t)\leq\log N$ and \Eqsub{26c}, \Eqsub{27c} turn into \Eqsub{11c}, \Eqsub{12c} for $\varepsilon_0\rightarrow 0$. \subsection{Manifolds} Let $\mathcal{M}\subset\mathbb{R}^n$ be a $d$-dimensional manifold obtained as the image $\mathcal{M} = F(B_{\ell_2^d})$, for a smooth map $F:B_{\ell_2^d}\rightarrow \mathbb{R}^n$. More precisely, we assume that $\|F(x) - F(y)\|_2 \simeq \|x - y\|_2$ and the {\em Gauss map}, which sends $x\in\mathcal{M}$ to $E_x$, the tangent plane at $x$, is Lipschitz from the geodesic distance $\rho_{\mathcal{M}}$ to $\rho_{\mathrm{Fin}}$. Following \cite{Dirksen14}, we want to ensure that the sparse matrix $\Phi$ satisfies \begin{equation} (1-\varepsilon)|\gamma| \le |\Phi(\gamma)| \le (1+\varepsilon)|\gamma|\EquationName{28c} \end{equation} for any $C^1$-curve $\gamma\subset \mathcal{M}$, where $|\cdot|$ denotes curve length. Note that \Equation{28c} is equivalent to requiring \begin{equation} 1-\varepsilon \le \|\Phi(v)\|_2 \le 1+\varepsilon \EquationName{29c} \end{equation} for any tangent vector $v$ of $\mathcal{M}$ at a point $x\in \mathcal{M}$. Denote by $$ \Theta = \{E_x : x\in \mathcal{M}\} $$ the tangent bundle of $\mathcal{M}$, to which we apply the estimates on $m,s$ obtained above. By assumption, $$ \rho_{\mathrm{Fin}}(E_x, E_y) \lesssim \rho_{\mathcal{M}}(x,y) \simeq \|F^{-1}(x) - F^{-1}(y)\|_2, $$ so that for $0\le t \le 1/2$ by \Equation{volComp}, \begin{equation*} \log \mathcal{N}(\Theta, \rho_{\mathrm{Fin}}, t) \lesssim \log \mathcal{N}(\mathcal{M}, \|\cdot\|_2, ct) \simeq \log \mathcal{N}(B_{\ell_2^d}, \|\cdot\|_2, t)\lesssim d\log\frac 1t. \end{equation*} In this application \begin{equation} \alpha = \max_{x\in\mathcal{M}} \max_{\substack{v\in E_x\\\|v\|_2 = 1}} \max_{1\le j\le n}|\inprod{v, e_j}| \EquationName{30c} \end{equation} We assume \begin{equation} \alpha \ll \frac 1{\sqrt{d}} \EquationName{31c} \end{equation} to make the below of interest (otherwise apply the result of \cite{Dirksen14}). Taking $\varepsilon_0 = \alpha\sqrt{d}$ in \Eqsub{26c}, \Eqsub{27c}, it follows that \Equation{28c} may be ensured under parameter conditions \begin{align} m &\gtrsim \varepsilon^{-2}(\log m)^4(\log n)^5 d\cdot \log\Big(\frac{1}{\alpha\sqrt{d}}\Big) \EquationName{32c}\\ s &\gtrsim \varepsilon^{-2}(\log m)^6(\log n)^4 + \varepsilon^{-2}(\log m)^4 (\log n)^7(\alpha d)^2 \EquationName{33c} \end{align} Thus for $\alpha = o(1/\sqrt{d})$, the condition on $s$ becomes non-trivial. Recall from Remark~\ref{rem:coherence} that $\alpha \ge \sqrt{d/n}$ and therefore the $\log(1/(\alpha\sqrt{d}))$ term in \Equation{32c} is at most $\log n$. \begin{remark} \textup{ Returning to the assumptions on $F$, consider $DF : B_{\ell_2^d}\rightarrow \mathcal{L}(\mathbb{R}^d, \mathbb{R}^n)$, the space of linear operators from $\mathbb{R}^d$ to $\mathbb{R}^n$. The first statement means that uniformly for $x\in B_{\ell_2^d}$, \begin{equation} c^{-1} \|\xi\|_2 \le \|DF(x)\xi\|_2 \le c\|\xi\|_2\text{ for } \xi\in\mathbb{R}^d \EquationName{34c} \end{equation} The second statement follows from requiring \begin{equation} \|DF(x) - DF(y)\|_{2\rightarrow 2} \le c\|x - y\|_2 \text{ for } x,y\in B_{\ell_2^d} \EquationName{35c} \end{equation} } \textup{ Indeed, since $E_x = DF(x)(\mathbb{R}^d), E_y = DF(y)(\mathbb{R}^d)$, it follows from \Equation{35c} $$ \inf_{\substack{u\in E_x\\ \|u\|_2 = 1}} \rho_{\ell_2^n}(u, E_y) \lesssim \|x - y\|_2 $$ implying $$ \rho_{\mathrm{Fin}}(E_x, E_y) = \|P_{E_x} - P_{E_y} \|_{2\rightarrow 2} \lesssim \|x - y\|_2 . $$ } \end{remark} \begin{remark}\RemarkName{bad-manifold} \textup{ If $\alpha \ge 1/\sqrt{d}$, then necessarily $s\gg d$ in \Equation{33c} (up to polylogarithmic factors). Thus the manifold case is very different from the case of linear subspaces. We sketch a construction to demonstrate this. } \textup{ Let $n > d^{10}$. Denote by $0\le \varphi \le 1$ a smooth bump function on $\mathbb{R}$ such that \begin{equation} \varphi(t) = t \text{ for } \frac 14\le t \le \frac 12,\ \mathrm{supp}(\varphi) \subset [0,1] \EquationName{36c} \end{equation} } \textup{ By the lower bound in \Equation{volComp}, there is a collection $\{a_\beta\}_{1\le \beta\le 2^d}$ of $2^d$ points in $B_{\ell_2^d}(0, 1/2) \mathbin{\stackrel{\rm def}{=}} B_{1/2}\subset\mathbb{R}^d$ such that $$ \|a_{\beta'} - a_{\beta} \|_2 > \frac 1{10}\text{ for } \beta'\neq \beta $$ and let $(\eta_\beta)_{1\le \beta\le 2^d}$ be any collection of unit vectors in $\mathbb{R}^n$. } \textup{ Consider the map $f:\mathbb{R}^d\rightarrow \mathbb{R}^n$ defined by \begin{equation} f(x) = \sum_{\beta=1}^{2^d} \varphi(10^4\|x - a_\beta\|_2^2)\eta_\beta \EquationName{37c} \end{equation} Thus by construction, the summands in \Equation{37c} are disjointly supported functions of $x$. Clearly \begin{equation} Df(x)\xi = 2\cdot 10^4\sum_\beta \varphi'(10^4\|x - a_\beta\|_2^2) \inprod{x - a_\beta, \xi}\eta_\beta \EquationName{38c} \end{equation} implying that $$ \|Df(x)\|_{2\rightarrow 2} \le C $$ and \begin{equation} \|Df(x) - Df(y)\|_{2\rightarrow 2} \le C\|x - y\|_2 \EquationName{39c} \end{equation} } \textup{ Next, let $\theta_1,\ldots,\theta_d\in\mathbb{R}^n$ be orthogonal vectors such that \begin{equation} \|\theta_j\|_2 = 1, \|\theta_j\|_\infty = \frac 1{\sqrt{n}} \EquationName{40c} \end{equation} and define the map \begin{align*} &F : B_{\ell_2^d}\subset\mathbb{R}^d \rightarrow \mathbb{R}^{2n} = \mathbb{R}^n\times \mathbb{R}^n\\ &F(x) = \Big(\sum_{j=1}^d x_j\theta_j, f(x)\Big) \end{align*} } \textup{ In view of \Equation{39c}, $F$ satisfies conditions \Eqsub{34c},\Eqsub{35c}. Also, by \Eqsub{38c},\Eqsub{40c} \begin{equation} \alpha \lesssim \sup_{x\in B_{\ell_2^d}} \max_{\|\xi\|_2 = 1}\|DF(x)\xi\|_\infty \lesssim \sqrt{\frac dn} + \max_\beta \|\eta_\beta\|_\infty \EquationName{41c} \end{equation} } \textup{ Let $\mathcal{M} = F(B_{\ell_2^d})$. Fix some $1\le \beta\le 2^d$ and let for $1/200 \le t \le 1/(100\sqrt{2})$ $$ \gamma(t) = F(a_\beta + t e_1) = \Big(\sum_{j=1}^d a_{\beta,j}\theta_j + t\theta_1, 10^4t^2\eta_\beta\Big) $$ where we used \Equation{36c}. Thus $\gamma$ is a $C^1$-curve in $\mathcal{M}$ and \begin{equation} \gamma'\Big(\frac 1{200}\Big) = (\theta_1, 100\eta_{\beta}) \EquationName{42c} \end{equation} } \textup{ Let $\Phi$ be a sparse $m\times 2n$ matrix for which \Equation{28c} holds. Then $\Phi$ has to satisfy in particular $$ (1-\varepsilon)(1+10^4)^{1/2} \le \|\Phi(\theta_1, 100\eta_\beta)\|_2 \le (1+\varepsilon)(1+10^4)^{1/2} $$ and hence \begin{equation} \|\Phi(\theta_1, 100\eta_\beta)\|_2 < 10^3 \text{ for all } 1\le \beta\le 2^d \EquationName{43c} \end{equation} } \textup{ Choose $k$ such that $$ 2^d > \binom{n}{k}2^k,\text{ i.e. } k\simeq \frac d{\log n} $$ and let $(\eta_\beta)$ be the collection of $2^k\binom{n}{k}$ vectors in $\mathbb{R}^n$ of the form \begin{equation} \eta = \frac 1{\sqrt{k}} \sum_{j\in I} \pm e_j\text{ with } I\subset \{1,\ldots,n\}, |I| = k \EquationName{44c} \end{equation} } \textup{ By \Equation{43c}, $\Phi$ needs to satisfy \begin{equation} \|\Phi(0,\eta)\|_2 \le 20 \EquationName{45c} \end{equation} for all vectors $\eta$ of the form \Equation{44c}. But if $m < n/d$, \Equation{45c} implies that $s\ge k/400 \gtrsim d/\log n$. On the other hand, \Equation{41c} gives $$ \alpha \lesssim \sqrt{\frac dn} + \frac 1{\sqrt{k}} \simeq\ \sqrt{\frac{\log n}d} $$ } \end{remark} \subsubsection{Geodesic distances}\SectionName{geodesic} In this section we show that not only do sparse maps preserve curve lengths on manifolds, but in fact they preserve geodesic distances. \begin{lemma}\LemmaName{lem1} Fix $x\in\mathbb{R}^n$ such that $|x_j| \ge 1$ for $j\in J\subset\{1,\ldots,n\}, |J| = r$. Then \begin{equation} \Big|\Big\{ 1\le i\le m: \sum \Phi_{ij}^2 x_j^2 \ge \frac 1s \Big\} \Big| > \min\Big\{\frac{sr}3, cm\Big\} \mathbin{\stackrel{\rm def}{=}} r_1 \end{equation} with probability (with respect to $\Phi$) at least \begin{equation} 1 - 2^{-sr} \end{equation} \end{lemma} \begin{proof} Fix $I\subset \{1,\ldots,m\}$, $|I|\le r_1$ and assume $\sum_j \Phi_{ij}^2 x_j^2 < 1/s$ for $i\notin I$. This means that for any $j\in J$, $S = \mathbin{\stackrel{\rm def}{=}}\{i ; \Phi_{ij} \neq 0\} \subset I$. The probability (with respect to $\Phi$) that $\{i ; \Phi_{ij} \neq 0\} \subset I$ (with $j$ fixed) is (by \Equation{neg-cor}) \begin{align*} \sum_{\substack{K\subset I\\|K|=s}}\Pr_{\Phi}(S = K) &= \sum_{\substack{K\subset I\\|K|=s}} \mathbb{E} \prod_{i\in K} \delta_{ij}\\ {}&\le \sum_{\substack{K\subset I\\|K|=s}} \Big(\frac sm\Big)^s\\ {}&\le \binom{|I|}{s} \cdot \Big(\frac sm\Big)^s\\ {}&\le \Big(\frac{e\cdot |I|}m\Big)^s \end{align*} Since for different $j$ the events are independent, it follows the probability that $$ \Big\{ 1\le i\le m ; \sum \Phi_{ij}^2 x_j^2 \ge \frac 1s \Big\} \subset I $$ is bounded by $((e|I|/m)^s)^r \le (er_1/m)^{sr}$. Taking a union bound over all $I\subset\{1,\ldots,m\},|I|\le r_1$ gives by the choice of $r_1$ $$ \binom{m}{|I|}\cdot \Big(\frac{er_1}m\Big)^{sr} \le \Big(\frac {em}{r_1}\Big)^{r_1} \Big(\frac{er_1}m\Big)^{sr} \le \Big(\frac{e^2r_1}m\Big)^{2sr/3} < 2^{-sr} $$ \end{proof} For the following lemma recall that for $a\in\mathbb{R}^n$ and $(\sigma_j)$ a Rademacher vector \begin{equation} \EquationName{Berger} \Pr_\sigma\Big(\Big|\sum_j a_j \sigma_j\Big| < \frac 12 \|a\|_2\Big) < \frac 45. \end{equation} This is a consequence of the Paley-Zygmund inequality, see e.g.\ \cite[Theorem 3.6]{Berger97}. \begin{lemma}\LemmaName{lem2} Let $x\in\mathbb{R}^n$ satisfy the assumption of \Lemma{lem1}. Then \begin{equation} \Pr_{\Phi}\Big(\|\Phi x\|_2 < \frac 1{2\sqrt{s}}\Big) < 2^{-cr_1} \EquationName{three} \end{equation} \end{lemma} \begin{proof} We may assume that $\sum \Phi_{ij}^2x_j^2 \ge 1/s$ for $i$ in a subset $I\subset \{1,\ldots,m\}$, $|I| \ge r_1$. Exploiting the random signs of $\Phi_{ij}$, we find by \Equation{Berger} for each $i\in I$ \begin{equation} \Pr\Big(\Big|\sum_j \Phi_{ij}x_j\Big| < \frac 1{2\sqrt{s}}\Big) < \frac 45\EquationName{four} \end{equation} If $\|\Phi x\|_2 < 1/(2\sqrt{s})$, then $|\sum_j \Phi_{ij}x_j| < 1/(2\sqrt{s})$ for all $i$, in particular for all $i\in I$. By \Equation{four} the probability for this event is at most $$ \Big(\frac 45\Big)^{|I|} < 2^{-cr_1} , $$ proving \Equation{three}. \end{proof} \begin{corollary}\CorollaryName{cor1} Let $\xi\subset\mathbb{R}^n$ be a set of vectors $x$ with following properties: \begin{enumerate} \item Each $x\in\xi$ has a decomposition $x = x' + x''$, $x'\in \xi'$ and there is a set $J = J_{x'}\subset \{1,\ldots,n\}$, $|J|\ge r$ so that $|x_j'| \ge 1$ for $j\in J$. Moreover \begin{equation} \|x''\|_2 < \frac 1{10n} \EquationName{smallnorm} \end{equation} \item \begin{equation}|\xi'| < 2^{c\min\{sr, m\}}\EquationName{xiprime}\end{equation} \end{enumerate} Then \begin{equation} \|\Phi x\|_2 > \frac 1{4\sqrt{s}}\text{ for all } x\in\xi \EquationName{normlb} \end{equation} with probability at least $1 - 2^{-c\min\{sr,m\}}$. \end{corollary} \begin{proof} Write $$ \|\Phi x\|_2 \ge \|\Phi x'\|_2 - \|\Phi x''\|_2 \ge \|\Phi x'\|_2 - \sqrt{n}\cdot \|x''\|_2 \ge \|\Phi x'\|_2 - \frac 1{10\sqrt{n}} . $$ since clearly $\|\Phi\|_2 \le \sqrt{n}$. Next, \Lemma{lem2} and a union bound will ensure that $\|\Phi x'\|_2 \ge 1/(2\sqrt{s})$ for all $x'\in\xi'$ with the desired probability. \end{proof} \begin{lemma}\LemmaName{smalllb} Let $s\ge c(\log n)^2$. Let $\xi\subset \mathbb{R}^n$ be a finite set of unit vectors satisfying \begin{equation} \log|\xi|< cm \end{equation} Then with probability at least $1 - e^{-cs}$ for some constant $c>0$, $\Phi$ satisfies \begin{equation} \|\Phi x\|> e^{-c(\log n)^2}\text{ for } x\in\xi . \end{equation} \end{lemma} \begin{proof} Given $x\in\mathbb{R}^n$, let $x^*$ be the decreasing rearrangement of $|x_i|$. Let $K = c\log n$. We partition $\xi$ as $$ \xi = \bigcup_{\ell} (\xi_\ell\backslash (\xi_0\cup \xi_1 \cup \ldots \cup \xi_{\ell-1})) $$ where $\xi_{-1}=\emptyset$, and for $\ell\geq 0$ satisfying $2^\ell K^2 < m$, $$\xi_\ell = \{x \in\xi : x^*_{m/(2^\ell K^2)} > n^{-K + 2\ell}\}$$ For the vectors $x\in\xi_0$, apply \Corollary{cor1} with $x=x', r = m/K^2$ (after rescaling by $n^{-K}$). Since $csr > cm> \log|\xi| \ge \log|\xi_0|$, \Eqsub{xiprime} holds and by \Equation{normlb}, we ensure that $\|\Phi x\|_2 > n^{-K}/(4\sqrt{s})$ for all $x\in\xi_0$. Consider the set $\xi_\ell' = \xi_\ell\backslash(\xi_0\cup\xi_1 \cup\ldots\cup\xi_{\ell-1})$. Thus each $x\in\xi_\ell'$ has a decomposition $$ x = y + z $$ where $y$ is obtained by considering the $m/(2^{\ell-1}K^2)$ largest coordinates of $x$ and $\|z\|_\infty\le x^*_{m/(2^{\ell-1}K^2)} \le n^{-K+2\ell-2}$ since $x\notin \xi_{\ell-1}$. Note also that $$ y\in \bigcup_{\substack{S\subset\{1,\ldots,n\}\\|S| = \frac m{2^{\ell-1}K^2}}} B_{X_S}, \text{ with } X_S = \{e_j : j\in S\} . $$ Let $\mathfrak{f}_S \subset B_{X_S}$ be a finite subset such that \begin{align} \nonumber \mathrm{dist}(y, \mathfrak{f}_S) &< n^{-K}\text{ for all } y\in B_{X_S}\\ \log|\mathfrak{f}_S| &\lesssim \frac{m}{2^{\ell-1}K^2}\log n^K = \frac{m\log n}{2^{\ell-1}K} \EquationName{finite} \end{align} Hence $y = x' + w$ with \begin{equation} x' \in \bigcup_S \mathfrak{f}_S,\ \|w\|_2 < n^{-K} \EquationName{contained} \end{equation} Apply \Corollary{cor1} to the set $n^{K - 2\ell}\xi'_\ell$, considering the decomposition $$ n^{K-2\ell}x = n^{K-2\ell}x' + n^{K-2\ell}(w + z) $$ satisfying by \Equation{contained} \begin{align*} \|n^{K - 2\ell}(w+z)\|_2 &< n^{-2\ell} + n^{K-2\ell}\|z\|_\infty\\ {}&< n^{-2\ell} + n^{-2} < \frac 1{10 n} \end{align*} which is condition \Eqsub{smallnorm}. Also $n^{K-2\ell}x'\in\xi'$, where by \Equation{finite} (and choice of $s,K$) $$ \log|\xi'| \le \log\binom{n}{m/(2^{\ell-1}K^2)} + \frac{m\log n}{2^{\ell-1}K} \lesssim \frac{m\log n}{2^{\ell-1}K} < c\min\Big\{\frac{sm}{2^\ell K^2}, m\Big\} , $$ which is condition \Eqsub{xiprime} with $r = m/(2^{\ell}K^2)$. Let $J_x$ be the set of $r = m/(2^\ell K^2)$ largest coordinates of $x$ (which are also the $r$ largest coordinates of $y$). Hence, for $j\in J_x$ \begin{align*} n^{K-2\ell}|x_j'| &\ge n^{K-2\ell}|y_j| - n^{K-2\ell}|w_j|\\ {}&= n^{K-2\ell}|x_j| - n^{K-2\ell}|w_j|\\ {}& > n^{K-2\ell}n^{-K+2\ell} - n^{K-2\ell}n^{-K} > \frac 12 \end{align*} since $x\in\xi_\ell$. By \Equation{normlb} $$ \|\Phi x\| > n^{-K+2\ell}\cdot \frac 1{4\sqrt{s}}\text{ for all } x\in\xi_\ell $$ with probability at least $1 - e^{-c\min\{m, sr\}} \geq 1 - e^{-cs}$, since $m/(2^\ell K^2) \ge 1$. \end{proof} Let $\mathcal{M}\subset\mathbb{R}^n$ be a $d$-dimensional manifold obtained as a bi-Lipschitz image of the unit ball $B_{\ell_2^d}$, i.e.\ $F:B_{\ell_2^d}\rightarrow \mathcal{M}$ satisfies \begin{equation} \|F(x) - F(y)\|_2 \simeq \|x - y\|_2 .\EquationName{bilip} \end{equation} We assume moreover $F$ is smooth, more specifically $DF:B_{\ell_2^d}\rightarrow \mathcal{L}(\mathbb{R}^d, \mathbb{R}^n)$ (i.e.\ linear maps from $\mathbb{R}^d$ to $\mathbb{R}^n$ under operator norm) is Lipschitz, i.e., \begin{equation} \|DF(x) - DF(y)\|_{\ell_2^d\rightarrow\ell_2^n} \lesssim \|x - y\|_2\EquationName{lipschitz-derivative} \end{equation} \begin{lemma} Let $\mathcal{M}$ be as above. Assume \begin{align} \nonumber m&\gtrsim d(\log n)^2\\ s&\gtrsim (\log n)^2 \EquationName{parameters} \end{align} Then with probability at least $1 - e^{-cs}$ for some constant $c>0$, $\Phi|_{\mathcal{M}}$ is bi-Lipschitz, and more specifically \begin{equation} \|\Phi(x) - \Phi(y)\|_2 \ge e^{-c(\log n)^2}\cdot \|x-y\|_2\text{ for }x,y\in\mathcal{M} \EquationName{terrible-distortion} \end{equation} \end{lemma} \begin{proof} We treat separately the pairs $x,y\in\mathcal{M}$ which are at a ``large'' and ``small'' distance from each other. Fix $\varepsilon_1 > 0$ and $\varepsilon_2>\varepsilon_1$ to be specified later. Let $A_{\varepsilon_1} \subset\mathcal{M}$ be an $\varepsilon_1$-net for $\mathcal{M}$. By \Equation{bilip} and \Equation{volComp}, we can assume that $$ \log|A_{\varepsilon_1}| \lesssim d\log(1/\varepsilon_1). $$ Assume that $x,y\in\mathcal{M}$, $\|x-y\|_2>\varepsilon_2$. Take $x_1,y_1\in A_{\varepsilon_1}$ s.t.\ $\|x - x_1\|_2 < \varepsilon_1$, $\|y - y_1\|_2 < \varepsilon_1$. Since $\Phi$ is linear $$ \|\Phi(x) - \Phi(y)\|_2 \ge \|\Phi(x_1 - y_1)\|_2 - 2\sqrt{n}\varepsilon_1 $$ Apply \Lemma{smalllb} to the set $$ \xi = \Big\{\frac{x_1 - y_1}{\|x_1 - y_1\|_2} : x_1,y_1 \in A_{\varepsilon_1}\Big\} . $$ Assuming \begin{equation} m \gtrsim d\log(1/\varepsilon_1)> c\log|A_{\varepsilon_1}|\EquationName{basic-entropy} \end{equation} we ensure $\Phi$ to satisfy $$ \Big\|\Phi\Big(\frac{x_1 - y_1}{\|x_1 - y_1\|_2}\Big)\Big\|_2 > e^{-c(\log n)^2}\text{ for }x_1,y_1\in A_{\varepsilon_1} . $$ Therefore, if $x,y\in\mathcal{M}$, $\|x - y\|_2 > \varepsilon_2$ \begin{align} \nonumber \|\Phi(x) - \Phi(y)\|_2 &\ge e^{-c(\log n)^2}\cdot \|x_1 - y_1\|_2 - 2\sqrt{n}\frac{\varepsilon_1}{\varepsilon_2}\cdot\|x-y\|_2\\ {}&\ge \frac 12 e^{-c(\log n)^2}\cdot\|x - y\|_2 \end{align} choosing $\varepsilon_2 = 5\sqrt{n} e^{c(\log n)^2}\varepsilon_1$. This takes care of large distances. In order to deal with small distances, we first ensure that \begin{equation} \|\Phi(v)\|> e^{-c(\log n)^2}\text{ for any unit tangent vector }v\text{ of } \mathcal{M} \EquationName{tangents} \end{equation} Consider \begin{equation} \mathcal{T}_{\varepsilon_1} = \bigcup_{x_1\in A_{\varepsilon_1}} \Big\{v\in T_{x_1}: \|v\| = 1\Big\}\EquationName{all-tangents} \end{equation} Under the assumption \Eqsub{basic-entropy} on $m$ (and making an $\varepsilon_1$-discretization of each $\{v\in T_{x_1} : \|v\|_2 = 1\}$), another application of \Lemma{smalllb} will ensure that \Eqsub{tangents} holds for all $v\in\mathcal{T}_{\varepsilon_1}$. Next, if $x\in\mathcal{M}$, $x_1\in A_{\varepsilon_1}$, $\|x - x_1\|_2 < \varepsilon_1$, it follows from \Equation{lipschitz-derivative} that $\|DF(x) - DF(x_1)\|_{\ell_2\rightarrow\ell_2} \lesssim \varepsilon_1$ and hence $\rho_{\mathrm{Fin}}(T_x, T_{x_1}) \lesssim \varepsilon_1$. Therefore, if $v\in T_x$, $\|v\|_2 = 1$, there is $v_1\in\mathcal{T}_{\varepsilon_1}$ s.t.\ $\|v - v_1\|_2 \lesssim \varepsilon_1$. Therefore $$ \|\Phi(v)\|_2 \ge \|\Phi(v_1)\|_2 - \sqrt{n}\varepsilon_1 \ge e^{-c(\log n)^2} - \sqrt{n}\varepsilon_1 > \frac 12 e^{-c(\log n)^2} $$ since $\varepsilon_1 < n^{-1/2} e^{-c(\log n)^2}/2$. Thus $\Phi$ satisfies \Eqsub{tangents}. Assume $x,y\in\mathcal{M}$, $\|x-y\|_2<\varepsilon_2$. Let $u = F^{-1}(x), w = F^{-1}(y) \in B_{\ell_2^d}$. By \Equation{bilip}, $\|u-w\|_2 \simeq \|x-y\|_2$. Write \begin{align*} y-x = F(w) - F(u) &= \int_0^1 \partial_t F(tw + (1-t)u) dt\\ {}&= \int_0^1 DF(tw + (1-t)u)(w - u)dt. \end{align*} Hence, again invoking \Equation{lipschitz-derivative}, \begin{align} & \|F(w) - F(u) - DF(u)(w-u)\|_2 \nonumber \\ & \qquad \le \|w-u\|_2\cdot \sup_t \|DF(tw + (1-t)u) - DF(u)\|_{\ell_2\rightarrow\ell_2} \nonumber\\ & \qquad \lesssim \|u - w\|_2^2. \EquationName{squared-dist} \end{align} Denote $$ v = DF(u)\Big(\frac{w-u}{\|w-u\|_2}\Big) \in T_u . $$ By \Equation{squared-dist} and using $\|\Phi\|\leq \sqrt{n}$, \begin{align*} \|y - x - \|u-w\|_2 v\|_2 &< c\varepsilon_2\|x-y\|_2\\ \|\Phi(y) - \Phi(x) - \|u-w\|_2\Phi(v)\|_2 &< c\sqrt{n}\varepsilon_2\|x-y\|_2 \end{align*} and by \Eqsub{tangents} \begin{align*} \|\Phi(x) - \Phi(y)\|_2 &\ge \|u-w\|_2 e^{-c(\log n)^2} - c\sqrt{n}\varepsilon_2\|x-y\|_2\\ {}&\ge \Big(ce^{-c(\log n)^2} - c\sqrt{n}\varepsilon_2\Big)\|x-y\|_2\\ {}&\gtrsim e^{-c(\log n)^2}\|x-y\|_2 \end{align*} provided \begin{equation} \varepsilon_2 < c\frac 1{\sqrt{n}} e^{-c(\log n)^2} \end{equation} Thus we may take $\log(1/\varepsilon_1),\log(1/\varepsilon_2)\simeq (\log n)^2$ and condition \Eqsub{basic-entropy} will hold for $$ m\gtrsim d(\log n)^2. $$ \end{proof} \begin{theorem} Let $\mathcal{M}$ be as above and $m,s$ satisfy \Eqsub{parameters}. Assume moreover that $m,s$ satisfy the appropriate conditions to ensure that \begin{equation} 1-\varepsilon \le \|\Phi(v)\|_2 \le 1+\varepsilon \EquationName{curves} \end{equation} for all unit tangent vectors $v$ of $\mathcal{M}$. Then with probability at least $1 - e^{-cs}$ for some constant $c>0$, $\Phi$ preserves geodesic distances up to factor $1+\varepsilon$. \end{theorem} \begin{proof} By \Eqsub{curves}, $(1-\varepsilon)|\gamma| < |\Phi(\gamma)| < (1+\varepsilon)|\gamma|$ for any $C^1$-curve $\gamma$ in $\mathcal{M}$. Let $\rho_{\mathcal{M}}$ refer to the geodesic distance in $\mathcal{M}$. Clearly \begin{equation} \rho_{\Phi(\mathcal{M})}(\Phi(x), \Phi(y)) \le (1+\varepsilon)\rho_{\mathcal{M}}(x,y) \end{equation} from the above. We need to show the reverse inequality. Let $\gamma_1$ be a $C^1$-curve in $\Phi(\mathcal{M})$ joining $\Phi(x),\Phi(y)$ such that $\rho_{\Phi(\mathcal{M})}(\Phi(x), \Phi(y)) = |\gamma_1|$. Since $\Phi$ satisfies \Eqsub{terrible-distortion}, $\Phi|_{\mathcal{M}} : \mathcal{M}\rightarrow\Phi(\mathcal{M})$ is bi-Lipschitz and hence a diffeomorphism (since $\Phi$ is also smooth). It follows that $\gamma = \Phi^{-1} \gamma_1$ is a $C^1$-curve joining $x$ and $y$ and $$ \rho_{\mathcal{M}}(x,y) \le |\gamma| \le (1+\varepsilon) |\Phi(\gamma)| = (1+\varepsilon)|\gamma_1| = (1+\varepsilon)\rho_{\Phi(\mathcal{M})}(\Phi(x), \Phi(y)) . $$ \end{proof} \section{Discussion}\SectionName{discussion} We have provided a general theorem which captures sparse dimensionality reduction in Euclidean space and qualitatively unifies much of what we know in specific applications. There is still much room though for quantitative improvement. We here list some known quantitative shortcomings of our bounds and discuss some avenues for improvement in future work. First, our dependence on $1/\varepsilon$ in $s$ in all our theorems is, up to logarithmic factors, quadratic. Meanwhile the works \cite{KN14,NN13b} show that the correct dependence in the case of small $m$ should be linear. Part of the reason for this discrepancy may be our use of the chaining result of \cite{KMR14}. Specifically, chaining is a technique that in general converts tail bounds into bounds on the expected supremum of stochastic processes (see \cite{Talagrand05} for details). Perhaps one could generalize the tail bound of \cite{KN14} to give a broad understanding of the decay behavior for the error random variable in the SJLT as a function of $s,m$ then feed such a quantity into a chaining argument to improve our use of \cite{KMR14}. Another place where we lost logarithmic factors is in our use of the duality of entropy numbers \cite[Proposition 4]{BPST89} in \Equation{3.3}. It is believed that in general if $K,D$ are symmetric convex bodies and $\mathcal{N}(K,D)$ is the number of translations of $D$ needed to cover $K$ then $$ \mathcal{N}(D^\circ, a K^\circ)\lesssim \mathcal{N}(K, D) \lesssim \mathcal{N}(D^\circ, a^{-1} K^\circ) $$ for some universal constant $a>0$. This is known as Pietsch's duality conjecture \cite[p.\ 38]{P72}, and unfortunately resolving it has been a challenging open problem in convex geometry for over 40 years. We have also lost logarithmic factors in our use of dual Sudakov minoration, which bounds $\sup_{t>0} t [\log \mathcal{N}(B_E, \|\cdot\|_X, t)]^{1/2}$ by a constant times $\mathbb{E}_g \|P_E g\|_X$ for any norm $\|\cdot\|_X$ even though what we actually wish to bound is the $\gamma_2$ functional. Passing from this functional via \Equation{Dudley} to $\sup_{t>0} t [\log \mathcal{N}(B_E, \|\cdot\|_X, t)]^{1/2}$ costs us logarithmic factors. The majorizing measures theory (see \cite{Talagrand05}) shows that the loss can be avoided when $\|\cdot\|_X$ is the $\ell_2$ norm; it would be interesting to explore how the loss can be avoided in our case. Finally, note that the SJLT $\Phi$ considered in this work is a randomly signed adjacency matrix of a random bipartite graph with $m$ vertices in the left bipartition, $n$ in the right, and with all right vertices having equal degree $s$. Sasha Sodin has asked in personal communication whether taking a random signing of the adjacency matrix of a random biregular graph (degree $s$ on the right and degree $ns/m$ on the left) can yield improved bounds on $s,m$ for some $T$ of interest. We leave this as an interesting open question. \bigskip \paragraph{Acknowledgement} We would like to thank Huy L.\ Nguy$\tilde{\hat{\mbox{e}}}$n for pointing out the second statement in Lemma~\ref{lem:compare}. Part of this work was done during a visit of the second-named author to Harvard University. He wishes to thank the Theory of Computation group for its hospitality. \newcommand{\etalchar}[1]{$^{#1}$}
\section{Introduction}\label{sec:INTRODUCTION} The weak coupling expansion of the QCD pressure --- a central quantity for thermodynamics --- shows few signs of apparent convergence, and only at extremely high temperatures (see e.g.~\cite{ConvergenceWeakCouplingExpansion_and_BlaizotDR}). It is therefore crucial to try to improve the situation by attempting to continue the perturbative results as close to the phase transition region as possible. The importance of such work is highlighted by the fact that the quark-gluon plasma created in heavy ion experiments is argued~\cite{SatzHeavyIonCollisions,MullerQCDMatter} to reach a temperature of only a few times the pseudo-critical temperature of the deconfinement transition, $T_\trmi{c}=154\pm 9$MeV~\cite{PeikertTc,BazavovTc}, where the plasma is clearly not extremely weakly coupled. When the quark chemical potentials $\mu_f$ are turned on to study the phase diagram of the theory (see \cite{RHIC,LHC} for present and \cite{FAIR1,FAIR2,NICA} for future collider experiments), lattice Monte Carlo simulations become inapplicable due to the so-called sign problem of QCD \cite{ForcrandSignPB,GuptaSignPB}. There have been many attempts to find a solution to this problem, but so far the most successful approach amounts to merely Taylor expanding the pressure in powers of $\mu_f/T$, giving access to only moderate densities. The coefficients of such Taylor expansions are, however, related to cumulants of globally conserved quantum numbers in the system, and thus give information about their correlations and fluctuations. Consequently, they turn to be very practical probes of changes occurring in the degrees of freedom of the system across the transition region. The limitation of lattice methods on the $T$-$\mu$ plane is one of the main reasons, why it is interesting to attack the problem of fluctuations and correlations of conserved charges --- and ultimately the equation of state --- from a resummed perturbation theory point of view. Using such a first principle framework, where there is no sign problem, these quantities can be extended to very large values of the ratio $\mu_f/T$, hence giving information about how the quark-gluon plasma behaves, as it becomes more and more dense. In this proceedings contribution, I will present recent results for two quantities related to cumulants, obtained using two different frameworks of resummed perturbation theory. I will further comment on the insights that these results can provide for heavy ion physics, as they display highly non-trivial and interesting features. \section{Statistical QCD and cumulants}\label{sec:RELEVANCEHIC} Given the Hamiltonian $H_\trmi{QCD}$ of Quantum Chromodynamics, the corresponding partition function $Z_\trmi{QCD}$ can be written in the form \beq Z_\trmi{QCD}\left(T,\mu_f;V\right) \equiv \mbox{Tr\ }\exp\left[-\frac{1}{T}\Bigg(H_\trmi{QCD}-\sum_f \mu_f\ Q_f\Bigg)\right] = \mbox{Tr\ }\Big(\rho_\trmi{QCD}\Big)\, \label{PartitionFunction} , \eeq where we have employed the usual notation for thermal averages, $\left\langle\vartheta\right\rangle\equiv\mbox{Tr\ }\left(\vartheta\cdot\rho_\trmi{QCD}\right)/Z_\trmi{QCD}$, that can also be expressed using a path integral representation for $Z_\trmi{QCD}$. Here, $Q_f$ and $\mu_f$ denote respectively conserved charges and the corresponding chemical potentials. In the present work, we will mainly consider the up, down and strange quark conserved charges (with their chemical potentials $\mu_\trmi{u},\ \mu_\trmi{d},\ \mu_\trmi{s}$). However, one can equivalently express the partition function in term of the baryon, electric charge and strangeness conserved numbers (with their chemical potentials $\mu_\trmi{B},\ \mu_\trmi{Q},\ \mu_\trmi{S}$). In addition, the partition function gives access to a host of equilibrium thermodynamic quantities such as the pressure and entropy, following the relations \beqa p_\trmi{QCD}&=& \frac{\partial \left(T \log Z_\trmi{QCD}\right)}{\partial V}\ \xrightarrow[V\rightarrow\infty]{ }\ \frac{T}{V}\,\log Z_\trmi{QCD} \, ,\\ S_\trmi{QCD}&=& \frac{\partial \left(T \log Z_\trmi{QCD}\right)}{\partial T} \, . \eeqa From equation~(\ref{PartitionFunction}), it is easy to see that the mean and (co)variance of two conserved charges, i.e. $\left\langle Q_f \right\rangle$ and $\left\langle \left(Q_f-\left\langle Q_f \right\rangle\right) \cdot \left(Q_g-\left\langle Q_g \right\rangle\right) \right\rangle$, can be expressed in terms of derivatives with respect to the corresponding chemical potentials. Thus, we can write \beqa \left\langle Q_f \right\rangle&=&T\,\frac{\partial}{\partial \mu_f} \log Z_\trmi{QCD} \, , \label{Mean} \\ \left\langle \left(Q_f-\left\langle Q_f \right\rangle\right) \cdot \left(Q_g-\left\langle Q_g \right\rangle\right) \right\rangle&=&T^2\,\frac{\partial^2}{\partial \mu_f \partial \mu_g} \log Z_\trmi{QCD} \, , \label{Covariance} \eeqa defining quantities called the first and second order cumulants of the related conserved charges. Based on the above, it is clear that when considering the covariance of two different charges, such as $Q_\trmi{u}$ and $Q_\trmi{d}$, one is probing the correlation between the two related flavors, while when considering the variance of the same charge, one is probing its fluctuations. The former quantity gives information about possible bound-state survival (see~\cite{BOUNDSTATESURVIVAL1,BOUNDSTATESURVIVAL2} and references therein), while the latter is related to how the system reacts to small increases of density. It also follows that in the infinite volume limit, equations~(\ref{Mean}) and~(\ref{Covariance}) relate the cumulants to the equation of state of the system via derivatives with respect to chemical potentials. These quantities, typically referred to as susceptibilities, are denoted by \beq \chi_\trmi{u$_i$\,d$_j$\,s$_k$\, ...}\left(T,\left\{\mu_f\right\}\right) \equiv \frac{\partial^n p_\trmi{QCD}\left(T,\left\{\mu_f\right\}\right)} {\partial\mu_\trmi{u}^i\, \partial\mu_\trmi{d}^j \, \partial\mu_\trmi{s}^k\, ...} \ , \eeq where $n=i+j+k+...$ and where instead of quark numbers one may also consider any other conserved charge. The existence and location of a possible critical point on the phase diagram of QCD is often investigated through the behavior of cumulants. Indeed, the variance of the baryon number $\chi_\trmi{B$_2$}\left(T,\mu_\trmi{B}\right)$ is expected to display a critical-like behavior with a sharp peak at the critical point, as it is highly sensitive to changes of density in the medium. This quantity can be estimated via a Taylor expansion containing higher order cumulants at vanishing chemical potential (see e.g.~\cite{SimonTwoFlavors} for discussion and results in the two flavor case). Also, as the convergence radius of the Taylor series of $\chi_\trmi{B$_2$}\left(T,\mu_\trmi{B}\right)$ has been argued to determine the location of a possible critical point (see~\cite{Sayantan} for a recent article), it is clearly interesting to consider higher order cumulants at vanishing chemical potential. For some recent reviews on heavy ion collisions and in particular the use of cumulants in statistical QCD, the reader is referred to~\cite{SatzHeavyIonCollisions,KochCumulant}. \section{Resummed perturbative QCD}\label{sec:FRAMEWORKS} \subsection{Dimensional Reduction inspired resummation} It is well known~\cite{DimensionalReductionPhenomenon1,DimensionalReductionPhenomenon2} that dimensional reduction allows the dynamics of the energy scales of order $gT$ and smaller to be properly accounted for through an effective field theory, which for QCD is called Electrostatic QCD (EQCD). Indeed, by integrating out the hard degrees of freedom, one obtains a three-dimensional SU($N_\trmi{c}$) Yang-Mills theory coupled to an adjoint Higgs field~\cite{BraatenNietoEQCD,KajantieEQCD}. This effective theory turns out to account for all of the infrared divergences typically encountered in naive weak coupling expansions~\cite{LindeIRcatastrophe}. Thus, EQCD not only provides a way out of the so-called infrared catastrophe, but in addition allows for weak coupling calculations to be carried out to high orders in perturbation theory. This, of course, requires a careful matching procedure of the effective theory with full QCD. Taking advantage of the above fact, one can write the QCD pressure in the form \beq p_\trmi{QCD}= p_\trmi{hard}(\mbox{\small$g$}) + T\, p_\trmi{EQCD}(\mbox{\small$m_\trmi{E},g_\trmi{E},\lambda_\trmi{E},\zeta$}) \, , \eeq where the EQCD parameters $m_\trmi{E}(g),g_\trmi{E}(g),\lambda_\trmi{E}(g)$ and $\zeta(g)$ are in turn functions of temperature and the quark chemical potentials. They admit expansions in powers of the four-dimensional gauge coupling $g$. The function $p_\trmi{hard}$ gives the contribution of the hard scale ($\sim T$) only, and is available through a strict loop expansion in QCD, while $p_\trmi{EQCD}$ gives the contribution of the soft ($\sim gT$) and the ultrasoft ($\sim g^2T$) scales, being accessible from the partition function of EQCD. Typically, when computing a physical quantity through EQCD, one expands the EQCD parameters --- and ultimately the entire result --- in powers of $g$. However, as first suggested in~\cite{ConvergenceWeakCouplingExpansion_and_BlaizotDR} and later successfully applied to the case of the pressure at zero chemical potential in~\cite{MikkoYorkQuarkMassThresholds}, one can alternatively simply consider both $p_\trmi{hard}$ and $p_\trmi{EQCD}$ functions of the EQCD parameters, and not expand them in $g$. This indeed resums certain higher order contributions to the pressure, containing all correct contributions up to and including order $g^6 \log g$~\cite{KeijoMikkoYork,Aleksi1} when expanded in powers of $g$. This procedure leads to a considerable decrease in the theoretical uncertainties of the result through a substantial reduction of its renormalization scale dependence, and thus also to an improvement in its convergence properties. In the following, we will denote the results originating from the above procedure the `DR ones'. More details about the implementation of the resummation at finite chemical potentials can be found in~\cite{Paper2}, while for general details about dimensional reduction and EQCD, the reader is referred to the original references~\cite{BraatenNietoEQCD,KajantieEQCD} as well as to~\cite{Aleksi1,HartLainePhilipsen,Ipp,AleksiPhDThesis} for a generalization to finite density. \subsection{Hard-thermal-loop perturbation theory} Another way of resumming important higher order contributions to thermodynamic quantities is to make use of a variationally improved perturbation theory framework (see~\cite{RefOptKneurAndOthers,RefOptPeterForSPT} and references therein). The basic idea of such a framework is to introduce a physically relevant (usually mass) term that is added and subtracted from the Lagrangian density of the original theory. Treating the added piece with the non-interacting part of the action while adding the subtracted one to the interaction terms of it, one ends up interpolating between the original theory and a theory having dressed propagators and vertices. The procedure is of course such that working consistently to a given order in a perturbative expansion, one in the end recovers results corresponding to the correct original theory. In a gauge field theory such as QCD, the situation is somewhat more complicated, as one cannot simply add and subtract a local mass term for the gauge fields without breaking the gauge invariance. Instead, one can make use of the well-known hard-thermal-loop effective action, first derived in~\cite{FrenkelTaylorHTL,BraatenPisarskiHTL}, which plays the role of a non-local (and hence momentum dependent) mass term. With the QCD pressure, this approach was first applied more than a decade ago in~\cite{JensMikeBraatenFirstHTLpt,JensMikeBraatenFirstHTLptbis,JensMikeFirstTwoLoop}, and is known as Hard-Thermal-Loop perturbation theory (HTLpt). In short, HTLpt amounts to reorganizing the Lagrangian density of QCD as \beq {\cal L_{\rm HTLpt}}=\Big[{\cal L}_{\rm QCD}+(1-\delta)\ {\cal L}_{\rm HTL}\Big]\Big|_{g\rightarrow\sqrt{\delta}g}+\Delta{\cal L}_{\rm HTL} \, , \eeq where ${\cal L}_{\rm HTL}$ is a gauge invariant HTL improvement term, given by the HTL effective action, and $\delta$ a formal expansion parameter set to one after having Taylor expanded the path integral up to some order in it. Finally, $\Delta{\cal L}_{\rm HTL}$ is a counter term necessary to cancel further ultraviolet divergences introduced by the resulting reorganization of the perturbative series. Essentially, this reorganization amounts to shifting the ground state of the expansion from an ideal gas of massless particles to an ideal gas of massive quasiparticles~\cite{HTLptFiniteMUThreeLoop,NanHTLpt1,NanHTLpt2}. More details about the implementation of the HTLpt procedure at finite chemical potentials can be found in~\cite{Paper1,Paper2} as well as in~\cite{HTLptFiniteMUThreeLoop,HTLptFiniteMUTwoLoop1,HTLptFiniteMUTwoLoop2}. For related developments, see also~\cite{BlaizotHTLResummation1,BlaizotHTLResummation2}. \subsection{Technical details} Working at four- and one-loop levels in the DR and HTLpt schemes, respectively, we must assign values to various parameters within them. As is customary in perturbative thermal QCD calculations, the renormalization scale is chosen to be of the order of $2\pi T$, and then varied by a factor of two around this central value in order to estimate the sensitivity of the results with respect to this choice. Following~\cite{FAC}, we choose to apply the Fastest Apparent Convergence (FAC) scheme to the next-to-leading order expression of the gauge coupling of EQCD. Thereby, we obtain $\bar{\Lambda}_{\rm central}=1.445\times 2\pi T$ and $1.291\times 2\pi T$ for the central values with two and three flavors, respectively~\cite{Paper1,Paper2}. Regarding the running of the coupling, we use a two-loop expression in the DR case~\cite{MikkoYorkTwoLoopG} and one-loop running in the HTLpt result. Finally, to fix the value of $\Lambda_{\rm QCD}$ (in the $\overline{\trmi{MS}}$ scheme), we use the recent lattice result $\alpha_\trmi{s}(1.5\;{\rm GeV})=0.326$~\cite{AlphaS}, demanding that our one- and two-loop couplings agree with it for $\bar{\Lambda}=1.5$ GeV. This respectively gives $\Lambda_{\rm QCD}=176$ MeV and 283 MeV for $N_\trmi{f}=3$, as well as $\Lambda_{\rm QCD}=204$ MeV and 324 MeV for $N_\trmi{f}=2$. These values are then varied by $\pm$ 30 MeV. In the plots appearing in the following section, the bands corresponding to our perturbative results are composed by varying all of the parameters mentioned here in their respective ranges. \section{Results}\label{sec:RESULTS} \subsection{Kurtoses} Roughly speaking, the kurtosis is a measure of how strongly peaked a given quantity is, and is defined as the ratio of the corresponding fourth and second order cumulants, the latter one squared. In our figure~\ref{ratios}, we display results for this quantity for the (light) quark and baryon numbers; we further multiply the results by the second order cumulants, so that in effect we simply plot ratios of the fourth and second order cumulants. Such a quantity can be used as a measure of how fast a phase transition is; see e.g.~the discussion of \cite{StephanovKurtosis} with regards to the behavior of the kurtosis relevant to the order parameter of the deconfinement phase transition near the critical point. In figure~\ref{ratios} (left), we plot our DR and HTLpt results together with lattice data from~\cite{RatioBNLB1,RatioBNLB2,RatioWuppBudLight}, which at temperatures around $T=$ 300 -- 400 MeV seem to agree nicely with the one-loop HTLpt band, but soon start approaching the DR prediction. It should be noted that the latter reproduces the overall trend of the lattice data better. On the right figure, we see similar trends in both predictions, with the difference that the DR one seems to converge to the Stefan Boltzmann limit much faster than in the case of the light quark kurtosis. This trend is seen to be in accordance with the displayed lattice data from~\cite{RatioBNLB2,RatioWuppBudBaryon}, and agrees with the expectation that the medium is much less sensitive to the hadronic degrees of freedom in this range of temperatures. Moreover, in figure~\ref{ratios} (right), we display the corresponding three-loop HTLpt prediction obtained using the cumulants computed in~\cite{HTLptFiniteMUThreeLoop}. Although the band is relatively large at low temperature, this result is also seen to reproduce the qualitative trend of the lattice data, furthermore agreeing quite well with the DR one. From the one-loop to the three-loop HTLpt prediction, we notice a reasonably good convergence of the series for this quantity. \begin{figure}[!ht]\centering\includegraphics[scale=0.30]{chiu4_over_chiu2_Nf3.pdf}\!\!\!\!\includegraphics[scale=0.30]{chiB4_over_chiB2_Nf3.pdf} \caption{Ratios of cumulants related to the light quark (left) and baryon (right) numbers, with lattice data taken from~\cite{RatioBNLB1,RatioBNLB2} (BNL-B) and~\cite{RatioWuppBudLight,RatioWuppBudBaryon} (WB). The three-loop HTLpt result is obtained from the corresponding cumulants of~\cite{HTLptFiniteMUThreeLoop}. The dashed curves inside the bands correspond to the central values of the renormalization and QCD scales, while the black dashed (straight) lines denote the Stefan Boltzmann limits.\label{ratios}} \end{figure} \subsection{Sixth order cumulants} In figure~\ref{chi6Nf3and2}, we next display sixth order cumulants for the light quark number in the cases of three (left) and two (right) quark flavors. Such higher order cumulants are expected to be very sensitive probes of the hadronic freeze-out~\cite{KarschFluctu}. For these quantities, the two first orders of perturbation theory vanish, and the weak coupling expansions only start at order $g^3$, rendering the properties of the weak coupling series much poorer than for the lower order cumulants. Indeed, we see from both the left and right figures that our DR predictions are positive for all displayed temperatures, while the HTLpt results are consistently negative. The latter apparently agree better with the (unfortunately not continuum extrapolated) lattice data of~\cite{SimonTwoFlavors,Peterchi6}. However, it should be recalled our HTLpt predictions are of one-loop order, and based on the very good agreement between our DR results and the recent three-loop HTLpt calculations of~\cite{HTLptFiniteMUThreeLoop} (see~\cite{Paper2} for a detailed comparison), it is probable that the forthcoming three-loop HTLpt prediction of~\cite{THREELOOPHTLptforth} for the sixth order cumulant will also turn out to be positive. It is interesting to note that the leading plasmon contribution to the weak coupling expansion of the sixth order cumulant is negative, while it can be seen by truncating our DR results at various orders that the sign of the result is turned positive by the higher order contributions. Noting that lattice studies seem to favor negative values for sixth order cumulants~\cite{SimonTwoFlavors,KarschFluctu,Peterchi6} above the transition region, we suspect that including further perturbative orders to the current DR result, it will eventually change in the temperature range considered. This will, however, likely not be verified in the near future, as the technical challenges involved in e.g.~a full five-loop determination of the quantity are formidable. \begin{figure}[!ht]\centering\includegraphics[scale=0.30]{chiu6_Nf3.pdf}\!\!\!\!\includegraphics[scale=0.30]{chiu6_Nf2.pdf} \caption{The sixth order cumulant for the light quark number, evaluated for $N_\trmi{f}=3$ (left) and $N_\trmi{f}=2$ (right), with lattice data from~\cite{Peterchi6} (left) and~\cite{SimonTwoFlavors} (right). The right figure has been taken from~\cite{Paper2}. The dashed curves inside the bands correspond to the central values of the renormalization and QCD scales.\label{chi6Nf3and2}} \end{figure} \section{Conclusion}\label{sec:CONLUSION} In this proceedings contribution, which is based on the recent resummed perturbation theory investigations~\cite{Paper1,Paper2}, we have reviewed predictions for two quantities highly relevant for the physics of heavy ion collisions. First, considering kurtoses corresponding to the light quark and baryon number operators for three flavor QCD, we saw how resummed perturbation theory provides a very successful description of lattice data down to 2 -- 3 $T_\trmi{c}$. This is clearly an indication that a weakly interacting quasiparticles picture is well suited for the fermionic sector of the theory, even close to the transition region. Also, information on the relevant degrees of freedom was highlighted by the qualitative difference of fluctuations between light quark and baryon conserved charges. Next, we looked at sixth order cumulants in the cases of two and three flavor QCD. Here, we encountered quantities with which resummed perturbation theory has a much harder time, which could be understood from the fact that the weak coupling expansions only start at a relatively high order. Quantities like the ones considered here are important for understanding the phase transition leading to the formation of the quark-gluon plasma, and also give crucial information about changes in the degrees of freedom of the system right above the transition. I therefore hope that these results, and more generally resummed perturbation theory as a complementary first principle tool, will turn out to be of practical use in the present and future analysis of heavy ion collision data. \ack I am grateful to P\'eter Petreczky and Aleksi Vuorinen for valuable discussions, helpful in the preparation of this manuscript. I would also like to thank Nan Su for introducing this workshop to me, as well as the Helsinki Institute of Physics for its hospitality, where this proceedings contribution was written. This work has been supported by the mobility grant of the Bielefeld Graduate School in Theoretical Sciences. \section*{References}
\section{Introduction.} Given a potential $V:\mathbb{R}\to \mathbb{R}$ and $\beta>0$, we consider the $\beta$-matrix model \begin{equation}\label{betamodel}\mathbb P^N_V(d\lambda_1,\ldots,d\lambda_N):=\frac{1}{Z^N_V}\prod_{1\le i<j\le N}|\lambda_i-\lambda_j|^\beta e^{-N\sum_{i=1}^N V(\lambda_i)} \,d\lambda_1\dots\,d\lambda_N,\end{equation} where $Z_V^N:=\int \prod_{1\le i<j\le N}|\lambda_i-\lambda_j|^\beta e^{-N\sum_{i=1}^N V(\lambda_i)} \,d\lambda_1\dots\,d\lambda_N$. It is well known (see e.g. \cite{Vi}) that for any $V,W:\mathbb{R}\to\mathbb{R}$ such that $Z_V^N,Z_{V+W}^N<\infty$ there exists maps $T^N:\mathbb R^N{\rightarrow}\mathbb R^N$ that transport $\mathbb P^N_V$ into $\mathbb P^N_{V+W}$, that is, for any bounded measurable function $f:\mathbb R^N{\rightarrow}\mathbb R$ we have $$\int f\circ T^N (\lambda_1,\ldots,\lambda_N) \,\mathbb P^N_V(d\lambda_1,\ldots,d\lambda_N)=\int f(\lambda_1,\ldots,\lambda_N) \, \mathbb P^N_{V+W}(d\lambda_1,\ldots,d\lambda_N)\,.$$ However, the dependency in the dimension $N$ of this transport map is in general unclear unless one makes very strong assumptions on the densities \cite{Caff} that unfortunately are never satisfied in our situation. Hence, it seems extremely difficult to use these maps $T^N$ to understand the relation between the asymptotic of the two models. The main contribution of this paper is to show that a variant of this approach is indeed possible and provides a very robust and flexible method to compare the asymptotics of local statistics. In the more general context of several-matrices models, it was shown in \cite{GS} that the maps $T^N$ are asymptotically well approximated by a function of matrices independent of $N$, but it was left open the question of studying corrections to this limit. In this article we consider one-matrix models, and more precisely their generalization given by $\beta$-models, and we construct {\it approximate} transport maps with a very precise dependence on the dimension. This allows us to compare local fluctuations of the eigenvalues and show universality.\\ Local fluctuations were first studied in the case $\beta=2$. After the work of Mehta \cite{ME} where the sine kernel law was exhibited for the GUE, Pastur and Shcherbina \cite{PS} proved universality for $V\in C^3$. The Riemann Hilbert approach was then developed in this context in \cite{DKMcLVZ2} for analytic potentials. The paper \cite{LL} provides the most general result for universality in the case $\beta=2$. In the case $\beta=1,4$, universality was proved in \cite{DeGibulk} in the bulk, and \cite{DeGiedge} at the edge, for monomials $V$ (see \cite{DeGi2} for a review). For general one-cut potentials, the first proof of universality was given in \cite{Shch} in the case $\beta=1$, whereas \cite{KrSh} treated the case $\beta=4$ in the one-cut case. The multi-cut case was then considered in \cite{Sh}. The local fluctuations of more general $\beta$-ensembles were only derived recently \cite{VV,RRV} in the Gaussian case. Universality in the $\beta$-ensembles was first addressed in \cite{BEY1} (in the bulk, $\beta> 0$, $V\in C^4$), then in \cite{BEY2} (at the edge, $\beta\ge 1$, $V\in C^4$) and \cite{KRV} (at the edge, $\beta>0$, $V$ convex polynomial) and finally in \cite{Shch} (in the bulk, $\beta>0$, $V$ analytic, multi-cut case included). The paper \cite{Shch} and the current one were completed essentially at the same time, and the method developed in \cite{Shch} is closely related with the one followed in the present work in the sense that it uses a transport map, but still they are very different. Indeed, \cite{Shch} uses a global transport map $T^N=T_0^{\otimes N}$, independent of the dimension, which maps the probability $\mathbb P^N_{V}$ to an intermediate probability measure $\tilde {\mathbb P}^N_{V+W}$ which is absolutely continuous with respect to $\mathbb P^N_{V+W}$ but with a non-trivial density, and then shows that this density does not perturb the local fluctuations of local statistics. On the other hand, in the present paper the map $T^N$ is constructed to map $\mathbb P^N_{V}$ to an intermediate probability measure $\tilde {\mathbb P}^N_{V+W}$ with density with respect to $\mathbb P^N_{V+W}$ going uniformly to one: the dependence of $T_N$ in the dimension $N$ then allows us to retrieve the universality of the local fluctuations. \\ The main idea of this paper is that we can build a smooth approximate transport map which at the first order is simply a tensor product, and then it has first order corrections of order $N^{-1}$ (see Theorem \ref{thm:T}). This continues the long standing idea developed in loop equations theory to study the correlation functions of $\beta$-matrix models by clever change of variables, see e.g. \cite{johansson, KrSh, AM, BEMP}. On the contrary to loop equations, the change of variable has to be of order one rather than infinitesimal. The first order term of our map is simply the monotone transport map between the asymptotic equilibrium measures; then, corrections to this first order are constructed so that densities match up to a priori smaller fluctuating terms. As we shall see, our transport map is constructed as the flow of a vector field obtained by approximately solving a linearized Jacobian equation and then making a suitable ansatz on the structure of the vector field. The errors are controlled by deriving bounds on covariances and correlations functions thanks to loop equations, allowing us to obtain a self-contained proof of universality (see Theorem \ref{thm:univ}). Although optimal transportation will never be really used, it will provide us the correct intuition in order to solve this problem (see Sections \ref{sect:MA2} and \ref{sect:approx}). We notice that this last step could also be performed either by using directly central limit theorems, see e.g. \cite{Shch,borot-guionnet,Shc12}, or local limit laws \cite{BEY1,BEY2}. However, our approach has the advantage of being pretty robust and should generalize to many other mean field models. In particular, it should be possible to generalize it to the several-matrices models, at least in the perturbative regime considered in \cite{GS}, but this would not solve yet the question of universality as the transport map would be a non-commutative function of several matrices. Even in the case of GUE matrices, there are not yet any results about the local statistics of the eigenvalues of such non-commutative functions, except for a few very specific cases. \\ We now describe our results in detail. Given a potential $V:\mathbb R\to \mathbb R$, we consider the probability measure \eqref{betamodel}. We assume that $V$ goes to infinity faster than $\beta\log |x|$ (that is $V(x)/\beta\log |x|\to +\infty$ as $|x|\to +\infty$) so that in particular $Z^N_V$ is finite. We will use $\mu_V$ to denote the equilibrium measure, which is obtained as limit of the spectral measure and is characterized as the unique minimizer (among probability measures) of \begin{equation}\label{entr}I_V(\mu):=\frac{1}{2} \int \bigl( V(x)+V(y) -\beta \log |x-y|\bigr)d\mu(x)d\mu(y)\,.\end{equation} We assume hereafter that another smooth potential $W$ is given so that $V+W$ goes to infinity faster than $\beta\log |x|$. We set $V_t:=V+tW$, and we shall make the following assumption: \begin{hypo}\label{hypo1} We assume that $\mu_{V_0}$ and $\mu_{V_1}$ have a connected support and are non-critical, that is, there exists a constant $\bar c>0$ such that, for $t=0,1$, $$\frac{d\mu_{V_t}}{dx}=S_t(x)\sqrt{ (x-a_t)(b_t-x)} \qquad \text{with $S_t \geq \bar c$ a.e. on $[a_t,b_t]$.}$$ \end{hypo} \begin{rem}{\rm The assumption of a connected support could be removed here, following the lines of \cite{Shch,borot-guionnet2}. Indeed, only a generalization of Lemma \ref{central} is required, which is not difficult. However, the non-criticality assumption cannot be removed easily, as criticality would result in singularities in the transport map.} \end{rem} Finally, we assume that the eigenvalues stay in a neighborhood of the support $[a_t-{\epsilon},b_t+{\epsilon}]$ with large enough $\mathbb P^N_V$-probability, that is with probability greater than $1-C\,N^{-p}$ for some $p$ large enough. By \cite[Lemma 3.1]{borot-guionnet2}, the latter is fulfilled as soon as: \begin{hypo} \label{hypo2} For $t=0,1$, \begin{equation}\label{noncr} U_{V_t}(x):=V_t(x)-\beta\int d\mu_{V_t}(y)\log |x-y|\end{equation} achieves its minimal value on $[a,b]^c$ at its boundary $\{a,b\}$ \end{hypo} All these assumptions are verified for instance if $V_t$ is uniformly convex for $t=0,1$. \\ The main goal of this article is to build an approximate transport map between $\mathbb P^N_V$ and $\mathbb P^N_{V+W}$: more precisely, we construct a map $T^N:\mathbb R^N{\rightarrow}\mathbb R^N$ such that, for any bounded measurable function $\chi$, \begin{equation}\label{aptr}\left|\int \chi\circ T^N\, d\mathbb P^N_{V}-\int \chi\, d\mathbb P^N_{V+W}\right|\le C\,\frac{(\log N)^3 }{{N}} \|\chi\|_\infty \end{equation} for some constant $C$ independent of $N$, and which has a very precise expansion in the dimension (in the following result, $T_0:\mathbb{R}\to\mathbb{R}$ is a smooth transport map of $\mu_{V}$ onto $\mu_{V+W}$, see Section \ref{sect:flow}): \begin{thm} \label{thm:T} Assume that $V',W$ are of class $C^{30}$ and satisfy Hypotheses \ref{hypo1} and \ref{hypo2}. Then there exists a map $T^N=(T^{N,1},\ldots,T^{N,N}):\mathbb R^N{\rightarrow}\mathbb R^N$ which satisfies \eqref{aptr} and has the form $$T^{N,i}(\hat\lambda)= T_0(\lambda_i) +\frac{1}{N} T_1^{N,i}(\hat\lambda)\quad \forall\,i=1,\ldots,N,\qquad \hat\lambda:=(\lambda_1,\ldots,\lambda_N),$$ where $T_0:\mathbb{R}\to \mathbb{R}$ and $T_1^{N,i}:\mathbb{R}^N\to \mathbb{R}$ are smooth and satisfy uniform (in $N$) regularity estimates. More precisely, $T^N$ is of class $C^{23}$ and we have the decomposition $T^{N,i}_1= X^{N,i}_1+\frac{1}{N} X^{N,i}_2$ where \begin{equation}\label{boi1int} \sup_{1\leq k \leq N}\|X_{1}^{N,k}\|_{L^4(\mathbb P_{V}^N)} \leq C\log N,\qquad \|X_{2}^{N}\|_{L^2(\mathbb P_{V}^N)} \leq CN^{1/2}(\log N)^2, \end{equation}for some constant $C>0$ independent of $N$. In addition, with probability greater than $1- N^{-N/C}$, \begin{equation}\label{boi2int}\max_{1\le k,k'\le N} |X_{1}^{N,k}(\lambda)-X_{1}^{N,k'}(\lambda)|\le C\,\log N \sqrt{N}|\lambda_k-\lambda_{k'}|.\end{equation} \end{thm} As we shall see in Section \ref{sect:univ}, this result implies universality as follows (compare with \cite[Theorem 2.4]{BEY2}): \begin{thm} \label{thm:univ} Assume $V',W\in C^{30}$, and let $T_0$ be as in Theorem \ref{thm:T} above. Denote $\tilde P_V^N$ the distribution of the increasingly ordered eigenvalues $\lambda_i$ under $\mathbb P_V^N$. There exists a constant $\hat C>0$, independent of $N$, such that the following two facts hold true: \begin{enumerate} \item Let $M\in (0,\infty)$ and $m\in \mathbb N$. For any Lipschitz function $f:\mathbb{R}^m\to \mathbb{R}$ supported inside $[-M,M]^m$, \begin{multline*} \bigg|\int f\bigl(N(\lambda_{i+1}-\lambda_i),\ldots,N(\lambda_{i+m}-\lambda_i)\bigr)\, d\tilde P_{V+W}^N\\ \qquad -\int f\bigl(T_{0}'(\lambda_i)N(\lambda_{i+1}-\lambda_i),\ldots,T_{0}'(\lambda_i)N(\lambda_{i+m}-\lambda_i)\bigr) d\tilde P_{V}^N\bigg|\\ \le \hat C \,\frac{(\log N)^3 }{N} \|f\|_\infty + \hat C\,\biggl(\sqrt{m}\,\frac{(\log N)^2}{N^{1/2}}+M\,\frac{\log N}{N^{1/2}} +\frac{M^2}{N}\biggr) \|\nabla f\|_{\infty}. \end{multline*} \item Let $a_V$ (resp. $a_{V+W}$) denote the smallest point in the support of $\mu_V$ (resp. $\mu_{V+W}$), so that ${\rm supp}(\mu_V)\subset [a_V,\infty)$ (resp. ${\rm supp}(\mu_{V+W})\subset [a_{V+W},\infty)$). Let $M\in (0,\infty)$. Then, for any Lipschitz function $f:\mathbb{R}^m\to \mathbb{R}$ supported inside $[-M,M]^m$, \begin{multline*} \bigg| \int f\bigl(N^{2/3}(\lambda_1-a_{V+W}), \ldots,N^{2/3}(\lambda_m-a_{V+W})\bigr) \,d\tilde P_{V+W}^N\\ -\int f\Bigl(N^{2/3}T_{0}'(a_V)\bigl(\lambda_1-a_V\bigr), \ldots,N^{2/3}T_{0}'(a_V)\bigl(\lambda_m-a_V\bigr)\Bigr)\, d\tilde P_{V}^N\bigg|\\ \le \hat C \,\frac{(\log N)^3}{{N}} \|f\|_\infty+ \hat C\biggl(\sqrt{m}\,\frac{(\log N)^2}{N^{5/6}}+M\,\frac{\log N}{N^{5/6}} +\frac{M^2}{N^{4/3}} + \frac{\log N}{N^{1/3}}\biggr) \|\nabla f\|_{\infty} . \end{multline*} The same bound holds around the largest point in the support of $\mu_V$. \end{enumerate} \end{thm} \begin{rem} {\em The condition that $V',W\in C^{30}$ in the theorem above is clearly non-optimal (compare with \cite{BEY1}). For instance, by using Stieltjes transform instead of Fourier transform in some of our estimates, we could reduce the regularity assumptions on $V',W$ to $C^{21}$ by a slightly more cumbersome proof. In addition, by using \cite[Theorem 2.4]{BEY2} we could also weaken our regularity assumptions in Theorem \ref{thm:T}, as we could use that result to estimate the error terms in Section \ref{sect:error}. However, the main point of this hypothesis for us is to stress that we do not need to have analytic potentials, as often required in matrix models theory. Moreover, under this assumption we can provide self-contained and short proofs of Theorems \ref{thm:T} and \ref{thm:univ}. } \end{rem} Note that Theorem \ref{thm:T} is well suited to prove universality of the spacings distribution in the bulk as stated in Theorem \ref{thm:univ}, but it is not clear how to directly deduce the universality of the rescaled density, see e.g \cite[Theorem 2.5(i)]{BEY1}. Indeed, this corresponds to choosing test functions whose uniform norm blows up like some power of the dimension, so to apply Theorem \ref{thm:T} we should have an a priori control on the numbers of eigenvalues inside sets of size of order $N^{-1}$ under both $\mathbb P^N_{V}$ and $\mathbb P^N_{V+W}$. Notice however that \cite[Theorem 2.5(ii)]{BEY1} requires $\beta \geq 1$, while our results hold for any $\beta>0$. \cite{KRV} holds for $\beta>0$ but convex polynomial. In particular, the edge universality proved in Theorem \ref{thm:univ}(2) is new for $\beta \in (0,1)$ and $V \in C^{30}$ which is not a polynomial. In addition our strategy is very robust and flexible. For instance, although we shall not pursue this direction here, it looks likely to us that one could use it prove the universality of the asymptotics of the law of $\{N(\lambda_i-x) \}_{1\le i\le m}$ under $\tilde P_V^N$ for given $i$ and $x$, as in \cite{BY14,BY14B}. Indeed, $N(\lambda_i-x)$ can be seen via the approximate transport map to be given by the same kind of difference under the source measure (e.g. the Gaussian one), up to a term coming from the transport map $T^{N,i}_1$. Proving that $T^{N,i}_1$ does not depend on $i$ locally would suffice to prove universality, as the sine-kernel law is translation invariant. \\ The paper is structured as follows: In Section \ref{sect:MA} we describe the general strategy to construct our transport map as the flow of vector fields obtained by approximately solving a linearization of the Monge-Amp\`ere equation (see \eqref{eq:laplace}). As we shall explain there, this idea comes from optimal transport theory. In Section \ref{sect:laplace} we make an ansatz on the structure of an approximate solution to \eqref{eq:laplace} and we show that our ansatz actually provides a smooth solution which enjoys very nice regularity estimates that are uniform as $N \to \infty$. In Section \ref{sect:flow} we reconstruct the approximate transport map from $\mathbb P^N_V$ to $\mathbb P^N_{V+W}$ via a flow argument. The estimates proved in this section will be crucial in Section \ref{sect:univ} to show universality.\\ \textit{Acknowledgments:} AF was partially supported by NSF Grant DMS-1262411. AG was partially supported by the Simons Foundation and by NSF Grant DMS-1307704. Both authors are thankful to the anonymous referees for several useful comments that helped to improve the presentation of the paper. \section{Approximate Monge-Amp\`ere equation} \label{sect:MA} \subsection{Propagating the hypotheses} The central idea of the paper is to build transport maps as flows, and in fact to build transport maps between $\mathbb P_V^N$ and $\mathbb P_{V_t}^N$ where $t\mapsto V_t$ is a smooth function so that $V_0=V$, $V_1=V+W$. In order to have a good interpolation between $V$ and $V+W$, it will be convenient to assume that the support of the two equilibrium measures $\mu_V$ and $\mu_{V+W}$ (see \eqref{entr}) are the same. This can always be done up to an affine transformation. Indeed, if $L:\mathbb{R}\to \mathbb{R}$ is the affine transformation which maps $[a_1,b_1]$ (the support of $\mu_{V_1}$) onto $[a_0,b_0]$ (the support of $\mu_{V_0}$), we first construct a transport map from $ \mathbb P_{V}^N$ to $L^{\otimes N}_\sharp \mathbb P_{V+W}^N=\mathbb P_N^{V+\tilde W}$ where \begin{equation}\label{pol} \tilde W=V\circ L^{-1} +W\circ L^{-1} -V,\end{equation} and then we simply compose our transport map with $(L^{-1})^{\otimes N}$ to get the desired map from $ \mathbb P_{V}^N$ to $\mathbb P_{V+W}^N$. Hence, without loss of generality we will hereafter assume that $\mu_V$ and $\mu_{V+W}$ have the same support. We then consider the interpolation $\mu_{V_t}$ with $V_t=V+tW$, $t\in [0,1]$. We have: \begin{lem}\label{propa} If Hypotheses \ref{hypo1} and \ref{hypo2} are fulfilled for $t=0,1$, then Hypothesis \ref{hypo1} is also fulfilled for all $t\in [0,1]$. Moreover, we may assume without loss of generality that $V$ goes to infinity as fast as we want up to modify $\mathbb P_{V}^N$ and $\mathbb P_{V+W}^N$ by a negligible error (in total variation). \end{lem} \begin{proof} Let $\Sigma$ denote the support of $\mu_V$ and $\mu_{V+W}$. Following \cite[Lemma 5.1]{borot-guionnet}, the measure $\mu_{V_t}$ is simply given by $$\mu_{V_t}=(1-t)\mu_V+t \mu_{V+W}.$$ Indeed, $\mu_V$ is uniquely determined by the fact that there exists a constant $c$ such that $$\beta\int\log |x-y|d\mu_V(x)-V\le c$$ with equality on the support of $\mu_V$, and this property extends to linear combinations. As a consequence the support of $\mu_{V_t}$ is $\Sigma$, and its density is bounded away from zero on $\Sigma$. This shows that Hypothesis \ref{hypo1} is fulfilled for all $t\in [0,1]$. Furthermore, we can modify $\mathbb P_{V}^N$ and $\mathbb P_{V+W}^N$ outside an open neighborhood of $\Sigma$ without changing the final result, as eigenvalues will quit this neighborhood only with very small probability under our assumption of non-criticality according to the large deviation estimates developped in \cite{BAGD,borot-guionnet} and culminating in \cite{borot-guionnet2} as follows: \begin{align*} \limsup_{N{\rightarrow}\infty}\frac{1}{N}\ln \mathbb P^{V}_{N}\left[\exists\, i\,:\,\lambda_i \in F\right] & \le -\frac{\beta}{2}\,\inf_{x \in F} \tilde{U}_{V}(x), \\ \liminf_{N{\rightarrow}\infty}\frac{1}{N}\ln \mathbb P^{V}_{N}\left[\exists\, i\,:\,\lambda_i \in \Omega\right] & \ge -\frac{\beta}{2}\,\inf_{x \in \Omega} \tilde{U}_{V}(x). \end{align*} where $\tilde{U}_V:=U_V-\inf U_V$, and $U_V$ is defined as in \eqref{hypo2}. \end{proof} Thanks to the above lemma and the discussion immediately before it, we can assume that $\mu_{V}$ and $\mu_{V+W}$ have the same support, that $W$ is bounded, and that $V$ goes to infinity faster than $x^{p}$ for some $p>0$ large enough. \subsection{Monge-Amp\`ere equation} \label{sect:MA2} Given the two probability densities $\mathbb P^N_{V_t}$ to $\mathbb P^N_{V_s}$ as in \eqref{betamodel} with $0 \leq t \leq s \leq 1$, by optimal transport theory it is well-known that there exists a (convex) function $\phi_{t,s}^N$ such that $\nabla \phi_{t,s}^N$ pushes forward $\mathbb P^N_{V_t}$ onto $\mathbb P^N_{V_s}$ and which satisfies the Monge-Amp\`ere equation $$ \det(D^2 \phi_{t,s}^N)=\frac{\rho_t}{\rho_s(\nabla \phi_{t,s}^N)},\qquad \rho_\tau:=\frac{d\mathbb P_{V_\tau}^N}{d\lambda_1\ldots\,d\lambda_N} $$ (see for instance \cite[Chapters 3 and 4]{Vi} or the recent survey paper \cite{DF} for an account on optimal transport theory and its link to the Monge-Amp\`ere equation). Because $\phi_{t,t}(x)=|x|^2/2$ (since $\nabla\phi_{t,t}$ is the identity map), we can differentiate the above equation with respect to $s$ and set $s=t$ to get \begin{equation} \label{eq:laplace}\Delta \psi^N_t= c_t^N-\beta\sum_{i<j}\frac{\partial_i \psi^N_t-\partial_j\psi^N_t}{\lambda_i-\lambda_j} +N\sum_i W(\lambda_i) +N \sum_iV_t'(\lambda_i)\partial_i \psi^N_t, \end{equation} where $\psi_t^N:=\partial_s\phi_{t,s}^N|_{s=t}$ and $$c^N_t:=-N\int \sum_iW(\lambda_i) \,d\mathbb P^N_{V_t}= \partial_t \log Z^N_{V_t}\,.$$ Although this is a formal argument, it suggests to us a way to construct maps $T_{0,t}^N:\mathbb{R}^N\to\mathbb{R}^N$ sending $\mathbb P_{V}^N$ onto $\mathbb P_{V_t}^N$: indeed, if $T_{0,t}^N$ sends $\mathbb P_{V}^N$ onto $\mathbb P_{V_t}^N$ then $\nabla \phi_{t,s}^N\circ T_{0,t}^N$ sends $\mathbb P_{V}^N$ onto $\mathbb P_{V_s}^N$. Hence, we may try to find $T_{0,s}^N$ of the form $T_{0,s}^N=\nabla \phi_{t,s}^N\circ T_{0,t}^N+o(s-t)$. By differentiating this relation with respect to $s$ and setting $s=t$ we obtain $\partial_t T^N_{0,t}=\nabla \psi^N_t(T^N_{0,t})$. Thus, to construct a transport map $T^N$ from $\mathbb P^N_{V}$ onto $\mathbb P^N_{V+W}$ we could first try to find $\psi_t^N$ by solving \eqref{eq:laplace}, and then construct $T^N$ solving the ODE $\dot X^N_t=\nabla \psi^N_t(X^N_t)$ and setting $T^N:=X_1^N$. We notice that, in general, $T^N$ is not an optimal transport map for the quadratic cost. Unfortunately, finding an exact solution of \eqref{eq:laplace} enjoying ``nice'' regularity estimates that are uniform in $N$ seems extremely difficult. So, instead, we make an ansatz on the structure of $\psi_t^N$ (see \eqref{eq:psiN} below): the idea is that at first order eigenvalues do not interact, then at order $1/N$ eigenvalues interact at most by pairs, and so on. As we shall see, in order to construct a function which enjoys nice regularity estimates and satisfies \eqref{eq:laplace} up to a error that goes to zero as $N\to \infty$, it will be enough to stop the expansion at $1/N$. Actually, while the argument before provides us the right intuition, we notice that there is no need to assume that the vector field generating the flow $X_t^N$ is a gradient, so we will consider general vector fields ${\mbox{{\bf Y}}}_t^N=({\mbox{{\bf Y}}}_{1,t}^N,\ldots,{\mbox{{\bf Y}}}_{N,t}^N):\mathbb{R}^N\to\mathbb{R}^N$ that approximately solve \begin{equation} \label{eq:laplace2} {\rm div} {\mbox{{\bf Y}}}^N_t= c_t^N-\beta\sum_{i<j}\frac{{\mbox{{\bf Y}}}^N_{i,t}-{\mbox{{\bf Y}}}^N_{j,t}}{\lambda_i-\lambda_j} +N\sum_i W(\lambda_i) +N \sum_i V_t'(\lambda_i) {\mbox{{\bf Y}}}^N_{i,t}, \end{equation} We begin by checking that the flow of an approximate solution of \eqref{eq:laplace2} gives an approximate transport map. \subsection{Approximate Jacobian equation} \label{sect:approx} Here we show that if a $C^1$ vector field ${\mbox{{\bf Y}}}^N_t$ approximately satisfies \eqref{eq:laplace2}, then its flow $$\dot X^N_t={\mbox{{\bf Y}}}^N_t(X^N_t),\qquad X^N_0=\operatorname{Id},$$ produces almost a transport map. More precisely, let ${\mbox{{\bf Y}}}^N_t:\mathbb{R}^N\to\mathbb{R}^N$ be a smooth vector field and denote $${\mathcal R}^N_t({\mbox{{\bf Y}}}^N):= c_t^N-\beta\sum_{i<j}\frac{{\mbox{{\bf Y}}}^N_{i,t}-{\mbox{{\bf Y}}}^N_{j,t}}{\lambda_i-\lambda_j} +N\sum_i W(\lambda_i) +N \sum_i V_t'(\lambda_i) {\mbox{{\bf Y}}}^N_{i,t}-{\rm div} {\mbox{{\bf Y}}}^N_{t}.$$ \begin{lem} \label{lemma:flow} Let $\chi:\mathbb R^N\to \mathbb R$ be a bounded measurable function, and let $X^N_t$ be the flow of ${\mbox{{\bf Y}}}^N_{t}$. Then $$\left|\int \chi(X^N_t)\, d\mathbb P^N_V- \int \chi\, d\mathbb P^N_{V_t}\right|\le\|\chi\|_\infty\int_0^t \|{\mathcal R}^N_s({\mbox{{\bf Y}}}^N)\|_{L^1(\mathbb P^N_{V_s})}ds.$$ \end{lem} \begin{proof} Since ${\mbox{{\bf Y}}}^N_t\in C^1$, by Cauchy-Lipschitz Theorem its flow is a bi-Lipchitz homeomorphism. If $JX^N_t$ denotes the Jacobian of $X^N_t$ and $\rho_t$ the density of $\mathbb P^N_{V_t}$, by the change of variable formula it follows that $$\int \chi\, d\mathbb P^N_{V_t}= \int \chi(X^N_t)JX^N_t \rho_t(X^N_t) \,dx$$ thus \begin{equation}\label{coco} \left|\int \chi(X^N_t) \,d\mathbb P^N_V- \int \chi\, d\mathbb P^N_{V_t}\right|\le\|\chi\|_\infty \int |\rho_0-JX^N_t \rho_t(X^N_t)| \,dx=:\|\chi\|_\infty A_t\end{equation} Using that $\partial_t(JX^N_t )={\rm div} {\mbox{{\bf Y}}}^N_{t}\, JX^N_t$ and that the derivative of the norm is smaller than the norm of the derivative, we get \begin{align*} |\partial_t A_t|&\le\int \Bigl|\partial_t \bigl(JX^N_t \rho_t(X^N_t)\bigr)\Bigr| \,dx\\ &= \int|{\rm div} {\mbox{{\bf Y}}}^N_{t}\, JX^N_t\, \rho_t(X^N_t)+JX^N_t \,(\partial_t\rho_t)(X^N_t)+JX^N_t\,\nabla \rho_t(X^N_t)\cdot \partial_t X^N_t|\,dx\\ &=\int |{\mathcal R}^N_t({\mbox{{\bf Y}}})|(X^N_t)\,JX^N_t \,\rho_t(X^N_t)\,dx\\ &=\int |{\mathcal R}^N_t({\mbox{{\bf Y}}})|\,d\mathbb P^N_{V_t}. \end{align*} Integrating the above estimate in time completes the proof. \end{proof} By taking the supremum over all functions $\chi$ with $\|\chi\|_\infty \leq 1$, the lemma above gives: \begin{cor}\label{cortr} Let $X^N_t$ be the flow of ${\mbox{{\bf Y}}}_t^N$, and set $\hat {\mathbb P}_t^N:=(X^N_t)_\#\mathbb P^N_V$ the image of $\mathbb P^N_V$ by $X^N_t$. Then $$\|\hat{\mathbb P}_t^N - \mathbb P^N_{V_t}\|_{TV}\leq \int_0^t \|{\mathcal R}^N_s({\mbox{{\bf Y}}}^N)\|_{L^1(\mathbb P^N_{V_s})}ds.$$ \end{cor} \section{Constructing an approximate solution to \eqref{eq:laplace}} \label{sect:laplace} Fix $t \in [0,1]$ and define the random measures \begin{equation} \label{defemp} L_N:=\frac{1}{N}\sum_{i}\delta_{\lambda_i}\quad\mbox{ and }\quad M_N:=\sum_i\delta_{\lambda_i}-N\mu_{V_t}.\end{equation} As we explained in the previous section, a natural ansatz to find an approximate solution of \eqref{eq:laplace} is given by \begin{equation} \label{eq:psiN} \psi^N_t(\lambda_1,\ldots,\lambda_N):=\int \Bigl[\psi_{0,t}(x) +\frac{1}{N} \psi_{1,t} (x) \Bigr]\, dM_N(x) + \frac{1}{2N}\iint \psi_{2,t} (x,y) \,dM_N(x)\,dM_N(y), \end{equation} where (without loss of generality) we assume that $ \psi_{2,t} (x,y) = \psi_{2,t} (y,x)$. Since we do not want to use gradient of functions but general vector fields (as this gives us more flexibility), in order to find an ansatz for an approximate solution of \eqref{eq:laplace2} we compute first the gradient of $\psi$: $$ \partial_i \psi^N_{t}=\psi_{0,t}'(\lambda_i)+\frac{1}{N}\psi_{1,t}'(\lambda_i)+ \frac{1}{N}\xi^N_{1,t}(\lambda_i,M_N),\qquad \xi^N_{1,t}(x,M_N):=\int \partial_1\psi_{2,t}(x,y)\,dM_N(y). $$ This suggests us the following ansatz for the components of ${\mbox{{\bf Y}}}^N_t$: \begin{equation} \label{eq:YN} {\mbox{{\bf Y}}}^N_{i,t}(\lambda_1,\ldots,\lambda_N):={\mbox{{\bf y}}}_{0,t}(\lambda_i)+\frac{1}{N}{\mbox{{\bf y}}}_{1,t}(\lambda_i)+ \frac{1}{N}{\mbox{\boldmath$\xi$}}_{t}(\lambda_i,M_N),\quad {\mbox{\boldmath$\xi$}}_{t}(x,M_N):=\int {\mbox{{\bf z}}}_t(x,y)\,dM_N(y), \end{equation} for some functions ${\mbox{{\bf y}}}_{0,t},{\mbox{{\bf y}}}_{1,t}:\mathbb{R}\to\mathbb{R}$, ${\mbox{{\bf z}}}_t:\mathbb{R}^2\to \mathbb{R}$. Here and in the following, given a function of two variables $\psi,$ we write $\psi\in C^{s,v}$ to denote that it is $s$ times continuously differentiable with respect to the first variable and $v$ times with respect to the second. The aim of this section is to prove the following result: \begin{prop}\label{prop:key}Assume $V',W\in C^{r}$ with $r\ge 30$. Then, there exist ${\mbox{{\bf y}}}_{0,t}\in C^{r-2}$, ${\mbox{{\bf y}}}_{1,t}\in C^{r-8}$, and ${\mbox{{\bf z}}}_t\in C^{s,v}$ for $s+v \leq r-5$, such that $${\mathcal R}^N_t:= \biggl( c_t^N-\beta\sum_{i<j}\frac{{\mbox{{\bf Y}}}^N_{i,t}-{\mbox{{\bf Y}}}^N_{j,t}}{\lambda_i-\lambda_j} +N\sum_i W(\lambda_i) +N \sum_i V_t'(\lambda_i) {\mbox{{\bf Y}}}^N_{i,t}\biggr)-{\rm div} {\mbox{{\bf Y}}}^N_{t}$$ satisfies $$\|{\mathcal R}^N_t\|_{L^1({\mathbb P^N_{V_t}})}\le C\,\frac{(\log N)^3 }{N}$$ for some positive constant $C$ independent of $t\in [0,1]$. \end{prop} The proof of this proposition is pretty involved, and will take the rest of the section. \subsection{Finding an equation for ${\mbox{{\bf y}}}_{0,t}, {\mbox{{\bf y}}}_{1,t},{\mbox{{\bf z}}}_{t}$.} Using \eqref{eq:YN} we compute $$ {\rm div}{\mbox{{\bf Y}}}^N_t=N\int {\mbox{{\bf y}}}_{0,t}'(x)\,dL_N(x)+\int {\mbox{{\bf y}}}_{1,t}'(\lambda)\,dL_N(x) +\int \partial_1{\mbox{\boldmath$\xi$}}_{t}(x,M_N)\,dL_N(x) +{\mbox{\boldmath$\eta$}}(L_N), $$ where, given a measure $\nu$, we set \begin{align*} {\mbox{\boldmath$\eta$}}(\nu):=\int\partial_{2}{\mbox{{\bf z}}}_{t}(y,y)\,d\nu(y). \end{align*} Therefore, recalling that $M_N=N (L_N-\mu_{V_t})$, we get \begin{align*} {\mathcal R}^N_t&= -\frac{\beta N^2}{2} \iint\frac{{\mbox{{\bf y}}}_{0,t}(x)-{\mbox{{\bf y}}}_{0,t}(y)}{x-y}\,dL_N(x)\,dL_N(y) +N^2\int V'_t \,{\mbox{{\bf y}}}_{0,t} \,dL_N+N^2 \int W \,dL_N\\ & -\frac{\beta N}{2} \iint\frac{{\mbox{{\bf y}}}_{1,t}(x)-{\mbox{{\bf y}}}_{1,t}(y)}{x-y}\,dL_N(x)\,dL_N(y) +N\int V'_t \,{\mbox{{\bf y}}}_{1,t} \,dL_N\\ & -\frac{\beta N}{2} \iint\frac{{\mbox{\boldmath$\xi$}}_t(x,M_N) - {\mbox{\boldmath$\xi$}}_t(y,M_N)}{x-y}\,dL_N(x)\,dL_N(y) +N\int V'_t(x)\,{\mbox{\boldmath$\xi$}}_t(x,M_N)\, dL_N(x)\\ & -\frac{1}{N} {\mbox{\boldmath$\eta$}}(M_N) - N\Bigl(1-\frac{\beta}2\Bigr) \int {\mbox{{\bf y}}}_{0,t}'\,dL_N\\ & - \Bigl(1-\frac{\beta}2\Bigr) \int {\mbox{{\bf y}}}_{1,t}'\,dL_N -\Bigl(1-\frac{\beta}2\Bigr) \int \partial_1{\mbox{\boldmath$\xi$}}_{t}(x,M_N)\,dL_N(x) -{\mbox{\boldmath$\eta$}}(\mu_{V_t}) +\tilde c^{N}_t, \end{align*} where $\tilde c^N_t$ is a constant and we use the convention that, when we integrate a function of the form $\frac{f(x)-f(y)}{x-y}$ with respect to $L_N\otimes L_N$, the diagonal terms give $f'(x)$. We now observe that $L_N$ converges towards $\mu_{V_t}$ as $N\to \infty$ \cite{Boutet}, see also \cite{BAG, AGZ} for the corresponding large deviation principle, and the latter minimizes $I_{V_t}$ (see \eqref{entr}). Hence, considering $\mu_\varepsilon:=(x+\varepsilon f)_\#\mu_{V_t}$ and writing that $I_{V_t}(\mu_{\varepsilon})\ge I_{V_t}(\mu_{V_t})$, by taking the derivative with respect to $\varepsilon$ at $\varepsilon =0$ we get \begin{equation}\label{bn} \int V'_t (x)f(x)\,d\mu_{V_t}(x)=\frac{\beta}{2}\iint \frac{f(x)-f(y)}{x-y}\, d\mu_{V_t}(x)\,d\mu_{V_t}(y)\end{equation} for all smooth bounded functions $f:\mathbb{R}\to\mathbb{R}$. Therefore we can recenter $L_N$ by $\mu_{V_t}$ in the formula above: more precisely, if we set \begin{equation}\label{defXi}\Xi f(x):=-\beta \int\frac{f(x)-f(y)}{x-y}\, d\mu_{V_t}(y)+V'_t(x)f (x),\end{equation} then \begin{multline*} N^2\int V_t'\, f \,dL_N- \frac{\beta N^2}{2}\iint \frac{f(x)-f(y)}{x-y}\,dL_N(x)\,dL_N(y)\\ =N\int \Xi f\, dM_N-\frac{\beta}{2}\iint \frac{f(x)-f(y)}{x-y}\,dM_N(x)\,dM_N(y) \end{multline*} Applying this identity to $f={\mbox{{\bf y}}}_{0,t},{\mbox{{\bf y}}}_{1,t},{\mbox{\boldmath$\xi$}}_t(\cdot,M_N)$ and recalling the definition of ${\mbox{\boldmath$\xi$}}_t(\cdot,M_N)$ (see \eqref{eq:YN}), we find \begin{align*} {\mathcal R}^N_t&= N \int [\Xi {\mbox{{\bf y}}}_{0,t}+W]\,dM_N\\ & +\int \left( \Xi{\mbox{{\bf y}}}_{1,t}+\Big(\frac{\beta}2 - 1 \Big)\bigg[{\mbox{{\bf y}}}_{0,t}' +\int \partial_{1}{\mbox{{\bf z}}}_{t}(z,\cdot)d\mu_{V_t}(z)\biggr]\right)\,dM_N\\ &+\iint dM_N(x)\,dM_N(y)\left(\Xi {\mbox{{\bf z}}}_{t}(\cdot,y)[x]-\frac{\beta}{2}\frac{{\mbox{{\bf y}}}_{0,t}(x)-{\mbox{{\bf y}}}_{0,t}(y)}{x-y}\right)+C^N_t +E_N, \end{align*} where $$ \Xi {\mbox{{\bf z}}}_t(\cdot, y) [x] =-\beta\int \frac{{\mbox{{\bf z}}}_t(x,y)-{\mbox{{\bf z}}}_t(\tilde x,y )}{x-\tilde x }\,d\mu_{V_t}(\tilde x)+V'_t(x){\mbox{{\bf z}}}_t(x,y),$$ $C^N_t$ is a deterministic term, and $E_N$ is a remainder that we will prove to be negligible: \begin{equation} \label{eq:EN} \begin{split} E_N&:= -\frac{1}{N}\int \partial_{2}{\mbox{{\bf z}}}_{t}(x,x)\,dM_N(x)-\frac{1}{N}\Bigl(1-\frac{\beta}2\Bigr) \int {\mbox{{\bf y}}}_{1,t}'\,dM_N\\ &-\frac{1}{N}\Bigl(1-\frac{\beta}2\Bigr)\iint \partial_1{\mbox{{\bf z}}}_t(x,y)\,dM_N(x)\,dM_N(y)\\ &-\frac{\beta}{2N}\iint \frac{{\mbox{{\bf y}}}_{1,t}(x)-{\mbox{{\bf y}}}_{1,t}(y)}{x-y}\, dM_N(x)\,dM_N(y)\\ &-\frac{\beta}{2N}\iiint\frac{{\mbox{{\bf z}}}_t(x,y)-{\mbox{{\bf z}}}_t(\tilde x,y)}{x-\tilde x}\,dM_N(x)\,dM_N(y)\,dM_N(\tilde x). \end{split} \end{equation} Hence, for ${\mathcal R}^N_t$ to be small we want to impose \begin{equation} \label{defpsi} \begin{split} \Xi{\mbox{{\bf y}}}_{0,t}&=-W+c, \\ \Xi {\mbox{{\bf z}}}_t(\cdot, y) [x]& =-\frac{\beta}{2} \frac{{\mbox{{\bf y}}}_{0,t}(x)-{\mbox{{\bf y}}}_{0,t}(y)}{x-y}+c(y),\\ \Xi{\mbox{{\bf y}}}_{1,t}&=-\Bigl(\frac{\beta}{2}-1\Bigr)\bigg[{\mbox{{\bf y}}}_{0,t}' + \int \partial_{1}{\mbox{{\bf z}}}_{t}(z,\cdot)\,d\mu_{V_t}(z)\biggr]+c', \end{split} \end{equation} where $c,c'$ are some constant to be fixed later, and $c(y)$ does not depend on $x$. \subsection{Inverting the operator $\Xi$.} We now prove a key lemma, that will allow us to find the desired functions ${\mbox{{\bf y}}}_{0,t},{\mbox{{\bf y}}}_{1,t},{\mbox{{\bf z}}}_{t}$. \begin{lem} \label{central} Given $V:\mathbb{R}\to \mathbb{R}$, assume that $\mu_{V}$ has support given by $[a,b]$ and that $$ \frac{d\mu_V}{dx}(x)=S(x)\sqrt{(x-a)(b-x)} $$ with $S(x)\geq \bar c>0$ a.e. on $[a,b]$. Let $g:\mathbb R{\rightarrow}\mathbb R$ be a $C^k$ function and assume that $V$ is of class $C^p$. Set $$\Xi f(x):=-\beta\int\frac{f(x)-f(y)}{x-y} d\mu_V(x) +V'(x)f(x)$$ Then there exists a unique constant $c_g$ such that the equation $$\Xi f(x)=g(x)+c_g$$ has a solution of class $C^{(k-2)\wedge (p-3)}$. More precisely, for $j\le (k-2)\wedge (p-3)$ there is a finite constant $C_j$ such that \begin{equation}\label{toto2b} \|f\|_{C^j(\mathbb{R})}\le C_j \|g\|_{C^{j+2}(\mathbb{R})} , \end{equation} where, for a function $h$, $\|h\|_{C^j(\mathbb{R})}:=\sum_{r=0}^j\|h^{(r)}\|_{L^\infty(\mathbb R)}$. Moreover $f$ (and its derivatives) behaves like $(g(x)+c_g)/V'(x)$ (and its corresponding derivatives) when $|x|\to +\infty$. This solution will be denoted by $\Xi^{-1}g$. \end{lem} Note that $Lf(x)=\Xi f'(x)$ can be seen as the asymptotics of the infinitesimal generator of the Dyson Brownian motion taken in the set where the spectral measure approximates $\mu_V$. This operator is central in our approach, as much as the Dyson Brownian motion is central to prove universality in e.g. \cite{ESYY,BEY1,BEY2}. \begin{proof} As a consequence of \eqref{bn}, we have \begin{equation} \label{eq:muV V'} \beta\,PV\int\frac{1}{x-y} \,d\mu_{V}(y)=V'(x)\qquad \text{on the support of $\mu_{V}$.} \end{equation} Therefore the equation $\Xi f(x)=g(x)+c_g$ on the support of $\mu_{V}$ amounts to \begin{equation} \label{eq:tri} \beta\,PV \int \frac{f(y)}{x-y} \,d\mu_{V}(y)= g(x)+c_g\qquad \forall\,x \in [a,b]. \end{equation} Let us write $$d(x):=d\mu_{V}/dx =S(x)\sqrt{(x-a)(b-x)}$$ with $S$ positive inside the support $[a,b]$. We claim that $S\in C^{p-3}([a,b])$. Indeed, by \eqref{bn} with $f(x)=(z-x)^{-1}$ for $z\in [a,b]^c$, we find that the Stieltjes transform $G(z)=\int (z-y)^{-1}\,d\mu_{V}(y)$ satisfies, for $z$ outside $[a,b]$, $$\frac{\beta}{2}G(z)^2=G(z)V'(\Re(z))+F(z),\qquad\mbox{ with } F(z)=\int \frac{V'(y)-V'(\Re(z))}{z-y}\,d\mu_{V}(y)\,.$$ Solving this quadratic equation so that $G\to 0$ as $|z|\to\infty$ yields \begin{equation} \label{eq:G} G(z)=\frac{1}{\beta} \Big(V'(\Re(z))-\sqrt{ [V'(\Re(z))]^2+2\beta F(z)}\Big). \end{equation} Notice that $V'(\Re(z))^2+2\beta F(z)$ becomes real as $z$ goes to the real axis, and negative inside $[a,b]$. Hence, since $-\pi^{-1} \Im G(z)$ converges to the density of $\mu_V$ as $z$ goes to the real axis (see e.g \cite[Theorem 2.4.3]{AGZ}), we get \begin{equation} \label{eq:S} -S(x)^2(x-a)(b-x)=(\beta\pi)^{-2}\left[V'(x)^2+2\beta F(x)\right]. \end{equation} This implies in particular that $\{a,b\}$ are the two points of the real line where $V'(x)^2+2\beta F(x)$ vanishes. Moreover $F(x)=-\int \int_0^1 V''(\alpha y+(1-\alpha)x) \,d\alpha\,d\mu_{V}(y)$ is of class $C^{p-2}$ on $\mathbb R$ (recall that $V\in C^p$ by assumption), therefore $(V')^2+2\beta F\in C^{p-2}(\mathbb{R})$. Since we assumed that $S$ does not vanish in $[a,b]$, from \eqref{eq:S} we deduce that $S$ is of class $C^{p-3}$ on $[a,b]$. To solve \eqref{eq:tri} we apply Tricomi's formula \cite[formula 12, p.181]{tricomi} and we find that, for $x\in [a,b]$, $$\beta f(x) {\sqrt{(x-a)(b-x)}} d(x)= PV \int_{a}^{b} \frac{{\sqrt {(y-a)(b-y)}}}{{y-x}} (g(y) +c_g)dy + c_2 := h(x) $$ for some constant $c_2$, hence $$ \begin{array}{rcl} h(x) &=& \beta f(x){{(x-a)(b-x)}} S(x)\\ &=& PV \int_a^b \frac{{\sqrt {(y-a)(b-y)}}}{{y-x}} (g(y) +c_g)dy + c_2\\ &=&PV \int_a^b {\sqrt {(y-a)(b-y)}} \frac{g(y) -g(x)}{y-x}dy + (g(x)+ c_g)PV \int_a^b \frac{{\sqrt {(y-a)(b-y)}}}{{y-x}} dy + c_2\\ &=& \int_a^b {\sqrt {(y-a)(b-y)}} \frac{g(y) -g(x)}{y-x}dy -\pi \left(x-\frac{a+b}{2} \right)(g(x)+ c_g) + c_2, \end{array} $$ where we used that, for $x\in [a,b]$, $$ PV \int_a^b \frac{{\sqrt {(y-a)(b-y)}}}{{y-x}}dy = -\pi \Bigl(x -\frac{ a+b}{2}\Bigr). $$ Set $$ h_0 (x)= \int_a^b {\sqrt {(y-a)(b-y)}}\, \frac{g(y) -g(x)}{y-x}dy.$$ Then $h_0$ is of class $C^{k-1}$ (recall that $g$ is of class $C^k$). We next choose $c_g$ and $c_2$ such that $h$ vanishes at $a$ and $b$ (notice that this choice uniquely identifies $c_g$). We note that $f \in C^{(k-2)\wedge (p-3)}([a,b])$. Moreover, we can bound its derivatives in terms of the derivatives of $h_0, g$ and $S$: if we assume $j\le p-3$, we find that there exists a constant $C_j$, which depends only on the derivatives of $S$, such that $$ \|f^{(j)}\|_{L^\infty([a,b])}\le C_j \max_{p\le j } \Bigl(\|h_0^{(p+1)}\|_{L^\infty([a,b])}+\|g^{(p+1)}\|_{L^\infty([a,b])}\Bigr)\le C_j \max_{p\le j+2} \|g^{(p)}\|_{L^\infty([a,b])}. $$ Let us define $$k(x):= \beta\,PV\int \frac{f(y)}{x-y} d\mu_{V}(y)-g(x)-c_g\qquad \forall\,x \in \mathbb{R}.$$ By \eqref{eq:tri} we see that $k\equiv 0$ on $[a,b]$. To ensure that $\Xi f=g+c_g$ also outside the support of $\mu_{V}$ we want $$f(x)\biggl(\beta\,PV\int\frac{1}{x-y}\,d\mu_{V}(y)-V'(x)\biggr)=k(x)\qquad\forall\,x \in [a,b]^c.$$ Let us consider the function $\ell:\mathbb{R}\to\mathbb{R}$ defined as \begin{equation} \label{eq:ell} \ell(x):= \beta\,PV\int\frac{1}{x-y}\,d\mu_V(y)-V'(x). \end{equation} Notice that, thanks to \eqref{eq:G}, $\ell(x)=\beta G(x) - V'(x)=-\beta\sqrt{ [V'(x)]^2+2\beta F(x)}$. Hence, comparing this expression with \eqref{eq:S} and recalling that $S \geq \bar c>0$ in $[a,b]$, we deduce that $[V'(x)]^2+2\beta F(x)$ is smooth and has simple zeroes both at $a$ and $b$, therefore $[V'(x)]^2+2\beta F(x)>0$ in $[a-{\epsilon},b+{\epsilon}]\backslash [a,b]$ for some ${\epsilon}>0$. This shows that $\ell$ does not vanish in $[a-{\epsilon},b+{\epsilon}]\backslash [a,b]$. Recalling that we can freely modify $V$ outside $[a-{\epsilon},b+{\epsilon}]$ (see proof of Lemma \ref{propa}), we can actually assume that $\ell$ vanishes at $\{a,b\}$ and does not vanish in the whole $[a,b]^c$. We claim that $\ell$ is H\"older $1/2$ at the boundary points, and in fact is equivalent to a square root there. Indeed, it is immediate to check that $\ell$ is of class $C^{p-1}$ except possibly at the boundary points $\{a,b\}$. Moreover \begin{align*} PV\int\frac{1}{x-y}\,d\mu_{V}(y)&= S(a) \int_a^b\frac{1}{x-y} \sqrt{(y-a)(b-y)} \,dy\\ &+\int_a^b \frac{y-a}{x-y}\biggl(\int_0^1 S'(\alpha a+(1-\alpha) y) d\alpha\biggr) \sqrt{(y-a)(b-y)} \,dy. \end{align*} The first term can be computed exactly and we have, for some $c\neq 0$, \begin{equation}\label{ipo} \int_a^b\frac{1}{x-y} \sqrt{(y-a)(b-y)} \,dy =c (b-a)\biggl( \frac{x-\frac{a+b}{2}}{b-a} -\sqrt{ \Bigl(\frac{x-\frac{a+b}{2}}{b-a}\Bigr)^2-\frac{1}{4}}\biggr)\end{equation} which is H\"older $1/2$, and in fact behaves as a square root at the boundary points. On the other hand, since $S$ is of class $C^{p-3}$ on $[a,b]$ with $p\ge 4$, the second function is differentiable, with derivative at $a$ given by $$\int_a^b \frac{1}{a-y}\biggl(\int_0^1 S'(\alpha a+(1-\alpha) y) d\alpha \biggr)\,\sqrt{(y-a)(b-y)} \,dy,$$ which is a convergent integral. The claim follows. Thus, for $x$ outside the support of $\mu_{V}$ we can set $$f(x):=\ell(x)^{-1}k(x).$$ With this choice $\Xi f=g + c_g$ and $f$ is of class $C^{(k-2)\wedge (p-3)}$ on $\mathbb{R}\setminus\lbrace a,b \rbrace$. We now want to show that $f$ is of class $C^{(k-2)\wedge (p-3)}$ on the whole $\mathbb{R}$. For this we need to check the continuity of $f$ and its derivatives at the boundary points, say at $a$ (the case of $b$ being similar). We take hereafter $r\le ( k-2)\wedge (p-3)$, so that $f$ has $r$ derivatives inside $[a,b]$ according to the above considerations. \\ Let us first deduce the continuity of $f$ at $a$. We write, with $f(a^+)=\lim_{x\downarrow a} f(x)$, $$k(x)=f(a^+) \ell(x)+ k_1(x)$$ with $$ k_1(x):=\beta\left( PV\int \frac{f(y)}{x-y} d\mu_{V}(y)- PV\int \frac{f(a^+)}{x-y} d\mu_V(y) \right) +g(x)+c_g+ f(a^+)V'_t(x). $$ Notice that since $f=\ell^{-1}k$ outside $[a,b]$, if we can show that $\ell^{-1}(x)k_1(x)\to 0$ as $x\uparrow a$ then we would get $f(a^-)=f(a^+)$, proving the desired continuity. To prove it we first notice that $k_1$ vanishes at $a$ (since both $k$ and $\ell$ vanish inside $[a,b]$), hence \begin{align*} k_1(x)&= \beta\left( PV\int \frac{f(y)-f(a^+) }{x-y} d\mu_{V}(y)- PV\int \frac{f(y)-f(a^+)}{a-y} d\mu_{V}(y) \right) +\tilde g(x)-\tilde g(a)\\ &=\beta (a-x)\,PV\int \frac{f(y)-f(a^+)}{(x-y)(a-y)} d\mu_{V}(y) +\tilde g(x)-\tilde g(a), \end{align*} with $\tilde g:=g+f(a^+)V' \in C^1$. Assume $1\le ( k-2)\wedge (p-3)$. Since $f$ is of class $C^1$ inside $[a,b]$ we have $|f(y)-f(a^+)|\leq C|y-a|$, from which we deduce that $|k_1(x)|\leq C|x-a|$ for $x \leq a$. Hence $\ell^{-1}(x)k_1(x)\to 0$ as $x\uparrow a$ (recall that $\ell$ behaves as a square root near $a$), which proves that $$\lim_{x\uparrow a} f(x)=\lim_{x\downarrow a}f(x)$$ and shows the continuity of $f$ at $a$. We now consider the next derivative: we write $$k(x)=\bigl[f(a)+f'(a^+) (x-a)\bigr] \ell(x)+ k_2(x)$$ with \begin{align*} k_2(x)&:= \beta (a-x)\,PV\int \frac{f(y)-f(a^+)-(y-a) f'(a^+)}{(x-y)(a-y)}\, d\mu_{V}(y)\\ & +\tilde g(x)-\tilde g(a)+f'(a^+) (x-a) V'_t(x). \end{align*} Since $k=\ell\equiv 0$ on $[a,b]$ we have $k_2(a)=k_2'(a^+)=k_2'(a^-)=0$. Hence, since $f$ is of class $C^2$ on $[a,b]$, we see that $|k_2(x)|\leq C|x-a|^2$ for $x \leq a$, therefore $k_2(x)/\ell(x)$ is of order $|x-a|^{3/2}$, thus $$ f(x)=f(a)+f'(a^+) (x-a)+ O(|x-a|^{3/2})\qquad \text{for $x \leq a$,} $$ which shows that $f$ has also a continuous derivative. We obtain the continuity of the next derivatives similarly. Moreover, away from the boundary point the $j$-th derivative of $f$ outside $[a,b]$ is of the same order than that of $g/ V'$, while near the boundary points it is governed by the derivatives of $g$ nearby, therefore \begin{equation}\label{toto2} \|f^{(j)}\|_{{L^\infty([a,b]^c)}}\le C'_j\max_{r\le j+2} \|g^{(r)}\|_{L^\infty(\mathbb R)} \,. \end{equation} Finally, it is clear that $f$ behaves like $(g(x)+c_g)/V'(x)$ when $x$ goes to infinity. \end{proof} \def{\mathcal L}{{\mathcal L}} \subsection{Defining the functions $ {\mbox{{\bf y}}}_{0,t},{\mbox{{\bf y}}}_{1,t},{\mbox{{\bf z}}}_{t}$} To define the functions ${\mbox{{\bf y}}}_{0,t},{\mbox{{\bf y}}}_{1,t},{\mbox{{\bf z}}}_{t}$ according to \eqref{defpsi}, notice that Lemma \ref{propa} shows that the hypothesis of Lemma \ref{central} are fulfilled. Hence, as a consequence of Lemma \ref{central} we find the following result (recall that $\psi\in C^{s,v}$ means that $\psi$ is $s$ times continuously differentiable with respect to the first variable and $v$ times with respect to the second). \begin{lem}\label{lem:regular} Let $r\ge 7$. If $W, V' \in C^r$, we can choose ${\mbox{{\bf y}}}_{0,t}$ of class $C^{r-2}$, ${\mbox{{\bf z}}}_t\in C^{s,v}$ for $s+v\le r-5,$ and ${\mbox{{\bf y}}}_{1,t}\in C^{r-8}$. Moreover, these functions (and their derivatives) go to zero at infinity like $1/V'$ (and its corresponding derivatives). \end{lem} \begin{proof} By Lemma \ref{central} we have ${\mbox{{\bf y}}}_{0,t}=\Xi^{-1} W\in C^{r-2}$. For ${\mbox{{\bf z}}}_{t}$, we can rewrite \begin{align*} \Xi {\mbox{{\bf z}}}_t(\cdot, y) [x] &=-\frac{\beta}{2}\int_0^1 {\mbox{{\bf y}}}_{0,t}'(\alpha x +(1-\alpha)y )\,d\alpha +c(y)\\ &= -\frac{\beta}{2}\int_0^1 [{\mbox{{\bf y}}}_{0,t}'(\alpha x +(1-\alpha)y )+c_\alpha(y)]\,d\alpha \end{align*} where we choose $c_\alpha(y)$ to be the unique constant provided by Lemma \ref{central} which ensures that $\Xi^{-1} [{\mbox{{\bf y}}}_{0,t}'(\alpha x +(1-\alpha)y )+c_\alpha(y)]$ is smooth. This gives that $c(y)=\int_0^1 c_\alpha(y) d\alpha$. Since $\Xi^{-1}$ is a linear integral operator, we have $${\mbox{{\bf z}}}_t(x,y)= -\frac{\beta}{2}\int_0^1 \Xi^{-1}[ {\mbox{{\bf y}}}_{0,t}'(\alpha \cdot +(1-\alpha)y )](x)\, d\alpha\,.$$ As the variable $y$ is only a translation, it is not difficult to check that ${\mbox{{\bf z}}}_t\in C^{s,v}$ for any $s+v\le r-5$. It follows that $$-\Bigl(\frac{\beta}{2}-1\Bigr)\bigg[{\mbox{{\bf y}}}_{0,t}' + \int \partial_{1}{\mbox{{\bf z}}}_{t}(z,\cdot )\,d\mu_{V_t}(z)\biggr]+c'$$ is of class $C^{r-6}$ and therefore by Lemma \ref{central} we can choose ${\mbox{{\bf y}}}_{1,t} \in C^{r-8}$, as desired. The decay at infinity is finally again a consequence of Lemma \ref{central}. \end{proof} \subsection{Getting rid of the random error term $E_N$} \label{sect:error} We show that the $L^1_{\mathbb P_{V_t}^N}$-norm of the error term $E_N$ defined in \eqref{eq:EN} goes to zero. This could be easily derived from \cite{BEY1}, but we here provide a self-contained proof. To this end, we first make some general consideration on the growth of variances. Following e.g. \cite[Theorem 1.6]{Edouard-mylene}, up to assume that $V_t$ goes sufficiently fast at infinity (which we did, see Lemma \ref{propa}), it is easy to show that there exists a constant $\tau_0>0$ so that for all $\tau \ge \tau_0$, $$ \mathbb P^N_{V_t}\biggl( D( L_N,\mu_{V_t})\ge \tau\sqrt{\frac{\log N}{N}} \biggr)\le e^{-c \tau^2 N\log N }\,, $$ where $D$ is the $1$-Wasserstein distance $$ D(\mu,\nu):=\sup_{\|f'\|_{\infty}\le 1} \biggl|\int f(d\mu-d\nu)\biggr|. $$ Since $M_N=N(L_N-\mu_{V_t})$ we get \begin{equation} \label{eq:W M} D( L_N,\mu_{V_t})=\frac{1}{N} \sup_{\|f'\|_{\infty}\le 1} \biggl|\int f\,dM_N\biggr|, \end{equation} hence for $\tau\ge \tau_0$ \begin{equation}\label{conc} \mathbb P^N_{V_t}\biggl( \sup_{\|f\|_{\rm Lip}\le 1} \biggl|\int f\,dM_N\biggr| \ge \tau\sqrt{N \log N} \biggr)\le e^{-c \tau^2 N\log N }.\end{equation} This already shows that, if $f$ is sufficiently smooth, $\int f(x,y) dM_N(x)\,dM_N(y)$ is of order at most $N\log N$. More precisely $$\int f(x,y)\, dM_N(x)\,dM_N(y)=\int \hat f(\zeta,\xi)\biggl( \int e^{i\zeta x} dM_N(x) \int e^{i\xi x} dM_N(x)\biggr)\, d\xi\, d\zeta,$$ so that with probability greater than $1-e^{-c\tau_0^2N \log N}$ we have \begin{equation}\label{bor1} \left|\int f(x,y) \,dM_N(x)\,dM_N(y)\right|\le \tau_0^2\, N\log N \,\int | \hat f(\zeta,\xi)|\, |\zeta|\,|\xi|\,d\zeta\, d\xi\,.\end{equation} To improve this estimate, we shall use loop equations as well as Lemma \ref{central}. Given a function $g$ and a measure $\nu$, we use the notation $\nu(g):=\int g\,d\nu$. \begin{lem}\label{conch} Let $g$ be a smooth function. Then, if $\tilde M_N=NL_N-N\mathbb E_{V_t}[ L_N]$, there exists a finite constant $C$ such that \begin{align*} \sigma^{(1)}_N(g)&:=\biggl| \int M_N(g) \,d \mathbb P^N_{V_t}\biggr|\le C \,m(g) =:B^1_N(g)\\ \sigma^{(2)}_N(g)&:= \int \left(\tilde M_N(g)\right)^2 d \mathbb P^N_{V_t}\le C\Bigl(m(g)^2 + m(g)\|g\|_\infty +\|\Xi^{-1} g\|_\infty \|g'\|_{\infty}\Bigr)=:B^2_N(g) \\ \sigma^{(4)}_N(g)&:=\int \left(\tilde M_N(g)\right)^4 \, d \mathbb P^N_{V_t}\\ &\le C \Bigl( \|\Xi^{-1} g\|_\infty \|g'\|_{\infty} \sigma_N^{(2)}(g)+ \|g\|_\infty^3m(g) +m(g)^2 \sigma_N^{(2)}(g)+m(g)^4\Bigr) =:B^4_N(g), \end{align*} where \begin{align*} m(g)&:=\Bigl|1-\frac{\beta}{2}\Bigr|\|(\Xi^{-1} g)'\|_\infty + \frac{\beta}{2} \log N \int |\hat \Xi^{-1} g|(\xi) \,|\xi|^3 \,d\xi. \end{align*} \end{lem} \begin{proof} First observe that, by integration by parts, for any $C^1$ function $f$ \begin{equation}\label{loop1} \int \biggl(N\sum_i V'(\lambda_i)f(\lambda_i)- \beta \sum_{i<j}\frac{f(\lambda_i)-f(\lambda_j)}{\lambda_i-\lambda_j} \biggr)\,d\mathbb P^N_{V_t}=\int \sum_{i} f'(\lambda_i) \,d\mathbb P^N_{V_t}\end{equation} which we can rewrite as the first loop equation \begin{equation} \label{lkjh} \int M_N(\Xi f)\,d \mathbb P^N_{V_t}=\int \biggl[ \Bigl(1-\frac{\beta}{2}\Bigr)\int f' dL_N + \frac{\beta}{2 N} \int \frac{f(x)-f(y)}{x-y} d M_N(x) d M_N(y)\biggr]\,d\mathbb P^N_{V_t}\,.\end{equation} We denote $$F_N(g):=\Bigl(1-\frac{\beta}{2}\Bigr)\int (\Xi^{-1} g)' \,dL_N + \frac{\beta}{2 N} \int \frac{\Xi^{-1} g(x)-\Xi^{-1} g(y)}{x-y} \,d M_N(x)\, d M_N(y)$$ so that taking $f:=\Xi^{-1} g$ in \eqref{lkjh} we deduce $$\int M_N(g)\,d \mathbb P^N_{V_t}=\int F_N(g)\, d \mathbb P^N_{V_t}.$$ To bound the right hand side above, we notice that $\Xi^{-1}g$ goes to zero at infinity like $1/V'$ (see Lemma \ref{central}). Hence we can write its Fourier transform and get \begin{multline*} \int \frac{\Xi^{-1} g(x)-\Xi^{-1} g(y)}{x-y} \,d M_N(x)\, d M_N(y)\\ = i\int d\xi\, \xi \,\hat \Xi^{-1}g(\xi)\int_0^1 d\alpha \int e^{i\alpha\xi x}\,dM_N(x) \int e^{i(1-\alpha)\xi y}\,dM_N(y), \end{multline*} so that we deduce (recall \eqref{eq:W M}) $$\sup_{D(L_N,\mu_{V_t})\le \tau_0 \sqrt{\log N/N}} F_N(g)\le (1+ \tau_0^2)\, m(g).$$ On the other hand, as the mass of $M_N$ is always bounded by $2N$, we deduce that $F_N(g)$ is bounded everywhere by $ N m(g) $. Since the set $\{D( L_N,\mu_{V_t})\ge \tau_0 \sqrt{N\log N}\}$ has small probability (see \eqref{conc}), we conclude that \begin{equation}\label{bor2} \biggl|\int M_N(g) \,d \mathbb P^N_{V_t}\biggr|\le N e^{-c\tau_0^2 N\log N }m(g) + (1+\tau_0^2) m(g) \leq C\,m(g),\end{equation} which proves our first bound. Before proving the next estimates, let us make a simple remark: using the definition of $M_N$ and $\tilde M_N$ it is easy to check that, for any function $g$, \begin{equation} \label{eq:difference MN tilde} \left|M_N(g) - \tilde M_N(g)\right| =\biggl|\int M_N(g) \,d \mathbb P^N_{V_t}\biggr|. \end{equation} To get estimates on the covariance we obtain the second loop equation by changing $V(x)$ into $V(x)+\delta \,g(x)$ in \eqref{loop1} and differentiating with respect to $\delta$ at $\delta=0$. This gives \begin{equation} \label{lkjp} \begin{split} &\int M_N(\Xi f) \tilde M_N(g) \,d \mathbb P^N_{V_t}=\int L_N(f g')\,d \mathbb P^N_{V_t} \\ &+\int \biggl[ \Bigl(1-\frac{\beta}{2}\Bigr)\int f' dL_N + \frac{\beta}{2 N} \int \frac{f(x)-f(y)}{x-y} d M_N(x) d M_N(y)\biggr]\tilde M_N(g)\, d\mathbb P^N_{V_t}. \end{split} \end{equation} We now notice that $M_N(\Xi f)-\tilde M_N(\Xi f)$ is deterministic and $\int \tilde M_N(g) \,d \mathbb P^N_{V_t}=0$, hence the left hand side is equal to $$ \int \tilde M_N(\Xi f) \tilde M_N(g) \,d \mathbb P^N_{V_t}. $$ We take $f:=\Xi^{-1}g$ and we argue similarly to above (that is, splitting the estimate depending whether $D( L_N,\mu_{V_t}) \geq \tau_0\sqrt{N\log N}$ or not, and use that $|\tilde M_N(g)|\leq N\|g\|_\infty$) to deduce that $\sigma_N^{(2)}(g):=\int |\tilde M_N( f) |^2 d \mathbb P^N_{V_t}$ satisfies \begin{equation} \label{bo1} \begin{split} \sigma_N^{(2)}(g)&\le \|g'\Xi^{-1} g\|_\infty +\int |F_N(g)| |\tilde M_N(g)| d\mathbb P^N_{V_t}\\ &\le \|\Xi^{-1} g\|_\infty\,\|g'\|_{\infty} + N^2 e^{-c\tau_0^2 N\log N}\|g\|_\infty m(g)+ C\,m(g)\int |\tilde M_N(g)| d\mathbb P^N_{V_t}\\ & = \|\Xi^{-1} g\|_\infty \|g'\|_{\infty} +N^2 e^{-c\tau_0^2 N\log N }m(g)\|g\|_\infty + C\,m(g)\sigma_N^{(2)}(g)^{1/2}. \end{split} \end{equation} Solving this quadratic inequality yields $$\sigma_N^{(2)}(g)\le C \Bigl[ m(g)^2 + m(g)\|g\|_\infty +\|\Xi^{-1} g\|_\infty \|g'\|_{\infty} \Bigr] $$ for some finite constant $C$. We finally turn to the fourth moment. If we make an infinitesimal change of potential $V(x)$ into $V(x)+\delta_1\, g_2(x) +\delta_2\, g_3(x)$ and differentiate at $\delta_1=\delta_2=0$ into \eqref{lkjp} we get, denoting $g=g_1$, \begin{equation} \label{lkj2} \begin{split} &\int M_N(\Xi f) \tilde M_N(g_1) \tilde M_N(g_2) \tilde M_N(g_3)\,d \mathbb P^N_{V_t}=\int \biggl[\sum_\sigma L_N(f g_{\sigma(1)}')\tilde M_N(g_{\sigma(2)}) \tilde M_N(g_{\sigma(3)}) \biggr]\,d\mathbb P^N_{V_t}+\\ &\int \biggl[ \Bigl(1-\frac{\beta}{2}\Bigr)\int f' \,dL_N + \frac{\beta}{2 N} \int \frac{f(x)-f(y)}{x-y}\, d M_N(x)\, d M_N(y)\biggr] M_N(g_1)\tilde M_N(g_2)\tilde M_N(g_3) \,d\mathbb P^N_{V_t} , \end{split} \end{equation} where we sum over the permutation $\sigma$ of $\{1,2,3\}$. Taking $\Xi f=g_1=g_2=g_3=g$, by \eqref{eq:difference MN tilde}, \eqref{bor2}, and Cauchy-Schwarz inequality we get $$\sigma_{N}^{(4)}(g)\le C\Bigl[ \|g'\Xi^{-1} g\|_\infty \sigma_N^{(2)}(g) + \|g\|_\infty^3m(g)+m(g) \sigma_N^{(4)}(g)^{3/4}+ m(g)^2\sigma_N^{(2)}(g)\Bigr], $$ which implies $$\sigma_{N}^{(4)}(g)\le C\Bigl[ \|g'\Xi^{-1} g \|_\infty \sigma_N^{(2)}(g)+ \|g\|_\infty^3m(g) +m(g)^2 \sigma_N^{(2)}(g)+m(g)^4\Bigr].$$ \end{proof} Applying the above result with $g=e^{i\lambda \cdot}$ we get the following: \begin{cor} Assume that $V',W \in C^r$ with $r\ge 8$. Then there exists a finite constant $C$ such that, for all $\lambda\in \mathbb R$, \begin{align} \int |M_N(e^{i\lambda\cdot})|^2 d \mathbb P^N_{V_t}&\le C[\log N(1+|\lambda|^{7})]^2\,,\label{conc3b}\\ \int |M_N(e^{i\lambda\cdot})|^4 d \mathbb P^N_{V_t}&\le C[\log N(1+|\lambda|^{7})]^4\,.\label{conc3bb} \end{align} \end{cor} \begin{proof} In the case $g(x)=e^{i\lambda x}$ we estimate the norms of $\Xi^{-1}g$ by using Lemma \ref{central}, and we get a finite constant $C$ such that $$\|\Xi^{-1} g\|_\infty\le C |\lambda|^2,\qquad \|\Xi^{-1}g'\|_\infty\le C |\lambda|^3,$$ whereas, since $\Xi^{-1}g$ goes fast to zero at infinity (as $1/V'$), for $j \leq r-3$ we have (see Lemma \ref{central}) $$|\hat\Xi^{-1} g|(\xi)\le C\frac{\| \Xi^{-1} g\|_{C^j}}{1+|\xi|^j}\le C'\frac{\| g\|_{C^{j+2}}}{1+|\xi|^j}\le C'\frac{1+ |\lambda|^{j+2}}{1+|\xi|^j}.$$ Hence, we deduce that there exists a finite constant $C'$ such that \begin{align*} m(g)&\le C\log N \biggl( |\lambda|^3+1+ \int d\xi \,\frac{1+|\lambda|^{7}}{1+ |\xi|^{5}}|\xi|^3 \biggr)=C'\log N\left(1+|\lambda|^7\right),\\ B_N^1(g)&\le C'\log N \left(1+|\lambda|^7\right),\\ B_N^2(g)&\le C' (\log N)^2 \left(1+|\lambda|^7\right)^2,\\ B_N^4(g)&\le C'(\log N)^4\left(1+|\lambda|^7\right)^4.\\ \end{align*} Finally, for $k=2,4$, using \eqref{eq:difference MN tilde} and \eqref{bor2} we have $$\int |M_N(e^{i\lambda\cdot})|^k \, d \mathbb P^N_{V_t}\le 2^{k-1}\left( \int |\tilde M_N(e^{i\lambda\cdot})|^k \,d \mathbb P^N_{V_t} +\bigl(B_N^1(g)\bigr)^k\right)$$ from which the result follows. \end{proof} We can now estimate $E_N$. The linear term can be handled in the same way as we shall do now for the quadratic and cubic terms (which are actually more delicate), so we just focus on them. We have two quadratic terms in $M_N$ which sum up into $$E^1_N=-\frac{1}{N}\Bigl(1-\frac\beta2\Bigr)\iint \partial_1{\mbox{{\bf z}}}_t(x,y)\,dM_N(x)\,dM_N(y) -\frac{\beta}{2N}\iint \frac{{\mbox{{\bf y}}}_{1,t}(x)-{\mbox{{\bf y}}}_{1,t}(y)}{x-y} \,dM_N(x)\,dM_N(y). $$ Writing $$ \frac{{\mbox{{\bf y}}}_{1,t}(x)-{\mbox{{\bf y}}}_{1,t}(y)}{x-y}=\int_0^1 {\mbox{{\bf y}}}_{1,t}'(\alpha x+(1-\alpha)y)\,d\alpha =\int_0^1\biggl( \int \widehat{{\mbox{{\bf y}}}_{1,t}'}(\xi) e^{i(\alpha x+(1-\alpha)y)\xi}d\xi\biggr)\,d\alpha $$ we see that $$ \iint \frac{{\mbox{{\bf y}}}_{1,t}(x)-{\mbox{{\bf y}}}_{1,t}(y)}{x-y} \,dM_N(x)\,dM_N(y)=\int d\xi\, \widehat{{\mbox{{\bf y}}}_{1,t}'}(\xi)\int_0^1d\alpha \,M_N(e^{i\alpha\xi \cdot})M_N(e^{i(1-\alpha)\xi \cdot}), $$ so using \eqref{conc3b} we get \begin{multline*} \int |E^1_N| \,d \mathbb P^N_{V_t}\le C\,\frac{(\log N)^2 }{N}\bigg(\int d\xi\, |\hat {\mbox{{\bf y}}}_{1,t}|(\xi)\, |\xi|\, \bigl(1+|\xi|^7\bigr)^2\\ + \iint d\xi \,d\zeta\, |\hat{\mbox{{\bf z}}}_{t}| (\xi,\zeta) \,|\xi|\left(1+|\xi|^7\right) \left(1+|\zeta|^7\right) \bigg). \end{multline*} It is easy to see that the right hand side is finite if ${\mbox{{\bf y}}}_{1,t}$ and ${\mbox{{\bf z}}}_{t}$ are smooth enough (recall that these functions and their derivatives decay fast at infinity). More precisely, to ensure that $$ |\hat {\mbox{{\bf y}}}_{1,t}|(\xi)\, |\xi|\, \bigl(1+|\xi|^7\bigr)^2 \leq \frac{C}{1+|\xi|^2} \in L^1(\mathbb{R}) $$ and $$ |\hat{\mbox{{\bf z}}}_{t}| (\xi,\zeta) \,|\xi|\left(1+|\xi|^7\right) \left(1+|\zeta|^7\right) \leq \frac{C}{1+|\xi|^3+|\zeta|^3}\in L^1(\mathbb{R}^2), $$ we need ${\mbox{{\bf y}}}_{1,t}\in C^{17}$ and ${\mbox{{\bf z}}}_{t}\in C^{11, 7}\cap C^{8,10}$, so (recalling Lemma \ref{lem:regular}) $V',W\in C^{25}$ is enough to guarantee that the right hand side is finite. Using \eqref{conc3b}, \eqref{conc3bb}, and H\"older inequality, we can similarly bound the expectation of the cubic term \begin{align*} E^2_N&=\frac{\beta}{2N}\iiint\frac{{\mbox{{\bf z}}}_t(x,y)-{\mbox{{\bf z}}}_t(\tilde x,y)}{x-\tilde x}\,dM_N(x)\,dM_N(y)\,dM_N(\tilde x)\\ &=i\frac{\beta}{2N}\iint d\xi \,d\zeta \,\widehat{\partial_1{\mbox{{\bf z}}}_{t}}(\xi,\zeta)\int_0^1 d\alpha \, M_N(e^{i\alpha\xi\cdot})M_N(e^{i(1-\alpha)\xi\cdot})M_N(e^{i\zeta\cdot}) \end{align*} to get $$ \int |E^2_N| \,d \mathbb P^N_{V_t}\le C\,\frac{(\log N)^3 }{{N}}\iint d\xi\, d\zeta\, |\hat{\mbox{{\bf z}}}_{t}(\xi,\zeta)| \,|\xi|\left(1+|\xi|^7\right)^2 \left(1+|\zeta|^7\right).$$ Again the right hand side is finite if ${\mbox{{\bf z}}}_{t} \in C^{18, 7}\cap C^{15,10}$, which is ensured by Lemma \ref{lem:regular} if $V',W$ are of class $C^{30}$. \subsection{Control on the deterministic term $C^N_t$} By what we proved above we have $$ \int| {\mathcal R}^N_t - C^N_t|\,d{\mathbb P^N_{V_t}}\leq C\,\frac{(\log N)^3 }{{N}}, $$ thus, in particular, $$ \bigl|C^N_t- \mathbb E[{\mathcal R}^N_t]\bigr|\leq C\,\frac{(\log N)^3 }{{N}}. $$ Notice now that, by construction, $${\mathcal R}^N_t =-{\mathcal L} {\mbox{{\bf Y}}}^N_t +N\sum_i W(\lambda_i)+c^N_t$$ with $c^N_t=-\mathbb E[ N\sum_i W(\lambda_i)]$ and $${\mathcal L} {\mbox{{\bf Y}}}:={\rm div}{\mbox{{\bf Y}}}+\beta\sum_{i<j}\frac{{\mbox{{\bf Y}}}^i-{\mbox{{\bf Y}}}^j}{\lambda_i-\lambda_j}-N\sum_i V'(\lambda_i) {\mbox{{\bf Y}}}^i,$$ and an integration by parts shows that, under $P_{V}^N$, $\mathbb E[{\mathcal L} {\mbox{{\bf Y}}}]=0$ for any vector field ${\mbox{{\bf Y}}}$. This implies that $\mathbb E[{\mathcal R}^N_t]=0$, therefore $|C^N_t|\leq C\,\frac{(\log N)^3 }{{N}}$. This concludes the proof of Proposition \ref{prop:key}. \section{Reconstructing the transport map via the flow} \label{sect:flow} In this section we study the properties of the flow generated by the vector field ${\mbox{{\bf Y}}}_t^N$ defined in \eqref{eq:YN}. As we shall see, we will need to assume that $W,V' \in C^r$ with $r \geq 15$. We consider the flow of ${\mbox{{\bf Y}}}^N_t$ given by $$ X_t^N:\mathbb R^N\to \mathbb R^N,\qquad \dot X_t^N={\mbox{{\bf Y}}}_t^N(X_t^N). $$ Recalling the form of ${\mbox{{\bf Y}}}_t^N$ (see \eqref{eq:YN}) it is natural to expect that we can give an expansion for $X_t^N$. More precisely, let us define the flow of ${\mbox{{\bf y}}}_{0,t}$, \begin{equation} \label{eq:X0} X_{0,t}:\mathbb R\to \mathbb R,\qquad \dot X_{0,t}={\mbox{{\bf y}}}_{0,t}(X_{0,t}),\quad X_{0,0}(\lambda)=\lambda, \end{equation} and let $X_{1,t}^N=(X_{1,t}^{N,1},\ldots,X_{1,t}^{N,N}):\mathbb R^N\to \mathbb R^N$ be the solution of the linear ODE \begin{equation} \label{eq:X1} \begin{split} \dot X_{1,t}^{N,k}(\lambda_1,\ldots,\lambda_N)&= {\mbox{{\bf y}}}_{0,t}'(X_{0,t}(\lambda_k))\cdot X_{1,t}^{N,k}(\lambda_1,\ldots,\lambda_N) + {\mbox{{\bf y}}}_{1,t}(X_{0,t}(\lambda_k))\\ &+\int {\mbox{{\bf z}}}_t(X_{0,t}(\lambda_k), y)\,dM_N^{X_{0,t}}(y)\\ &+\frac1N\sum_{j=1}^N \partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),X_{0,t}(\lambda_j)\Bigr)\cdot X_{1,t}^{N,j}(\lambda_1,\ldots,\lambda_N) \end{split} \end{equation} with the initial condition $X_{1,t}^N=0$, and $M_N^{X_{0,t}}$ is defined as $$ \int f(y) dM_N^{X_{0,t}}(y)=\sum_{i=1}^N \biggl[f(X_{0,t}(\lambda_i))-\int f d\mu_{V_t}\biggr]\qquad \forall\,f \in C_c(\mathbb R). $$ If we set $$ X_{0,t}^N(\lambda_1,\ldots,\lambda_N):=\bigl(X_{0,t}(\lambda_1),\ldots,X_{0,t}(\lambda_N)\bigr), $$ then the following result holds. \begin{lem} \label{the flow} Assume that $W,V' \in C^r$ with $r \geq 15$. Then the flow $X^N_t=(X_{t}^{N,1},\ldots,X_{t}^{N,N}):\mathbb R^N\to \mathbb R^N$ is of class $C^{r-8}$ and the following properties hold: Let $X_{0,t}$ and $X_{1,t}^N$ be as in \eqref{eq:X0} and \eqref{eq:X1} above, and define $X^N_{2,t}:\mathbb{R}^N\to\mathbb{R}^N$ via the identity $$X_t^N=X_{0,t}^{N}+ \frac1N X_{1,t}^N+\frac{1}{N^2} X_{2,t}^N\,.$$ Then \begin{equation}\label{boi1} \sup_{1\leq k \leq N}\|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)} \leq C\log N,\qquad \|X_{2,t}^{N}\|_{L^2(\mathbb P_{V}^N)} \leq CN^{1/2}(\log N)^2, \end{equation} where $$ \|X_{i,t}^{N}\|_{L^2(\mathbb P_{V}^N)}=\biggl(\int |X_{i,t}^N|^2d\mathbb P_{V}^N\biggr)^{1/2} ,\qquad |X_{i,t}^N|:=\sqrt{\sum_{j=1,\ldots,N}|X_{i,t}^{N,j}|^2},\quad i =0,1,2. $$ In addition, there exists a constant $C>0$ such that, with probability greater than $1- N^{-N/C}$, \begin{equation}\label{boi2}\max_{1\le k,k'\le N} |X_{1,t}^{N,k}(\lambda_1,\ldots,\lambda_N)-X_{1,t}^{N,k'}(\lambda_1,\ldots,\lambda_N)|\le C\,\log N \sqrt{N}|\lambda_k-\lambda_{k'}|.\end{equation} \end{lem} \begin{proof} Since ${\mbox{{\bf Y}}}_t^N \in C^{r-8}$ (see Lemma \ref{lem:regular}) it follows by Cauchy-Lipschitz theory that $X_t^N$ is of class $C^{r-8}$. Using the notation $\hat\lambda=(\lambda_1,\ldots,\lambda_N)\in \mathbb R^N$ and $$ X_t^{N,k,\sigma}(\hat\lambda):= X_{0,t}(\lambda_k)+ \sigma\frac{X_{1,t}^{N,k}}{N}(\hat\lambda)+\sigma\frac{X_{2,t}^{N,k}}{N^2}(\hat\lambda) =(1-\sigma)X_{0,t}(\lambda_k)+\sigma X_t^{N,k}(\hat\lambda) $$ and defining the measure $M_N^{X_{t}^{N,s}}$ as \begin{equation} \label{eq:MN1} \int f(y)\, dM_N^{X_{t}^{N,s}}(y)=\sum_{i=1}^N \biggl[f\bigl((1-s)X_{0,t}(\lambda_i)+sX_t^{N,i}(\hat\lambda)\bigr)-\int f \,d\mu_{V_t}\biggr]\qquad \forall\,f \in C_c(\mathbb R). \end{equation} by a Taylor expansion we get an ODE for $X_{2,t}^N$: \begin{equation} \label{eq:ODE X2 main} \begin{split} \dot X_{2,t}^{N,k}(\hat \lambda)&= \int_0^1 {\mbox{{\bf y}}}_{0,t}'\Bigl(X_t^{N,k,s}(\hat\lambda) \Bigr)\,ds\cdot X_{2,t}^{N,k}(\hat\lambda)\\ &+N\int_0^1\Bigl[{\mbox{{\bf y}}}_{0,t}'\Bigl(X_t^{N,k,s}(\hat\lambda)\Bigr) -{\mbox{{\bf y}}}_{0,t}'\Bigl(X_{0,t}(\lambda_k) \Bigr)\Bigr]\,ds \cdot X_{1,t}^{N,k}(\hat\lambda)\\ &+ \int_0^1 {\mbox{{\bf y}}}_{1,t}'\Bigl(X_t^{N,k,s}(\hat\lambda) \Bigr) \,ds\cdot \Bigl(X_{1,t}^{N,k}(\hat\lambda)+\frac{X_{2,t}^{N,k}(\hat\lambda)}{N} \Bigr)\\ &+\int_0^1 \bigg[ \int \partial_{1}{\mbox{{\bf z}}}_t\Bigl(X_t^{N,k,s}(\hat\lambda),y \Bigr)\,dM_N^{X_{t}^{N,s}}(y)\\ &\qquad \qquad \qquad \qquad - \int \partial_{1}{\mbox{{\bf z}}}_t\Bigl(X_{0,t}(\lambda_k),y \Bigr)\,dM_N^{X_{0,t}}(y)\bigg]\,ds\cdot \Bigl(X_{1,t}^{N,k}(\hat\lambda)+\frac{X_{2,t}^{N,k}(\hat\lambda)}{N} \Bigr)\\ &+ \int \partial_{1}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),y \Bigr)\,dM_N^{X_{0,t}}(y)\cdot \Bigl(X_{1,t}^{N,k}(\hat\lambda)+\frac{X_{2,t}^{N,k}(\hat\lambda)}{N} \Bigr)\\ &+\sum_{j=1}^N\int_0^1 \biggl[\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_t^{N,k,s}(\hat\lambda),X_t^{N,j,s}(\hat\lambda)\Bigr)-\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),X_{0,t}(\lambda_j)\Bigr)\biggr]\,ds\cdot X_{1,t}^{N,j}(\hat\lambda)\\ &+\sum_{j=1}^N\int_0^1 \biggl[\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_t^{N,k,s}(\hat\lambda),X_t^{N,j,s}(\hat\lambda)\Bigr)\biggr]\,ds\cdot \frac{X_{2,t}^{N,j}(\hat\lambda)}{N}, \end{split} \end{equation} with the initial condition $X_{2,0}^{N,k}=0$. Using that $$ \|{\mbox{{\bf y}}}_{0,t}\|_{C^{r-2}(\mathbb{R})}\leq C $$ (see Lemma \ref{lem:regular}) we obtain \begin{equation} \label{eq:bound X0} \|X_{0,t}\|_{C^{r-2}(\mathbb{R})}\leq C. \end{equation} We now start to control $X_{1,t}^N$. First, simply by using that $M_N$ has mass bounded by $2N$ we obtain the rough bound $|X_{1,t}^{N,k}|\leq C\,N$. Inserting this bound into \eqref{eq:ODE X2 main} one easily obtain the bound $|X_{2,t}^{N,k}|\leq C\,N^2$. We now prove finer estimates. First, by \eqref{conc} together with the fact that $X_{0,t}$ and $x\mapsto {\mbox{{\bf z}}}_{t}(y, x)$ are Lipschitz (uniformly in $y$), it follows that there exists a finite constant $C$ such that, with probability greater than $1- N^{-N/C}$, \begin{equation} \label{eq:psi2MN} \biggl\| \int {\mbox{{\bf z}}}_t(\cdot, \lambda)\,dM_N^{X_{0,t}}(\lambda)\biggr\|_\infty\leq C\, \log N \sqrt{N}. \end{equation} Hence it follows easily from \eqref{eq:X1} that \begin{equation} \label{eq:bounds X1} \max_k\|X_{1,t}^{N,k}\|_{\infty}\leq C\log N \sqrt{N} \end{equation} outside a set of probability bounded by $N^{-N/C}$. In order to control $X_{2,t}^N$ we first estimate $X_{1,t}^N$ in $L^4(\mathbb P_{V}^N)$: using \eqref{eq:X1} again, we get \begin{multline} \label{eq:ODE X1} \frac{d}{dt}\Bigl(\max_{k}\|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)}\Bigr)\\ \leq C\biggl( \max_{k}\|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)}+1 + \biggl\|\int{\mbox{{\bf z}}}_t(X_{0,t}(\lambda_k), y)\,dM_N^{X_{0,t}}(y)\biggr\|_{L^4(\mathbb P_{V}^N)}\biggr). \end{multline} To bound $X_{1,t}^N$ in $L^4(\mathbb P_{V}^N)$ and then to be able to estimate $X_{2,t}^N$ in $L^2(\mathbb P_{V}^N)$, we will use the following estimates: \begin{lem} For any $k=1,\ldots,N$, \begin{equation} \label{eq:bound psi2 p1} \biggl\|\int{\mbox{{\bf z}}}_t(X_{0,t}(\lambda_k), y)\,dM_N^{X_{0,t}}(y)\biggr\|_{L^4(\mathbb P_{V}^N)} \leq C \log N, \end{equation} \begin{equation} \label{eq:bound psi2} \biggl\|\int \partial_{1}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),y \Bigr)\,dM_N^{X_{0,t}}(y)\biggr\|_{L^4(\mathbb P_{V}^N)} \leq C\log N. \end{equation} \end{lem}\begin{proof} We write the Fourier decomposition of $\eta_{t}(x,y):={\mbox{{\bf z}}}_t(X_{0,t}(x),X_{0,t}(y))$ to get $$\int \eta_{t}(x,y)\, d M_N(y)=\int \hat \eta_{t}(x,\xi) \int e^{i\xi y} \,dM_N(y) \,d\xi\,.$$ Since ${\mbox{{\bf z}}}_t \in C^{u,v}$ for $u+v\le r-5$ and $X_{0,t}\in C^{r-2}$ (see \eqref{eq:bound X0}), we deduce that $$ | \hat\eta_{t}(x,\xi) |\le \frac{C}{1+ |\xi|^{r-5}},$$ so that using \eqref{conc3bb} we get \begin{align*} \biggl\|\sup_x\biggl|\int \eta_{t}(x,y) \,dM_N(y)\biggr|\biggr\|_{L^4(\mathbb P_{V}^N)} &\le \int \Bigl\| \hat\eta_{t}(\cdot,\xi)\Bigr\|_{\infty} \biggl\| \int e^{i\xi y} dM_N(y)\biggr\|_{L^4(\mathbb P_{V}^N)}\,d\xi \\ &\le C\log N \int \Bigl\|\hat\eta_{t}(\cdot,\xi)\Bigr\|_{\infty} \left(1+|\xi|^7\right)\,d\xi\\ & \le C\log N, \end{align*} provided $r>13$. The same arguments work for $\partial_{1} {\mbox{{\bf z}}}_{t}$ provided $r>14.$ Since by assumption $r \geq 15$, this concludes the proof. \end{proof} Inserting \eqref{eq:bound psi2 p1} into \eqref{eq:ODE X1} we get \begin{equation} \label{eq:X1eps} \|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)} \leq C\log N \qquad \forall\,k=1,\ldots,N, \end{equation} which proves the first part of \eqref{boi1}. We now bound the time derivative of the $L^2$ norm of $X_{2,t}^N$: using that $M_N$ has mass bounded by $2N$, in \eqref{eq:ODE X2 main} we can easily estimate $$ \biggl| N\int_0^1\Bigl[{\mbox{{\bf y}}}_{0,t}'\Bigl(X_t^{N,k,s}(\hat\lambda)\Bigr) -{\mbox{{\bf y}}}_{0,t}'\Bigl(X_{0,t}(\lambda_k) \Bigr)\Bigr]\,ds \cdot X_{1,t}^{N,k}(\hat\lambda)\biggr| \leq C|X_{1,t}^{N,k}|^2 + \frac{C}N|X_{1,t}^{N,k}|\,|X_{2,t}^{N,k}|, $$ \begin{multline*} \int_0^1 \bigg| \int \partial_{1}{\mbox{{\bf z}}}_{t}\Bigl(X_t^{N,k,s}(\hat\lambda),y \Bigr)\,dM_N^{X_{t}^{N,s}}(y)- \int \partial_{1}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),y \Bigr)\,dM_N^{X_{0,t}}(y)\bigg|\,ds\\ \leq C|X_{1,t}^{N,k}| + \frac{C}N|X_{2,t}^{N,k}|+\frac{C}N \sum_j \biggl(|X_{1,t}^{N,j}| + \frac{1}N|X_{2,t}^{N,j}|\biggr), \end{multline*} \begin{multline*} \sum_{j=1}^N\int_0^1 \bigg|\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_t^{N,k,s}(\hat\lambda),X_t^{N,j,s}(\hat\lambda)\Bigr) -\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),X_{0,t}(\lambda_j)\Bigr)\bigg|\,ds\,|X_{1,t}^{N,j}|\\ \leq \frac{C}N \sum_j \biggl(|X_{1,t}^{N,j}|^2 + \frac{1}N|X_{2,t}^{N,j}|\,|X_{1,t}^{N,j}|\biggr), \end{multline*} hence \begin{align*} \frac{d}{dt}\|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)}^2&=2\int \sum_k X_{2,t}^{N,k}\cdot \dot X_{2,t}^{N,k}\,d\mathbb P_{V}^N\\ & \leq C \int \sum_k |X_{2,t}^{N,k}|^2\,d\mathbb P_{V}^N + C \int \sum_{k}|X_{1,t}^{N,k}|^2|X_{2,t}^{N,k}|d\mathbb P_{V}^N\\ &+\frac{C}{N} \int \sum_{k}|X_{1,t}^{N,k}| |X_{2,t}^{N,k}|^2d\mathbb P_{V}^N +C \int \sum_{k}|X_{1,t}^{N,k}| |X_{2,t}^{N,k}|\,d\mathbb P_{V}^N\\ &+\frac{C}{N^2}\int\sum_{k}|X_{2,t}^{N,k}|^3\,d\mathbb P_{V}^N + \frac{C}N\int\sum_{k,j} |X_{1,t}^{N,j}|\,|X_{1,t}^{N,k}|\,|X_{2,t}^{N,k}|\,d\mathbb P_{V}^N\\ &+\frac{C}{N^3}\int\sum_{k,j}|X_{2,t}^{N,k}|^2\,|X_{2,t}^{N,j}|\,d\mathbb P_{V}^N\\ &+\sum_{k}\int X_{2,t}^{N,k}\cdot \int_0^1 \biggl[\int \partial_{1}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),y \Bigr)\,dM_N^{X_{0,t}}(y)\biggr]\,ds\cdot X_{1,t}^{N,k}\,d\mathbb P_{V}^N\\ &+ \frac{C}N\int\sum_{k,j} |X_{1,t}^{N,j}|^2\,|X_{2,t}^{N,k}|\,d\mathbb P_{V}^N+\frac{C}{N^2}\int\sum_{k,j}|X_{2,t}^{N,k}|\,|X_{2,t}^{N,j}|\,|X_{1,t}^{N,j}|\,d\mathbb P_{V}^N\\ &+\frac{C}N \int\sum_{k,j}|X_{2,t}^{N,k}|\,|X_{2,t}^{N,j}|\,d\mathbb P_{V}^N. \end{align*} Using the trivial bounds $|X_{1,t}^{N,k}|\leq C\,N$ and $|X_{2,t}^{N,k}|\leq C\,N^2$, \eqref{eq:bound psi2}, and elementary inequalities such as, for instance, $$ \sum_{k,j}|X_{1,t}^{N,j}|\,|X_{1,t}^{N,k}|\,|X_{2,t}^{N,k}| \leq \sum_{k,j}\Bigl(|X_{1,t}^{N,j}|^4+ |X_{1,t}^{N,k}|^4 + |X_{2,t}^{N,k}|^2\Bigr), $$ we obtain \begin{equation} \label{eq:ODEX2} \begin{split} \frac{d}{dt}\|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)}^2 & \leq C\bigg( \|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)}^2 + \int \sum_{k}|X_{1,t}^{N,k}|^4\,d\mathbb P_{V}^N\\ &+ \int \sum_{k}|X_{1,t}^{N,k}|^2\,d\mathbb P_{V}^N +\sum_{k}\log N\, \|X_{2,t}^{N,k}\|_{L^2(\mathbb P_{V}^N)} \|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)}\biggr). \end{split} \end{equation} We now observe, by \eqref{eq:X1eps}, that the last term is bounded by $$ \|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)}^2 + (\log N)^2\sum_{k}\|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)}^2 \leq \|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)}^2 + C\,N(\log N)^4. $$ Hence, using that $\|X_{1,t}^{N,k}\|_{L^2(\mathbb P_{V}^N)} \leq \|X_{1,t}^{N,k}\|_{L^4(\mathbb P_{V}^N)}$ and \eqref{eq:X1eps} again, the right hand side of \eqref{eq:ODEX2} can be bounded by $C\|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)} ^2+ C\,N(\log N)^4$, and a Gronwall argument gives $$ \|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)}^2 \leq C\, N(\log N)^4, $$ thus $$ \|X_{2,t}^N\|_{L^2(\mathbb P_{V}^N)} \leq C\,N^{1/2}(\log N)^2, $$ concluding the proof of \eqref{boi1}. We now prove \eqref{boi2}: using \eqref{eq:X1} we have \begin{align*} &|\dot X_{1,t}^{N,k}(\hat\lambda)-\dot X_{1,t}^{N,k'}(\hat\lambda)|\\ &\leq |{\mbox{{\bf y}}}_{0,t}'(X_{0,t}(\lambda_k))-{\mbox{{\bf y}}}_{0,t}'(X_{0,t}(\lambda_{k'})) |\,|X_{1,t}^{N,k}(\hat\lambda)|\\ &+|{\mbox{{\bf y}}}_{0,t}'(X_{0,t}(\lambda_{k'})) |\,|X_{1,t}^{N,k}(\hat\lambda)-X_{1,t}^{N,k'}(\hat\lambda)| + |{\mbox{{\bf y}}}_{1,t}(X_{0,t}(\lambda_k))-{\mbox{{\bf y}}}_{1,t}(X_{0,t}(\lambda_{k'}))|\\ &+\biggl|\int \Bigl({\mbox{{\bf z}}}_t(X_{0,t}(\lambda_k), y)- {\mbox{{\bf z}}}_t(X_{0,t}(\lambda_{k'}), y)\Bigr)\,dM_N^{X_{0,t}}(y)\biggr|\\ &+\frac1N\sum_{j=1}^N\int_0^1 \Bigl|\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_k),X_{0,t}(\lambda_j)\Bigr) -\partial_{2}{\mbox{{\bf z}}}_{t}\Bigl(X_{0,t}(\lambda_{k'}),X_{0,t}(\lambda_j)\Bigr) \Bigr|\,ds\,|X_{1,t}^{N,j}(\hat\lambda)|. \end{align*} Using that $|X_{0,t}(\lambda_k)-X_{0,t}(\lambda_{k'})| \leq C|\lambda_k - \lambda_{k'}|$, the bound \eqref{eq:bounds X1}, the Lipschitz regularity of ${\mbox{{\bf y}}}_{0,t}'$, ${\mbox{{\bf y}}}_{1,t}$, ${\mbox{{\bf z}}}_t$, and $\partial_{2}{\mbox{{\bf z}}}_{t}$, and the fact that $$ \biggl\| \int \partial_{1}{\mbox{{\bf z}}}_{t}(\cdot, \lambda)\,dM_N^{X_{0,t}}(\lambda)\biggr\|_\infty\leq C\, \log N \sqrt{N} $$ with probability greater than $1- N^{-N/C}$ (see \eqref{conc}), we get $$ |\dot X_{1,t}^{N,k}(\hat\lambda)-\dot X_{1,t}^{N,k'}(\hat\lambda)| \leq C|X_{1,t}^{N,k}(\hat\lambda)- X_{1,t}^{N,k'}(\hat\lambda)|+C\log N \sqrt{N}|\lambda_k - \lambda_{k'}| $$ outside a set of probability less than $ N^{-N/C}$, so \eqref{boi2} follows from Gronwall. \end{proof} \section{Transport and universality}\label{sect:univ} In this section we prove Theorem \ref{thm:univ} on universality using the regularity properties of the approximate transport maps obtained in the previous sections. \begin{proof}[Proof of Theorem \ref{thm:univ}] Let us first remark that the map $T_0$ from Theorem \ref{thm:T} coincides with $X_{0,1}$, where $X_{0,t}$ is the flow defined in \eqref{eq:X0}. Also, notice that $X^N_1:\mathbb{R}^N\to\mathbb{R}^N$ is an approximate transport of $\mathbb P^N_V$ onto $\mathbb P_{V+W}^N$ (see Lemma \ref{lemma:flow} and Proposition \ref{prop:key}). Set $\hat X_1^N:=X_{0,1}^N+\frac{1}{N}X_{1,1}^N$, with $X_{0,t}^N$ and $X_{1,t}^N$ as in Lemma \ref{the flow}. Since $X_1^N-\hat X_1^N=\frac{1}{N^2}X_{2,1}^N$, recalling \eqref{boi1} and using H\"older inequality to control the $L^1$ norm with the $L^2$ norm, we see that \begin{equation} \label{eq:Xhat} \begin{split} \biggl|\int g(\hat X_1^N)\,d\mathbb P_V^N - \int g(X_1^N)\,d\mathbb P_V^N \biggr|& \leq \|\nabla g\|_\infty \frac{1}{N^2} \int |X_{2,1}^N|\,d\mathbb P_V^N \\ &\leq \|\nabla g\|_\infty \frac{1}{N^2} \|X_{2,1}^N\|_{L^2(\mathbb P_V)} \\ &\leq C \|\nabla g\|_\infty \frac{(\log N)^2}{N^{3/2}}. \end{split} \end{equation} This implies that also $\hat X^N_1:\mathbb{R}^N\to\mathbb{R}^N$ is an approximate transport of $\mathbb P^N_V$ onto $\mathbb P_{V+W}^N$. In addition, we see that $\hat X_1^N$ preserves the order of the $\lambda_i$ with large probability. Indeed, first of all $X_{0,t}:\mathbb{R}\to \mathbb{R}$ is the flow of ${\mbox{{\bf y}}}_{0,t}$ which is Lipschitz with some constant $L$, hence differentiating \eqref{eq:X0} we get $$ \frac{d}{dt}\bigl|X_{0,t}'\bigr| \leq \bigl|{\mbox{{\bf y}}}_{0,t}'(X_{0,t})\bigr|\,\bigl|X_{0,t}'\bigr| \leq L\,\bigl|X_{0,t}'\bigr|, \qquad X_{0,0}'=1, $$ so Gronwall's inequality gives the bound $$ e^{-Lt}\leq \bigl|X_{0,t}'\bigr|\leq e^{Lt} $$ Since $X_{0,0}'=1$, it follows by continuity that $X_{0,t}'$ must remain positive for all time and it satisfies \begin{equation} \label{eq:x0prime} e^{-Lt}\leq X_{0,t}'\leq e^{Lt}, \end{equation} from which we deduce that $$ e^{-Lt}\bigl(\lambda_j - \lambda _i\bigr) \leq X_{0,t}(\lambda_j) - X_{0,t}(\lambda _i) \leq e^{Lt}\bigl(\lambda_j - \lambda _i\bigr), \qquad \forall\,\lambda_i<\lambda_j. $$ In particular, $$ e^{-L}\bigl(\lambda_j - \lambda _i\bigr) \leq X_{0,1}(\lambda_j) - X_{0,1}(\lambda _i) \leq e^{L}\bigl(\lambda_j - \lambda _i\bigr). $$ Hence, using the notation $\hat\lambda=(\lambda_1,\ldots,\lambda_N)$, since $$ \biggl|\frac{1}{N}X_{1,t}^{N,j}(\hat\lambda)-\frac{1}NX_{1,t}^{N,i}(\hat\lambda)\biggr|\le C\,\frac{\log N}{ \sqrt{N}}|\lambda_i-\lambda_{j}| $$ (see \eqref{boi2}) with probability greater than $1-N^{-N/C}$ we get $$ \frac{1}{C}\bigl(\lambda_j - \lambda _i\bigr) \leq \hat X_{1}^{N,j}(\hat \lambda) - \hat X_{1}^{N,i}(\hat\lambda ) \leq C\bigl(\lambda_j - \lambda _i\bigr) $$ with probability greater than $1-N^{-N/C}$. We now make the following observation: the ordered measures $\tilde P_V^N$ and $\tilde P_{V+W}^N$ are obtained as the image of $\mathbb P_V^N$ and $\mathbb P_{V+W}^N$ via the map $R:\mathbb{R}^N \to \mathbb{R}^N$ defined as $$ [R(x_1,\ldots,x_N)]_i:=\min_{\sharp J=i}\max_{j\in J} x_j. $$ Notice that this map is $1$-Lipschitz for the sup norm. Hence, if $g$ is a function of $m$-variables we have $ \|\nabla(g\circ R)\|_\infty \leq \sqrt{m} \|\nabla g\|_\infty$, so by Lemma \ref{lemma:flow}, Proposition \ref{prop:key}, and \eqref{eq:Xhat}, we get $$ \biggl|\int g\circ R (\hat X_1^N) \,d\mathbb P_V^N-\int g\circ R\, d\mathbb P_{V+W}^N\biggr| \leq C\frac{(\log N)^3 }{N}\|g\|_\infty+C\sqrt{m}\,\frac{(\log N)^2}{N^{3/2}} \|\nabla g\|_\infty. $$ Since $\hat X_1^N$ preserves the order with probability greater than $1-N^{-N/C}$, we can replace $g\circ R (N \hat X_1^N)$ with $g(N \hat X_1^N\circ R)$ up to a very small error bounded by $\|g\|_\infty N^{-N/C}$. Hence, since $R_\# \mathbb P_V^N=\tilde P_V^N$ and $R_\# \mathbb P_{V+W}^N=\tilde P_{V+W}^N$, we deduce that, for any Lipschitz function $f:\mathbb{R}^m\to \mathbb{R}$, \begin{multline*} \bigg|\int f\bigl(N(\lambda_{i+1}-\lambda_i),\ldots,N(\lambda_{i+m}-\lambda_i)\bigr)\, d\tilde P_{V+W}^N\qquad\\ \quad -\int f\Bigl(N\bigl(\hat X_1^{N,i+1}(\hat\lambda)-\hat X_1^{N,i}(\hat\lambda)\bigr), \ldots,N\bigl(\hat X_1^{N,i+m}(\hat\lambda)-\hat X_1^{N,i}(\hat\lambda)\bigr)\Bigr) d\tilde P_{V}^N\bigg| \\ \le C\, \frac{(\log N)^3 }{N}\|f\|_\infty+C\,\sqrt{m}\,\frac{(\log N)^2}{N^{1/2}}\|\nabla f\|_\infty. \end{multline*} Recalling that $$ \hat X_1^{N,j}(\hat\lambda)=X_{0,1}(\lambda_j)+\frac{1}{N}X_{1,1}^{N,j}(\hat\lambda), $$ we observe that, as $X_{0,1}$ is of class $C^2$, $$X_{0,1}(\lambda_{i+k})-X_{0,1}(\lambda_{i})=X_{0,1}'(\lambda_i)\,(\lambda_{i+k}-\lambda_{i})+O(|\lambda_{i+k}-\lambda_{i}|^2).$$ Also, by \eqref{boi2} we deduce that, out of a set of probability bounded by $ N^{-N/C}$, \begin{equation}\label{toto} |X_{1,1}^{N,i+k}(\hat\lambda)-X_{1,1}^{N,i}(\hat\lambda)|\le C\,\log N \sqrt{N}|\lambda_{i+k}-\lambda_i|\,. \end{equation} As $X_{0,1}'(\lambda_i) \geq e^{-L}$ (see \eqref{eq:x0prime}) we deduce $$\frac{1}{N}\,|X_{1,1}^{N,i+k}(\hat\lambda)-X_{1,1}^{N,i}(\hat\lambda)|\le C\,|X_{0,1}'(\lambda_i)\,(\lambda_{i+k}-\lambda_{i})| \,\frac{\log N}{{N}^{1/2}}$$ and $$ O(|\lambda_{i+k}-\lambda_{i}|^2)=O\bigl(\bigl|X_{0,1}'(\lambda_i)\,(\lambda_{i+k}-\lambda_{i})\bigr|^2\bigr) $$ hence with probability greater than $1- N^{-N/C}$ it holds \begin{align*} \hat X_1^{N,i+k}(\hat\lambda)-\hat X_1^{N,i}(\hat\lambda)= X_{0,t}'(\lambda_i)\,(\lambda_{i+k}-\lambda_{i}) \biggl[1+ O\biggl(\frac{\,\log N}{N^{1/2}} \biggr)+O\bigl(\bigl|X_{0,1}'(\lambda_i)\,(\lambda_{i+k}-\lambda_{i})\bigr|\bigr)\biggr]\,. \end{align*} Since we assume $f$ supported in $[-M,M]^m$, the domain of integration is restricted to $\hat\lambda$ such that $\{NX_{0,t}'(\lambda_i)\,(\lambda_i-\lambda_{i+k})\}_{1\le k \le m}$ is bounded by $2M$ for $N$ large enough, therefore $$\hat X_1^{N,i+k}(\hat\lambda)-\hat X_1^{N,i}(\hat\lambda)= X_{0,t}'(\lambda_i)\,(\lambda_{i+k}-\lambda_{i}) +O\biggl(2M\frac{\,\log N}{N^{3/2}}\biggr) +O\biggl(\frac{4M^2}{N^2}\biggr),$$ from which the first bound follows easily. For the second point we observe that $a_{V+W}=X_{0,1}(a_V)$ and, arguing as before, \begin{multline*} \bigg| \int f\bigl(N^{2/3}(\lambda_1-a_{V+W}), \ldots,N^{2/3}(\lambda_m-a_{V+W})\bigr) \,d\tilde P_{V+W}^N\\ -\int f\Bigl(N^{2/3}\bigl(\hat X_1^{N,1}(\hat\lambda)-X_{0,1}(a_V)\bigr), \ldots,N^{2/3}\bigl(\hat X_1^{N,m}(\hat\lambda)-X_{0,1}(a_V)\bigr)\Bigr)\, d\tilde P_{V}^N\bigg|\\ \le C\, \frac{(\log N)^3 }{N}\,\|f\|_\infty+C\,\sqrt{m}\frac{(\log N)^2}{N^{5/6}}\,\|\nabla f\|_\infty. \end{multline*} Since, by \eqref{boi1}, \begin{align*} \hat X_1^{N,i}(\lambda)&=X_{0,1}(\lambda_i) +O_{L^4(\mathbb P_{V}^N)}\biggl(\frac{\log N}{N}\biggr)\\ &= X_{0,1}(a_V) +X_{0,1}'(a_V)\,(\lambda_i-a_V)+O\bigl(|\lambda_i-a_V|^2\bigr)+O_{L^4(\mathbb P_{V}^N)}\biggl(\frac{\log N}{N}\biggr), \end{align*} we conclude as in the first point. \end{proof} \bibliographystyle{alpha}
\section{The BinaMIcS project} The BinaMIcS project has been awarded "Large Program" status with two high resolution spectropolarimeters: ESPaDOnS at the CFHT (Hawaii; PIs Alecian/Wade) for 604 hours over 4 years from 2013 to 2016, and Narval at TBL (France; PI Neiner) for 128 hours over 2 years from 2013 to 2014 (renewable for 2 more years). This large amount of time is being used to acquire an immense database of sensitive measurements of polarized and unpolarized spectra of spectroscopic binary (SB2) stars. The program includes 3 components: the detailed study of some known magnetic cool binaries, the detailed study of the few known magnetic hot close binaries, and a survey of a large number of hot binaries to search for magnetic fields. The detailed studies will allow us to obtain magnetic maps and test various models, while the survey part of the project will allow us to obtain statistical results. This database will be combined with new and archival complementary data (e.g. optical photometry, UV and X-ray spectroscopy) as well as theoretical studies, numerical simulations, and modeling (described hereafter), and applied to address the 4 main scientific objectives of the BinaMIcS project: i) what is the impact of magnetic fields during stellar formation, and vice versa; ii) how do tidally-induced internal flows impact fossil and dynamo fields; iii) how do magnetospheric star-star interactions modify stellar activity; iv) what is the magnetic impact on angular momentum exchanges and mass transfers. \section{Roadmap on the theoretical work of BinaMIcS} Magnetic fields are a crucial ingredient in a star's evolution, influencing its formation, the structure of its atmosphere and interior, as well as controlling the interaction with its environment. For binary stars, magnetism is even more significant, as magnetic fields in binary systems will be strongly affected by, and may also strongly affect, the transfer of energy, mass and angular momentum between the components. Therefore, the interplay between stellar magnetic fields and binarity has to be investigated in detail, from both the observational and theoretical point-of-view. First, the incidence and characteristics of magnetic fields are key parameters for understanding the physics of binaries. In higher-mass stars (above 1.5 $M_{\odot}$) the incidence of magnetic stars in binary systems provides a basic constraint on the detailed origin of the magnetic field, assumed to be a fossil remnant, and whether such strong magnetic fields suppress binary formation or are a result of mergers. Next, in low-mass stars, tidal interactions are expected to induce large-scale 3D shear and/or helical flows in stellar interiors that can significantly perturb the stellar dynamo. Similar flows may also influence the fossil magnetic fields of higher-mass stars. Finally, magnetically driven winds/outflows in cool and hot close binary systems have long been suspected to be responsible for their orbital evolution, while magnetospheric interactions have been proposed to enhance stellar activity. However, the crucial observational constraints required to test these hypotheses are, at present, nearly nonexistent, and dedicated theoretical studies are mandatory to bring the studies of binary (and multiple) stars to a new level of understanding. Within BinaMIcS, we will therefore study theoretically the complex interactions between tides and magnetic fields, i.e.: i) how magnetic fields modify tidal flows described by \cite[Zahn 1977, Ogilvie \& Lin 2007, Le Bars et al. 2010, and Remus et al. 2012]{Zahn 1977,OgilvieLin2007,LeBarsetal2010,Remusetal2012} and the associated torques applied on each component; ii) how flows driven by tides (as well as precession and libration) can modify dynamo mechanisms \cite[(e.g. Le Bars et al. 2011)], because of the angular momentum they transport ($\Omega$-effect) and of their helicity ($\alpha$-effect), and the stability of fossil fields; iii) how such external mechanical forcings can compete with convective driving as a function of the mass ratio and of the separation of the components; iv) how do the combined action of tides and magnetic fields modify the orbital dynamics of binary stars and the evolution of their components \cite[(e.g. de Mink et al. 2013, Mathis \& Remus 2013, Song et al. 2013)]{deMinketal2013,MathisRemus2013,Songetal2013}. Moreover, we will study the complex MHD interactions between the components of binary stars, i.e.: i) the interactions between stellar wind emitted by each component and the related torques \cite[(e.g. Strugarek et al. 2012)]{Strugareketal2012}; ii) the magnetospheric interactions and the associated applied torques, helicity exchanges, and modifications of the magnetic activity of the components \cite[(e.g. Lanza 2012 and the contribution by S. Gregory in this proceeding)]{Lanza2012}. Results will also be applied to the study of star-planet interactions.
\section{Introduction } Cold atomic ensembles interacting with electromagnetic fields are powerful tools in quantum information science with applications that include quantum memory \cite{FLukinMemory, Polzik2004, ChoiKimbleMemory}, quantum communication \cite{DLCZ, MatKuzNetwork}, continuous variable quantum computing \cite{BraunReview}, and metrology \cite{AppelPolzikMet, VulMet}. At the heart of these protocols is the strong coupling between a quantum mode of the field and an effective collective spin of the ensemble. This coupling can generate entanglement between atoms and photons, such that measurement of the light yields strong quantum backaction on the atoms. Photons can also enable a quantum data bus for entangling atoms with one another. Enhancing the atom-light interface is thus essential for improving the performance of quantum technologies and for reaching new regimes where a quantum advantage becomes manifest. This can be achieved through confined modes such as in optical cavities~\cite{Kimble2005, VulMet, Thompson2011} or waveguides in optical nanostructures~\cite{Rauschenbeutel2010, Waks2012, Kimble2013}. Strong atom-photon coupling can also occur in free space in the interaction between light and an extended ensemble of atoms. This occurs when photons are scattered collectively by the ensemble, and interference enhances the radiation into the probe mode relative to diffuse scattering into $4\pi$ steradians~\cite{Vuletic2011, Kaiser2013}. Early experiments demonstrated such strong coupling and entanglement in high pressure vapor cells where a one-dimensional description of plane wave modes and uniform atomic density is applicable \cite{Kuzmich1999}. This theory accurately describes a variety of experiments including the entanglement of macroscopic ensembles in remote vapor cells \cite{Polzik2001} and quantum memory for continuous variables~\cite{Polzik2004}. \begin{figure} \includegraphics[width=1\hsize]{figure_1} \caption{Schematic of a linearly polarized laser probe with a Gaussian spatial mode (blue) interacting dispersively with a cold, trapped atomic ensemble. The light that is collectively scattered by the average density distribution of the ensemble defines the radiated mode (red). The interference of the radiated and probe modes leads to a rotation of the field polarization according to the Faraday effect. When measured in a polarimeter, this can be used to generate spin squeezing in the ensemble. The spatial overlap of the collectively scattered field and the probe, a measure of the strong atom-light coupling, depends highly on geometry. In addition, diffusely scattered photons due to local density fluctuations in the ensemble lead to decoherence in the collective spin variables. } \label{Fig::Schematic} \end{figure} More recently, experiments have employed ensembles of ultracold atoms in pencil-shaped dipole traps probed by focused laser beams \cite{KamMul12, KosMit10}. When the radiation pattern of the atomic ensemble is effectively matched with the paraxial mode of the probe, the atomic dipoles are indistinguishable and the scattering is cooperative. Such geometries have the potential to strongly enhance the atom-photon quantum interface, but their description is more complex, requiring a full treatment of diffraction, inhomogeneous coupling, radiation patterns, and decoherence. Harnessing the advantages of these atomic ensembles thus requires a three-dimensional quantum theory of the underlying interaction, including both coherent coupling and quantum noise. In the last decade there has been significant progress in developing a three-dimensional quantum description of the atom-light interface. A rigorous field-theoretic treatment separates the mean-field classical effects from the quantum fluctuations and noise, including the spatial inhomogeneities of atoms and light modes~\cite{SorSor08}. Models that include spatial modes have been developed in a variety of contexts~\cite{Kuzmich2004, Windpassinger2008, Koschorreck2009, SauSta10}. Applications include remote entanglement via collective Raman scattering in a DLCZ-type protocol \cite{DuaZol02, SorSor09} and for multimode quantum memories \cite{ZueGro11}. From such studies, it is clear that one-dimensional models not only fail to describe the relevant coherent and incoherent effects, but they also do not take advantage of the resources associated with spatial modes~\cite{GroSor12, HigBuc12}. In this paper we revisit the three-dimensional atom-light interface with particular emphasis on spin squeezing through QND measurement of the collective spin via the Faraday effect \cite{KuzBig00,KosMitSq, Takano2009}, shown schematically in Fig.~\ref{Fig::Schematic}. In this protocol, the key interaction is the off-resonant scattering of horizontally polarized photons into vertical polarization. Measurement in a balanced polarimeter corresponds to a homodyne measurement of these scattered photons. The degree of scattering into the local oscillator, defined by the paraxial laser mode, determines the measurement strength and the resulting backaction that generates squeezing. Central to this problem are the spatial modes of the light and the collective spin waves of the atomic ensemble. In the one-dimensional model, one collective parameter defines the strong-coupling regime of the atom-photon interface, the optical density on resonance, OD $= \eta \sigma_0 L = N \sigma_0 /A$, where $\eta$ is the atomic density, $\sigma_0$ is the resonant scattering cross section, $L$ is the length of the vapor cell, and $N$ is the number of atoms in the volume $V=AL$ for a uniform beam of area $A$. In contrast, in a fully three-dimensional model, where the atomic density, $\eta(\mbf{r})$, and paraxial beam intensity distribution, $\beta(\mbf{r})$, are not uniform, there is a collection of parameters that dictate the strong-coupling regime. Different effective atom numbers, $N_{\rm eff}^{(K)} =\int d^3\mbf{r} \, \eta (\mbf{r}) \left[\beta(\mbf{r})\right]^K$, govern different physical effects. For example, $N_{\rm eff}^{(1)}$ determines the mean Faraday signal in the polarimeter, while $N_{\rm eff}^{(2)}$ determines the size of the measurement uncertainty from spin projection noise. The entangling strength of the atom-light interface is determined by the size of the spin projection uncertainty compared to the quantum uncertainty in the measured light quadratures (shot noise). This collective interaction is proportional to an effective optical density, $OD_\text{eff} = N^{(2)}_\text{eff} \sigma_0/A$. In contrast, decoherence acts locally on the atoms in a noncollective manner, and the noise injected into the system due to optical pumping and spin flips is governed by other parameters. A proper accounting of the balance between the coherent coupling and decoherence is especially challenging given the tensor nature of the atom-photon interaction of real alkali atoms. Previous treatments of quantum noise in a multimode Faraday-based atom-light interface have been carried out in a one-dimensional model~\cite{KupPol05, VasSor12}. Our goal is to extend this to the three-dimensional case. In this work, we derive a stochastic master equation describing the dynamics of the collective atomic state conditioned on balanced polarimetry measurements, including the effects of measurement backaction, collective decoherence from unmeasured paraxial light, and local decoherence from diffuse photon scattering that gives rise to optical pumping. While we apply this to study conditional spin squeezing generated by a QND measurement, the formalism we develop is broadly applicable to other protocols where a strong, free-space atom-light interface is essential, and where measurement backaction may be a tool for induced atom-atom interactions. The remainder of the article is organized as follows. We lay out the physical model for an ensemble of alkali atoms dispersively interacting with a coherent probe laser in Sec. \ref{Sec::ParaxialScattering}. We begin with a semiclassical model that can be used to describe the scattered paraxial fields and to identify the collective spin wave that is coupled to the laser mode. To understand the entangling Faraday interaction in a multimode geometry, we then present a fully quantum mechanical model. This serves as the cornerstone for a complete description of QND squeezing and allows us to account for the damaging effects of decoherence. When the output light is measured continuously, the quantum dynamics, including the combined effects of measurement backaction and decoherence, are described by a stochastic master equation. We use this fully quantum mechanical atom-light description to study effects of spatial modes on the squeezing of spin waves in Sec. \ref{Sec::SpinSqueezing}. In particular we use the multimode description to model the dynamics of spin squeezing and to analyze the dependence of peak squeezing on cloud and beam geometry. We use numerical simulations to help build physical intuition about the three-dimensional atom-light interface and to investigate how the model can be used to optimize an experimental design. Finally, we summarize our results and present future directions for this work in Sec.~\ref{Sec::Conclusion}. \section{Paraxial Atom-Light Interface} \label{Sec::ParaxialScattering} When driven by an off-resonant laser field such that the excited state probability is small, atoms elastically scatter electromagnetic waves in a manner equivalent to a set of linearly polarizable particles. Thus, a great deal of qualitative and quantitative information can be obtained from classical radiation theory. In a rigorous field-theoretic analysis, S$\o$rensen and S$\o$rensen showed that the mean-field effect of the light interacting with an atomic ensemble gives rise to an index of refraction of the gas, while fluctuations are due to the random positions of the atoms and the vacuum noise of the light \cite{SorSor08}. In particular, the index of refraction is due to the spatially-averaged local density of the atoms, while the diffuse scattering into 4$\pi$ arises from the random positions of the point atomic scatterers and is equivalent to decoherence by local spontaneous emission. This diffuse scattering, which leads to attenuation of the incident wave and optical pumping of the atomic state, is accounted for by an imaginary part of the polarizability according to the optical theorem. We can thus break up the problem into two pieces. First, the mean field effect is described by classical scattering of a laser beam incident on a linearly polarizable dielectric whose shape is determined by the atomic density distribution. For a paraxial probe beam and an extended cloud, the scattered field is also paraxial, and the solution is easily found by Fraunhofer scattering theory~\cite{Newton1982}. As we are interested in the Faraday effect, we include the tensor nature of the atomic polarizability. Scattering of an incident horizontal polarization to an orthogonal vertical polarization is the key effect that we seek to measure in the polarimeter. Second, to properly account for quantum backaction on the atoms resulting from measurement and to describe the decoherence due to diffuse scattering and optical pumping, we turn to the fully quantum theory. \subsection{Semiclassical theory} \label{Sec::SemiclassicalTheory} Consider the scattering of an incident paraxial laser beam with frequency $\omega_0$ and complex amplitude, $\mbf{E}_L(\mbf{r}_\perp, z) = \vec{\mathcal{E}}_L (\mathbf{r}_\perp , z) e^{ik_0 z}$, by a particle located at a position $\mbf{r}'$ with dynamical tensor electric-dipole polarizability $\tensor{\alpha}.$ The field envelope has the standard form $\vec{\mathcal{E}}_L(\mbf{r}_\perp, z) = \vec{\epsilon}\, \mathcal{E}_0 u_{00}(\mathbf{r}_\perp,z) $, where $\vec{\epsilon}$ is the laser polarization and $u_{00}$ is chosen to be the Gaussian TEM$_{00}$ mode given by \begin{equation} u_{00}(\mathbf{r}_\perp,z)= \frac{w_0}{w(z)}e^{-\frac{\left|\mbf{\mathbf{r}_\perp}\right|^2}{[w(z)]^2}}e^{\frac{ik_0 \left|\mbf{\mathbf{r}_\perp}\right|^2}{2 R(z)}}e^{-i\Phi(z)}. \end{equation} The $z$-dependent beam waist, the radius of curvature of the phase fronts, and the Guoy phase are given by \begin{subequations} \label{Eq::GaussianParameters} \begin{align} w(z) &= w_0 \sqrt{1+(z/z_R)^2} , \\ R(z) &= z\left(1+(z_R/z)^2\right) , \\ \Phi(z) &= \tan^{-1}(z/z_R), \end{align} \end{subequations} respectively, with beam waist $w_0$ and Rayleigh range $z_R \equiv k_0 w_0^2/2$. In the first Born approximation, the scattered field amplitude is that radiated by the induced dipole, \begin{align} \label{Eq::DipoleScattering} \mbf{E}_{\rm scat}(\mbf{r})&= k_0^2 \left[\tensor{\alpha}\cdot \mbf{E}_L(\mbf{r}')\right]_\perp \frac{e^{ik_0 \left|\mbf{r}-\mbf{r}'\right|}}{\left|\mbf{r}-\mbf{r}'\right|} \nonumber \\ & \approx k_0^2 \left[ \tensor{\alpha}\cdot \mbf{E}_L(\mbf{r}')\right]_\perp e^{ik_0(z-z')}\frac{e^{\frac{i k_0 \left|\mathbf{r}_\perp-\mathbf{r}_\perp'\right|^2}{2(z-z')}}}{z-z'} , \end{align} where the subscript $\perp$ denotes the component of the dipole transverse to the direction of observation. The last approximation is valid for paraxial points of observation, $z \gg \left|\mathbf{r}_\perp\right|$. Gaussian-cgs units for the electromagnetic field equations are used throughout. Because the dipole radiation is not mode-matched with the Gaussian laser beam, the light is scattered into all paraxial modes as well as off-axis nonparaxial modes. In the far field, $z \gg z'$, the total field takes the form, \begin{align} \mbf{E}_{\rm out}(\mbf{r}) & =\mbf{E}_L(\mbf{r}) + \mbf{E}_{\rm scat}(\mbf{r}) \nonumber \\ & = \big( \vec{\epsilon} + \vec{\Upsilon} \big) \mathcal{E}_0 e^{ik_0 z}u_{00}(\mathbf{r}_\perp,z)+ \mbf{E}'_{\rm scat}(\mbf{r}), \end{align} where $\mbf{E}'_{\rm scat}(\mbf{r})$ is the scattered field into all spatial modes other than the probe mode, and as shown in Appendix A, \erf{Eq::ClassicalScattering}, \begin{align} \vec{\Upsilon} \mathcal{E}_0 e^{ik_0 z} &\equiv \int \frac{d^2\mathbf{r}_\perp}{A} u_{00}^*(\mathbf{r}_\perp,z) \mbf{E}_{\rm scat}(\mbf{r}) \nonumber \\ & = i \frac{2\pi k_0}{A}\left(\tensor{\alpha}\cdot \vec{\epsilon}\right)_\perp \left| u_{00}(\mathbf{r}_\perp',z') \right|^2 \mathcal{E}_0 e^{ik_0 (z-z')} \end{align} is the field amplitude ``forward scattered'' into the laser mode. $A = \int d^2\mathbf{r}_\perp |u_{00}(\mathbf{r}_\perp,z)|^2 = \pi w_0^2/2$ is the effective beam area. The key physical effects are seen in these equations. The component of the radiated field vector $\vec{\Upsilon}$ along the laser polarization $\vec{\epsilon}$ gives rise to the scalar index of refraction and attenuation. The component of $\vec{\Upsilon}$ orthogonal to $\vec{\epsilon}$ gives rise to a rotation of the polarization on the Poincar\'{e} sphere -- Faraday rotation and birefringence. For example, suppose the laser is linearly polarized along $x$ ($\vec{\epsilon} =\mbf{e}_x)$. The total field thus can be written, \begin{align} \label{Eq::OutputField} & \mbf{E}_{\rm out} (\mbf{r}) = \mbf{E}'_{\rm scat}(\mbf{r}) + \\ &\quad \left[ \left(1+i\delta \phi-\frac{a}{2}\right) \mbf{e}_x + \left(\frac{\chi + i\beta}{2}\right) \mbf{e}_y \right] \mathcal{E}_0 e^{ik_0 z}u_{00}(\mathbf{r}_\perp,z), \nonumber \end{align} where, \begin{align} \delta \phi &=(2 \pi k_0/A) \left| u_{00}(\mathbf{r}_\perp',z') \right|^2 \text{Re} \left\{ \alpha_{xx} \right\} \nonumber \\ a &= (4 \pi k_0/A) \left| u_{00}(\mathbf{r}_\perp',z') \right|^2 \text{Im} \left\{ \alpha_{xx} \right\} \nonumber \\ \chi &=-(4 \pi k_0/A) \left| u_{00}(\mathbf{r}_\perp',z') \right|^2 \text{Im} \left\{ \alpha_{yx} \right\} \nonumber \\ \beta &=(4 \pi k_0/A) \left| u_{00}(\mathbf{r}_\perp',z') \right|^2 \text{Re} \left\{ \alpha_{yx} \right\} \nonumber \end{align} are respectively: $\phi$ is the index of refraction phase shift, $a$ is the Beer's law attenuation coefficient, $\chi$ is the rotation angle of the Stokes vector corresponding to the Faraday effect, and $\beta$ is the corresponding angle for birefringence, with the polarizability matrix elements denoted as $\alpha_{ij} = \mathbf{e}_i\cdot \tensor{\alpha} \cdot \mathbf{e}_j$ in the $x$-$y$ basis. The above description of the atom-field coupling is most easily generalized using the theory of scattering of paraxial waves \cite{MulPol05}; details can be found in Appendix \ref{Appendix::ParaxialScattering}. The mean field is described by the electric field envelope $\vec{\mathcal{E}}_L (\mathbf{r}_\perp , z,t)= \mathcal{A}(t-z/c) \vec{\mathcal{U}} (\mathbf{r}_\perp, z)$, where $\mathcal{A}(t)$ is the temporal pulse envelope evaluated at the retarded time, and $\vec{\mathcal{U}} (\mathbf{r}_\perp, z)$ is the spatial envelope satisfying the paraxial wave equation, \begin{equation} \frac{\partial}{\partial z}\vec{\mathcal{U}}(\mathbf{r}_\perp , z) \! = \!\frac{i}{2 k_0}\nabla_\perp^2 \vec{\mathcal{U}}(\mathbf{r}_\perp , z) +i2 \pi k_0 \tensor{\chi}(\mathbf{r}_\perp, z) \cdot \, \vec{\mathcal{U}}(\mathbf{r}_\perp , z), \end{equation} with spatially averaged dielectric susceptibility $\tensor{\chi}(\mathbf{r}_\perp, z)$. The scattering solution to this equation is well known \cite{Newton1982}. In the first Born approximation, i.e. for dilute samples where multiple scattering is negligible, given an incident field $\vec{\mathcal{U}}_{\rm in} (\mathbf{r}_\perp,z)$, the total field is \begin{align} &\vec{\mathcal{U}}(\mathbf{r}_\perp , z) = \vec{\mathcal{U}}_{\rm in} (\mathbf{r}_\perp,z) \label{paraxsol} \\ &\! + i 2 \pi k_0 \! \! \int_{-\infty}^z \! \! \! \! \! \! dz' \! \! \int \! \! d^2 \mathbf{r}_\perp' K(\mathbf{r}_\perp \! \! -\! \mathbf{r}_\perp', z\!-\!z') \tensor{\chi}( \mathbf{r}_\perp', z') \! \cdot \vec{\mathcal{U}}_{\rm in} (\mathbf{r}_\perp',z') , \nonumber \end{align} where $K(\mathbf{r}_\perp-\mathbf{r}_\perp', z-z')$ is the paraxial propagator. This solution is the superposition of incident and reradiated dipole fields. The solution for a paraxial field scattered from a point dipole at position $\mbf{r}'$, \erf{Eq::DipoleScattering}, is recovered by setting $\tensor{\chi}(\mbf{r}) = \tensor{\alpha} \, \delta^{(3)}(\mbf{r}-\mbf{r}') $. The diagonal matrix elements of the susceptibility give rise to the index of refraction and a slight distortion of the wavefront of the beam. We can neglect this effect for dilute gases, though it is easily accounted for. The Faraday effect arises from the scattering of initially $x$-polarized light into orthogonal $y$-polarization as discussed above, governed by the off-diagonal element of the dielectric susceptibility matrix, $\chi_{yx}$. To measure Faraday rotation, one employs a balanced polarimeter at $\pm 45^{\circ}$, so that the signal $\mathcal{M}$ is proportional to $\mathcal{U}_x^* \mathcal{U}_y +\mathcal{U}_y^* \mathcal{U}_x$, integrated across the detector surface at position $z_D$ in the far field, \begin{equation} \label{Eq::ClassicalMeasurementSignal} \mathcal{M} \propto \int d^2 \mathbf{r}_\perp \text{Re}\big\{ \mathcal{U}_x^*(\mathbf{r}_\perp, z_D) \mathcal{U}_y(\mathbf{r}_\perp, z_D) \big\}. \end{equation} The result is an effective homodyne detector for $\mathcal{U}_y$, with $\mathcal{U}_x$ playing the role of the local oscillator. Using the solution for $ \mathcal{U}_y(\mathbf{r}_\perp, z_D)$, \erf{paraxsol}, and the properties of the propagator, Eq. (\ref{backprop}), \begin{widetext} \begin{eqnarray} \label{Eq::ClassicalMeasurementSignal2} \mathcal{M} & \propto & \int_{-\infty}^{z_D} dz' \int d^2 \mathbf{r}_\perp' \left[ \int d^2 \mathbf{r}_\perp \, \mathcal{U}_x^* (\mathbf{r}_\perp, z_D) K(\mathbf{r}_\perp-\mathbf{r}_\perp', z_D-z') \right] \text{Re}\big\{ i \chi_{yx}(\mathbf{r}_\perp', z')\big\} \mathcal{U}_x (\mathbf{r}_\perp',z') \nonumber \\ &=& -\int d^3 \mbf{r}' \, \text{Im}\big\{ \chi_{yx}(\mathbf{r}_\perp', z')\big\} |\mathcal{U}_x (\mathbf{r}_\perp',z')|^2. \end{eqnarray} \end{widetext} The measured signal is thus proportional to the local value of the susceptibility component $\text{Im}\left\{ \chi_{yx}(\mathbf{r}_\perp, z)\right\}$ integrated over the dielectric, weighted by the local field intensity $|\mathcal{U}_x (\mathbf{r}_\perp,z)|^2$. For an ensemble of dilute cold atoms at fixed positions $\mbf{r}_i$, the dielectric susceptibility of the gas is \begin{equation} \tensor{\chi}(\mbf{r}) = \sum_i \expects{ \hat{\tensor{\alpha}}\phantom{}^{(i)} } \, \delta^{(3)}(\mbf{r}-\mbf{r}_i), \end{equation} where $\hat{\tensor{\alpha}}\phantom{}^{(i)}$ is the the dynamic polarizability tensor operator for the $i^{th}$ atom. We consider here atoms restricted to a subspace defined by a total (hyperfine) angular momentum $f$. In terms of the hyperfine spin operator $\hat{\mbf{f}}$, the polarizability operator can be decomposed into irreducible tensor components \cite{DeuJes09}, \begin{equation}\label{Eq::IrreducibleDecomp} \hat{\alpha}_{ij} \! = \! \alpha_0 \Big[ C^{(0)} \! +i C^{(1)} \epsilon_{ijk} \hat{f}_k + C^{(2)} \! \Big( \frac{ \hat{f}_i \hat{f}_j +\hat{f}_j \hat{f}_i}{2}- \delta_{ij}\frac {\hat{\mbf{f}}^2}{3} \Big) \Big] \end{equation} where $\alpha_0$ is the characteristic polarizability and $C^{(K)}$ is the coefficient of the irreducible rank-$K$ tensor component. The rank-0 component is a scalar, which does not influence spin and polarization dynamics. The vector (rank-1) component is responsible for the Faraday effect, while the tensor (rank-2) component induces birefringence. For alkali atoms driven on a fine-structure multiplet, $\alpha_0$ and the $C^{(K)}$ coefficients are given in \cite{DeuJes09}. The effect of the tensor component complicates both the collective coupling of the atoms to the probe as well as the internal spin dynamics. In special cases, the deleterious effects of the rank-2 component of the tensor polarizability can be removed via dynamical decoupling~\cite{Koschorreck2010}. More generally, a large bias field removes the rank-2 component of the interaction that couples the collective spin to the polarization of the probe~\cite{NorDeu12}, leaving only internal spin dynamics that can be compensated. We thus retain only the vector component of the off-diagonal element of the dielectric susceptibility, $\chi_{yx}$, which describes a pure Faraday interaction. Substituting $\text{Im} \{\chi_{yx}(\mbf{r}_\perp,z)\} \propto \sum_i \expects{ \hat{f}^{(i)}_z}\delta^{(3)}(\mbf{r}-\mbf{r}_i) $ into \erf{Eq::ClassicalMeasurementSignal2} yields \begin{equation} \label{Eq::SemiclassicalSpinWave} \mathcal{M} \propto \sum_i |\mathcal{U}_x (\mbf{r}_{\perp i}, z_i)|^2 \expect{ \hat{f}^{(i)}_z }. \end{equation} Equation (\ref{Eq::SemiclassicalSpinWave}) is the central result of the semiclassical model. In a plane wave, homogeneous, one-dimensional description, the measured observable is $\mathcal{M} \propto \sum_i \expects{ \hat{f}^{(i)}_z } = \expects{ \hat{F}_z } $, the symmetric collective spin of the ensemble. For paraxial beams, the polarimeter measures an effective {\em spin wave} determined by the inhomogeneous weighting of the atomic spin operators by the local intensity of the beam. The spin wave is stationary because it is coupled to the {\em forward-scattered light}, where the absorbed and emitted modes are the same. Physically, it is this collective observable that radiates indistinguishably into the probe mode and is effectively selected by the homodyne measurement of the polarimeter. \begin{figure} \includegraphics[width=1\hsize]{figure_2} \caption{ Scattered modes for various atomic cloud and beam geometries. The probe laser mode is indicated in blue and the mode scattered by a given dielectric distribution is shown in red. a) An atomic ensemble localized with respect to the probe beam scatters light like a point dipole. b) A ``pancake"-shaped cloud at a fixed $z$-plane radiates nearly perfectly into the probe mode. Extended ``pencil"-shaped clouds can radiate into the probe mode well as in c) or poorly as in d) depending on the geometry of the probe. } \label{Fig::ModeMatching} \end{figure} Further intuition can be gained from the semiclassical model. We recover symmetric atom-light coupling when the field intensity $ |\mathcal{U}_x (\mbf{r}_{\perp}, z)|^2$ is constant over the atomic ensemble. Geometrically, this is achieved when the beam waist, $w_0$, is much larger than the transverse extent of the cloud and the length of the cloud is short compared to twice the Rayleigh range, $z_R = k_0 w_0^2/ 2$. The mean-field radiation pattern of such a cloud described by \erf{paraxsol}, however, has poor overlap with the probe as depicted in Fig. \ref{Fig::ModeMatching}(a). The end result is that the polarimeter detects only a small fraction of the signal photons. On the other hand, perfect mode matching is achieved for atoms confined as a uniform dielectric sheet at a fixed $z$-plane as seen in Fig \ref{Fig::ModeMatching}(b). However, for a finite number of atoms, the realizable OD is low in this configuration. Indeed, a uniform dielectric slab of extent much larger than the beam waist achieves perfect mode matching, but one cannot achieve such an dielectric distribution with high OD using cold atomic gases. An intermediate ``pencil"-shaped geometry is more realistic, allowing for reasonable mode matching while maintaining a high OD, as in Figs. \ref{Fig::ModeMatching}(c-d). In addition to maximizing the signal, we must minimize the sources of noise. There are two fundamental effects: (i) the polarimeter has a finite shot noise resolution; (ii) atoms scatter photons diffusely into all directions (spontaneous emission). The latter is accompanied by optical pumping that can both depolarize the spins and inject noise into the measured spin wave. In order to address these effects, we must turn to the fully quantum theory. \subsection{Quantum theory} \label{Sec::QuantumTheory} Following Refs. \cite{Ham06, VasSor12}, we partition the quantized electric field into paraxial modes and nonparaxial, diffuse modes, \begin{align} \label{Eq::ModeDecomp} \hat{\mathbf{E}}^{(+)}(\mathbf{r}, t) = \hat{\mathbf{E}}_{\rm para}^{(+)}(\mathbf{r}, t) + \hat{\mathbf{E}}_{\rm diff}^{(+)}(\mathbf{r}, t). \end{align} This decomposition is motivated by the geometry we consider -- photon scattering of a paraxial laser beam by an extended atomic ensemble. The mean-field, spatially averaged atomic density, which plays the role of the index of refraction in the classical theory, appears as coherent radiation by a collective atomic observable in the quantum theory. The coupling of collective atomic observables to paraxial modes thus describes the coherent atom-photon light-shift interaction, mediated by the Hermitian part of the atomic polarizability operator. The diffuse modes, in contrast, couple to the density fluctuations in the ensemble due to the discrete atomic positions and thus act locally on each atom \cite{SorSor08}. In the usual Born-Markov approximation, tracing over these modes leads to decoherence and is described by the anti-Hermitian part of the atomic polarizability \cite{DeuJes09}. In this section we first derive a multimode generalization of the Faraday interaction that coherently entangles the atomic ensemble and the paraxial quantum field. Then, we employ a master equation to account for the effects of local decoherence (optical pumping) driven by diffuse scattering. Finally, we present the stochastic master equation describing the conditional collective atomic state given polarimetry measurements of the paraxial field, which we use to analyze spin squeezing in Sec. \ref{Sec::SpinSqueezing}. \subsubsection{Paraxial multimode Faraday interaction} Quantization of paraxial electromagnetic fields was discussed in \cite{DeuGar91}; relevant extensions to the current problem are summarized in Appendix \ref{Appendix::ParaxialQuantization}. We decompose the paraxial field operator into an orthogonal set of tranverse spatial modes, here the Laguerre-Gauss modes $ u_{pl}(\mathbf{r}_\perp,z)$, given in \erf{Eq::LGModes}, which are convenient for cylindrical symmetry. The positive-frequency component of the electric field restricted to the paraxial subspace is, \begin{align} \hat{\mbf{E}}&^{(+)}_{\rm para}(\mbf{r},t) = \nonumber \\ & \sum_{p,l,\alpha=x,y} \sqrt{\frac{2 \pi \hbar \omega_0}{c A}}\, \mbf{e}_\alpha \, \hat{a}_{pl,\alpha}(z,t) \, u_{pl}(\mathbf{r}_\perp,z) e^{i(k_0 z - \omega_0 t)}, \end{align} where the quantization area is chosen as the natural scale of the Gaussian beam, $A =\pi w_0^2/2$. The traveling wave creation/annihilation operators for each transverse mode freely propagate according the Hamiltonian \begin{equation} \hat{H}_{\text{free}} = \sum_{p,l, \alpha} \int dz \; \hat{a}^\dag_{pl, \alpha}(z,t) \left( -i \hbar c \frac{\partial}{\partial z} \right) \hat{a}_{pl, \alpha}(z,t), \end{equation} with solution, $\hat{a}_{pl, \alpha}(z,t) = \hat{a}_{pl, \alpha}(z-ct,0)=\hat{a}_{pl, \alpha}(0,t-z/c),$ and free-field commutation relations, \begin{equation} \left[ \hat{a}_{pl,\alpha}(z,t),\hat{a}^{\dag}_{p'l',\beta}(z',t')\right] =\delta_{p,p'} \delta_{l,l'} \delta_{\alpha, \beta} \, \delta(t-t'-(z-z')/c). \end{equation} We have normalized so that $\hat{a}^{\dag}_{pl,\alpha}(z,t)\hat{a}_{pl,\alpha}(z,t)$ is the local photon flux in transverse mode $pl$ with polarization $\alpha$. For weak excitation (linear atomic response), the interaction Hamiltonian governing the coupling of the quantized paraxial modes is \begin{equation} \label{Eq::ScatteringInteraction} \hat{H}_{\text{int}} = - \sum_i \hat{\mathbf{E}}^{(-)}_{\rm para}(\mbf{r}_i, t) \cdot \hat{\tensor{\alpha}}\phantom{}^{(i)} \cdot \hat{\mathbf{E}}^{(+)}_{\rm para}(\mbf{r}_i, t). \end{equation} As before, the index $i$ is summed over atoms in the ensemble at respective positions $\mathbf{r}_i$. Upon substituting the decomposition of $\hat{\alpha}_{ij}$ into its irreducible components given in Eq. (\ref{Eq::IrreducibleDecomp}), we find scalar (rank-0), vector (rank-1), and tensor (rank-2) contributions to the interaction. We retain only the vector contribution that leads to the Faraday effect, as the scalar contribution does not entangle photons with the atoms and the tensor contribution can in principle be removed \cite{KosMit10}. The Faraday interaction is then, \begin{widetext} \begin{align} \label{Eq::MultimodeFaraday} \hat{H}_{\text{int}} &= - \alpha_0 C^{(1)} \sum_i \hat{\mathbf{E}}^{(+)}_{\rm para}(\mbf{r}_i, t) \times\hat{\mathbf{E}}^{(-)}_{\rm para}(\mbf{r}_i, t) \cdot \hat{ \mbf{f}}^{(i)} \nonumber \\ &= \frac{\hbar \chi}{2} \sum_{ i,p,l, p',l' } \Big[ i u^*_{p'l'}(\mbf{r}_{\perp i}, z_i) u_{pl}(\mbf{r}_{\perp i}, z_i) \, \hat{a}^\dag_{p'l',y}(z_i,t) \hat{a}_{pl,x } (z_i,t) + \text{H.c.} \Big] \hat{f}_z^{(i)}, \end{align} \end{widetext} where \begin{equation} \chi = -C^{(1)} \frac{4 \pi \omega_0}{A c} \alpha_0 = -C^{(1)} \left(\frac{\sigma_0}{A}\right) \left(\frac{\Gamma}{2 \Delta}\right) \end{equation} is the Faraday rotation angle, $\sigma_0 = 3 \lambda^2/(2\pi)$ is the resonant cross-section for unit oscillator strength, $\Gamma$ is the atomic linewidth, and $\Delta$ is the detuning from resonance. For an $S_{1/2}\rightarrow P_J$ transition with $\Delta$ much larger than the excited state hyperfine splitting, $|C^{(1)}| = g_f/3$, where $g_f$ is the Land\'{e} g-factor. We can interpret \erf{Eq::MultimodeFaraday} as a scattering process, whereby an $x$-polarized photon in a given transverse mode, $pl$, is absorbed and a $y$-polarized photon in the mode $p'l'$ is emitted, and vice versa, as mediated by the collective atomic spin wave. Here, we consider an initial macroscopic occupation in the laser probe, again taken to be the fundamental Gaussian TEM$_{00}$ mode with $x$-polarization. In that case the interaction can be linearized by substituting $\hat{a}_{pl,x}(z,t) \rightarrow \sqrt{\dot{N}_L}\delta_{p,0}\delta_{l,0}$, where $\dot{N}_L = A I_0/(\hbar \omega_0)$ is the photon flux of the laser with peak intensity $I_0$. The quantum fluctuations in the field of interest are then represented by the $y$-polarized mode, $\hat{a}_{pl,y}(z,t)$, and the Faraday interaction then takes the form \begin{align}\label{Eq::TheFaradayInteraction} \hat{H}_{\text{int}} \!= \! -i \frac{\hbar \sqrt{\kappa}}{2} \! \sum_{i,p,l} \! \Big[ & \beta^*_{pl}(\mbf{r}_{\perp i}, z_i ) \hat{a}_{pl,y}(z_i,t) \!- \! \mbox{h.c.} \Big] \! \hat{f}_z^{(i)} \end{align} where the local amplitude for scattering from the fundamental (laser) mode $00$ into mode $pl$ is given by \begin{align} \label{Eq::Beta} \beta_{pl}(\mbf{r}) \equiv u^*_{pl}(\mbf{r}) u_{00}(\mbf{r}). \end{align} The interaction has been written in terms of the ``measurement strength" per atom, \begin{align} \label{Eq::MeasurementStrengthperAtom} \kappa = \chi^2 \dot{N}_L = \frac{1}{9f^2} \left( \frac{\sigma_0}{A} \right)\gamma_0, \end{align} which characterizes the rate at which photons are scattered into the paraxial modes, where $\gamma_0 = (\sigma_0 I_0 /\hbar \omega_0) (\Gamma^2/4\Delta^2)$ is the unit-oscillator-strength photon scattering rate at the peak intensity. The Heisenberg equation of motion for a $y$-polarized traveling wave mode interacting with the atomic media in the presence of the probe field is \begin{equation} \left(\frac{\partial}{\partial t} +c \frac{\partial}{\partial z}\right) \hat{a}_{pl,y}(z,t) = \frac{\sqrt{\kappa}}{2} \sum_i \beta_{pl} (\mbf{r}_{\perp i} z_i) \hat{f}_z^{(i)} (t) \delta(z-z_i), \end{equation} whose solution is \begin{align} \hat{a}_{pl,y}(z&,t) =\hat{a}_{pl,y}(0,t-z/c)&\\ &+\frac{\sqrt{\kappa}}{2} \sum_i \beta_{pl} (\mbf{r}_{\perp i} z_i) \hat{f}_z^{(i)} (t-(z-z_i)/c) \Theta(z-z_i),\nonumber \end{align} where $ \Theta(z)$ is the Heaviside step function. Neglecting the time it takes light to propagate across the sample, the mode amplitude at the detector plane, $z_D$, in the far field is \begin{equation} \hat{a}_{pl,y}(z_D,t) =\hat{a}_{pl,y}(0,t-z_D/c)+\frac{\sqrt{\kappa}}{2}\hat{F}_z^{pl}, \end{equation} a form familiar from input-output theory \cite{GardZoller}. The collective atomic spin wave that couples to this paraxial mode is\begin{equation} \hat{F}_z^{pl}= \sum_i \beta_{pl} (\mbf{r}_{\perp i}, z_i) \hat{f}_z^{(i)}. \end{equation} In the balanced polarimeter, the probe mode acts as a local oscillator so that one measures the Stokes vector component associated with the fundamental spatial mode defined by the laser beam, $(\hat{a}^\dag_{00,x}\hat{a}_{00,y}+\hat{a}^\dag_{00,y}\hat{a}_{00,x})/2 \approx \sqrt{\dot{N}/2}\hat{X}_{00}$, where $\hat{X}_{00}= (\hat{a}_{00,y}+\hat{a}^\dag_{00,y})/\sqrt{2}$ is the mode quadrature. The measured quadrature at the detector plane, $z_D$, is thus \begin{equation} \label{Eq::XplOut} \hat{X}_{00}(z_D,t) = \hat{X}_{00}(0,t-z_D/c) +\sqrt{\frac{\kappa}{2}} \hat{F}^{00}_z . \end{equation} Thus, the total polarimeter signal, integrated over a time $T$, is determined by the output operator \begin{equation}\label{Eq::Measurement} \hat{\mathcal{M}} = \int_0^T dt\, \hat{X}_{00}(0,t-z_D/c) + T \sqrt{\frac{\kappa}{2}} \hat{F}^{00}_z \end{equation} where $\hat{F}^{00}_z$ is the fundamental spin wave found in the semiclassical calculation, \erf{Eq::SemiclassicalSpinWave}. The fully quantum theory explicitly includes the additional vacuum noise entering the polarimeter, $\langle \Delta \hat{X}_{00}(0,t)\Delta \hat{X}_{00}(0,t') \rangle=\delta(t-t')/2$, that leads to a shot-noise (SN) variance of the polarimeter signal, in Eq. (\ref{Eq::Measurement}), $\Delta \mathcal{M}_{\rm SN}^2 = T/2$. Of particular interest here is the application to spin squeezing via QND measurement. In this case, the signal we seek to measure arises from different spin-projections associated with the eigenstates of $\hat{F}_z^{00}$. Whereas in magnetometry these shot-to-shot variations are known as ``projection noise" (PN), in the context of creating a spin squeezed state, these variations from the mean value represent the ``signal" one seeks to resolve over the laser shot noise. For the fundamental spin wave measured in the polarimeter, the projection noise variance is \begin{align}\label{Eq::variance} \left( \Delta F^{00}_z \right)^2_{\rm PN} = & \sum_{i} \beta^2_{00}(\mbf{r}_i) \expect{(\Delta \hat{f}_z^{(i)})^2} \\ & + \sum_{i\neq j} \beta_{00}(\mbf{r}_i) \beta_{00}(\mbf{r}_j )\expect{\Delta \hat{f}_z^{(i)}\Delta \hat{f}_z^{(j)}}. \nonumber \end{align} Given an initial spin coherent state polarized orthogonal to $z$, $\expects{\Delta \hat{f}_z^{(i)} \Delta \hat{f}_z^{(j)}}=(f/2) \delta_{ij}$, and thus, \begin{equation}\label{Eq::PNvariance} \left( \Delta F^{00}_z \right)^2_{\rm PN}= \frac{ N^{(2)}_\text{eff} f}{2}. \end{equation} Here we define a set of effective atom numbers \begin{align} \label{Eq::EffectiveAtomNumbers} N^{(K)}_\text{eff} & = \sum_{i} \big[ \beta_{00}(\mbf{r}_i) \big]^K = \sum_{i} |u_{00}(\mbf{r}_i)|^{2K} \\ &\rightarrow \int d^3\mbf{r} \, \eta(\mbf{r}) |u_{00}(\mbf{r}_i)|^{2K}, \end{align} where the sum becomes an integral in the continuum limit. The atomic density distribution $\eta(\mbf{r})$, is normalized so that $\int d^3\mbf{r} \, \eta(\mbf{r}) = N$, the total atom number. The effective atom number $N^{(2)}_\text{eff}$ determines the projection noise contribution to Eq. (\ref{Eq::Measurement}), $\Delta\mathcal{M}_{\rm PN}^2 = \kappa T^2 ( \Delta F_z^{00} )^2 _{\rm PN}/2 =\kappa T^2 N^{(2)}_\text{eff} f/4$. The coupling strength $\xi$ that sets the degree of entanglement one can attain between the atoms and photons is the ratio of the projection noise variance to the shot noise resolution \cite{DeuJes09}. Using Eqs. (\ref{Eq::MeasurementStrengthperAtom}) and (\ref{Eq::PNvariance}) we find \begin{equation}\label{Eq::CouplingStrength} \xi = \frac{\Delta \mathcal{M}_{\rm PN}^2}{\Delta \mathcal{M}_{\rm SN}^2}= \left( \Delta F_z^{00}\right)^2 \kappa T = \mbox{OD}_\text{eff} \, \frac{\gamma_s T}{18f}, \end{equation} where we have defined the effective optical density for the laser mode probing the spin wave on a unit-oscillator-strength transition, \begin{align} \label{Eq::ODeff} \mbox{OD}_\text{eff} = N^{(2)}_\text{eff} \frac{\sigma_0}{A}. \end{align} The key to achieving a large OD$_\text{eff}$ is choosing an atomic and beam geometry that addresses a large number of atoms and maximizes $N_\text{eff}^{(2)}$ while keeping the mode area $A$ small. It should be noted that whereas in the one-dimensional case the optical density is associated with both the coupling strength and the Beer's law attenuation of the probe, in the three-dimensional case different parameters are associated with each of these effects. Because the attenuation coefficient in \erf{Eq::OutputField} is proportional to the local intensity of the field, the total attenuation depends upon the effective atom number $N^{(1)}_\text{eff} $. While \erf{Eq::CouplingStrength} implies an ever increasing coupling strength with integration time $T$, we have neglected so far the decoherence that limits the total useful integration time and the strength of the atom-light interface. In the following section we treat these effects from a first-principles master equation, including spatial variations in the scattering rate which drives local decoherence. \subsubsection{Local decoherence and optical pumping } The discrete random atomic positions are associated with the density fluctuations that give rise to diffuse scattering into 4$\pi$ steradians \cite{SorSor08}. We consider light far detuned from any atomic resonance in a highly transparent regime, and thus we can safely neglect the small attenuation of the laser probe associated with this absorption. The scattering processes, however, cause decoherence of the spin wave due to optical pumping. This local decoherence breaks the collective symmetry of the problem and adds additional noise, which is detected in the polarimeter and competes with squeezing. To treat the decoherence due to diffuse scattering, we employ a master equation, \begin{equation} \frac{d \hat{\rho}}{dt} = -\frac{i}{\hbar} [\hat{H}_{\rm int}, \hat{\rho}] + \frac{d \hat{\rho}}{dt} \Big|_{\rm diff}, \end{equation} where $\hat{H}_{\rm int}$ is the multimode Faraday interaction given in \erf{Eq::TheFaradayInteraction}. The key feature of this equation is that the paraxial modes couple to collective spin waves, while the diffuse scattering couples to localized atoms and induces optical pumping according to \begin{equation} \label{Eq::GenME} \frac{d \hat{\rho}}{dt} \Big|_{\rm diff} = \sum_i \gamma_s(\mathbf{r}_i) \mathcal{D}_i [\hat{\rho}]. \end{equation} The map $\mathcal{D}_i$ acts on the $i^{th}$ atom, proportional to the local scattering rate, \begin{align} \label{Eq::LocalScatRate} \gamma_s(\mathbf{r}_i) = I(\mathbf{r}_i) \frac{\sigma_0}{\hbar \omega} \frac{\Gamma^2}{4 \Delta^2} = \gamma_{0}\beta_{00}(\mathbf{r}_i). \end{align} Here $I(\mathbf{r}_i) = I_0 \beta_{00}(\mbf{r}_i)$ is the local intensity at the position of the atom and $\gamma_{0}$ is the peak scattering rate. We consider here a probe driving an $S_{1/2} \rightarrow P_J$ transition in an alkali atom, with a detuning that is small compared to the ground state hyperfine splitting but large compared to any hyperfine splitting in the excited state. In this case, the light coherently couples substantially only to atoms in a given ground-electronic hyperfine manifold $f$ and the master equation is restricted to this subspace. As shown in Appendix \ref{Appendix::MultimodeJumpOperators}, with an $x$-polarized probe and applying a large bias magnetic field along the $z$-direction, the local decoherence in the master equation due to optical pumping is given by the map \begin{align} \label{Eq::DiffuseME} \mathcal{D}_i [\hat{\rho}] = \! -\frac{2}{9}\hat{\rho} + \frac{g_f^2}{9} \Big[ \hat{f}^{(i)}_z\hat{ \rho} \hat{f}^{(i)}_z + \frac{1}{2} \big( \hat{f}^{(i)}_x \hat{\rho} \hat{f}^{(i)}_x + \hat{f}^{(i)}_y \hat{\rho} \hat{f}^{(i)}_y \big) \Big]. \end{align} The first term on the right-hand side of \erf{Eq::DiffuseME} describes the decay of correlations due to optical pumping, while the second term represents a feeding due to ``transfer of coherences'' that can reduce this decay rate \cite{CohenTannoudji}. Note that for $f>1/2$, this master equation is not trace preserving, since atoms can be optically pumped to the other ground hyperfine manifold where they are lost to any further measurement. Given the master equation, we can find the effect of diffuse scattering on atomic correlations. Consider a inhomogeneous collective operator of the form $\hat{X} = \sum_i \beta_{00}(\mbf{r}_i) \hat{x}^{(i)}$. Because $\hat{X}$ is a weighted sum over single atom operators, the equation of motion for its expectation value depends upon the evolution of the single atom density operator, $\hat{\rho}^{(i)}$. By summing over a single index $i$ in Eq. (\ref{Eq::GenME}) we obtain \begin{align}\label{eq::rhoi} \frac{d\hat{\rho}^{(i)}}{dt}\Big|_{\rm diff}=\gamma_s(\mbf{r}_i) \mathcal{D}_i[\hat{\rho}^{(i)}], \end{align} from which the evolution of $\expects{\hat{X}}$ is given by \begin{align}\label{Eq::1stOrderEvol} \frac{d}{dt}\expects{\hat{X}}\Big|_{\rm diff} & =\sum_i\ \beta_{00}(\mbf{r}_i) \mbox{Tr} \bigg[ \hat{x}^{(i)}\frac{d\hat{\rho}^{(i)}}{dt}\Big|_{\rm diff}\bigg] \nonumber \\ &= \sum_i \gamma_s(\mbf{r}_i) \beta_{00}(\mbf{r}_i) \expect{\mathcal{D}_i[\hat{x}^{(i)}]}. \end{align} For inhomogeneous collective operators that depend on pairs of atoms, \begin{align} \hat{O} = \sum_{i \neq j} \beta_{00}(\mbf{r}_i) \beta_{00}(\mbf{r}_j) \hat{x}^{(i)} \hat{y}^{(j)}, \end{align} we require the joint density operator of the $i^{th}$ and $j^{th}$ atoms, $\hat{\rho}^{(i,j)}$, with equation of motion \begin{align}\label{Eq::rhoij} \frac{d}{dt}\hat{\rho}^{(i,j)}\Big|_{\rm diff} =\gamma_s(\mbf{r}_i)\mathcal{D}_i[\hat{\rho}^{(i,j)}]+ \gamma_s(\mbf{r}_j) \mathcal{D}_j[\hat{\rho}^{(i,j)}]. \end{align} The evolution of $\expects{\hat{O}}$ due to diffuse scattering is then \begin{align} \label{Eq::2ObservableEOM} \frac{d}{dt} \expects{\hat{O}} \Big|_{\rm diff} & = \sum_{i \neq j}\Big\{ \gamma_s(\mbf{r}_i) \beta_{00}(\mbf{r}_i) \beta_{00}(\mbf{r}_j) \expect{ \mathcal{D}_i [\hat{x}^{(i)}] \hat{y}^{(j)}} \nonumber \\ &+ \gamma_s(\mbf{r}_j) \beta_{00}(\mbf{r}_i) \beta_{00}(\mbf{r}_j) \expect{ \hat{x}^{(i)} \mathcal{D}_j[ \hat{y}^{(j)}] } \Big\}. \end{align} The degree of squeezing that one can ultimately produce is determined by a balance between QND measurement backaction on the spin wave mediated by the collective radiation and the damage to that observable caused by diffuse scattering. To properly treat this we must include the effects of measurement on the atoms, as discussed in the next section. \subsubsection{The conditional stochastic master equation} \label{Sec::HomodyneSME} The Faraday Hamiltonian, \erf{Eq::TheFaradayInteraction}, is an entangling interaction between the atomic spin waves and the paraxial modes of the field. When the light is measured in the polarimeter, quantum backaction leads to stochastic evolution of the atomic state, conditioned on the measurement result. A complete description of the dynamics is then described by a stochastic master equation (SME), with decoherence from unmeasured light and squeezing due to information gained from the continuous measurement record. In a balanced polarimeter, the measurement signal is proportional to the interference of the probe and scattered fields integrated over the detector faces, as in \erf{Eq::ClassicalMeasurementSignal}. Due to the orthogonality of the spatial modes, \erf{Eq::TransverseOrthogonality}, such a measurement selects only paraxial light that is scattered into the mode of the probe, $u_{00}$. The result is a continuous measurement of the quadrature $\hat{X}_{00}$. We derive the SME for the atoms following \cite{JacSte06, WieMilBook}, with details presented in Appendix \ref{Appendix::SMEDerivation}. Measurement of the quadrature $\hat{X}_{00}$ by the homodyne polarimeter generates a differential stochastic measurement record \begin{equation} dy_{00} =\expect{\hat{F}_z^{00}} dt + \frac{1}{ \sqrt{\kappa}}dW, \end{equation} where $dW$ is a Weiner interval in the It\={o} calculus and $\kappa$ is the measurement strength given in \erf{Eq::MeasurementStrengthperAtom}. Assuming unit measurement efficiency, the evolution of the ensemble conditioned upon the measurement record $y_{00}$ is given by \begin{align} \label{ApEq::HomodyneSMEtext} d \hat{\rho}&= \sqrt{ \frac{\kappa}{4} } \mathcal{H}_{00}[\hat{\rho}] \, dW + \frac{\kappa}{4} \sum_{p,l} \mathcal{L}_{pl}[\hat{\rho}] \, dt. \end{align} The effects of measurement backaction on the ensemble are taken into account by the superoperator $\mathcal{H}_{00}[\hat{\rho}] $, where \begin{align} \label{Eq::HSuperoperator} \mathcal{H}_{pl}[\hat{\rho}] = \hat{F}^{pl}_z \hat{\rho} + \hat{\rho} \hat{F}^{pl \dagger}_z - \Tr{ \big(\hat{F}^{pl}_z + \hat{F}^{pl\dagger}_z \big) \hat{\rho} } \hat{\rho}. \end{align} The Lindblad dissipator, \begin{align} \label{Eq::LSuperoperator} \mathcal{L}_{pl}[\hat{\rho}] = \hat{F}_z^{pl} \hat{\rho} \hat{F}_z^{pl\dagger} - \frac{1}{2} \hat{F}_z^{pl\dagger} \hat{F}_z^{pl} \hat{\rho} - \frac{1}{2} \hat{\rho} \hat{F}_z^{pl\dagger} \hat{F}_z^{pl}, \end{align} describes the effect on the atomic ensemble arising from collective radiation into all paraxial modes of the field. Including local decoherence from diffuse scattering, \erf{Eq::DiffuseME}, the full stochastic master equation for homodyne polarimetry measurements of the $00$-mode is \begin{align} \label{Eq::HomodyneSME} d \hat{\rho} =& - \frac{i}{\hbar} [\hat{H}, \hat{\rho}]dt + \sqrt{ \frac{\kappa}{4} } \mathcal{H}_{00}[\hat{\rho}] \, dW \\ &+ \frac{\kappa}{4} \sum_{p,l} \mathcal{L}_{pl}[ \hat{\rho}]\, dt + \sum_i \gamma_s(\mathbf{r}_i) \mathcal{D}_i [\hat{\rho}] \, dt. \nonumber \end{align} This SME is a complete description of the evolution of the collective atomic state, accounting for the three-dimensional nature of the atom-photon modes, decoherence, and measurement backaction. We see that through its interaction with the probe, the atomic ensemble undergoes an additional form of \emph{collective} decoherence, \erf{Eq::LSuperoperator}, corresponding to light radiated into paraxial modes $pl \neq 00$ that ultimately goes unmeasured. Thus we have arrived at the same conclusion as in Ref. \cite{DuaZol02}. That is, decoherence arises through two distinct processes - first, the inherent mode-mismatch that gives rise to collectively scattered light in spatial modes other than the probe mode and second, the diffuse scattering of photons that acts locally on atoms in the ensemble. \section{QND squeezing of spin waves} \label{Sec::SpinSqueezing} \subsection{Quantifying squeezing of the spin waves} One typically quantifies the amount of squeezing created in a QND measurement according to the Wineland squeezing parameter~\cite{Wineland94}, \begin{equation} \label{Eq::StandardParam} \zeta =\left(\frac{\Delta \phi}{\Delta \phi_{\text{SCS}}}\right)^2, \end{equation} where $\Delta \phi$ is the projection-noise limited resolution when measuring an angle of rotation for a generic spin $J$ of the given input state, and $\Delta \phi_{\text{SCS}}$ is the corresponding resolution when the input is a spin coherent state (SCS). For a mean value $J_\parallel \equiv |\expects{\hat{\mbf{J}}}|$, and variance $\Delta J_\perp^2$ orthogonal to the mean, the projection-noise resolution is $\Delta \phi =\Delta J_\perp/J_\parallel$. With $\Delta \phi_{\rm SCS} =1/\sqrt{2J}$, the Wineland squeezing parameter is then \begin{equation} \zeta = 2J\frac{\Delta J_\perp^2}{J_\parallel^2}. \label{simple_squeezing} \end{equation} For the spin waves of the inhomogeneous ensemble under consideration here, we must tie the squeezing parameter directly to the measured quantities. For an initial mean spin polarization along $x$ and a small rotation around $y,$ the polarimeter signal will be determined by the mean spin wave component $\expects{\hat{F}_x^{00}} = \sum_i \beta_{00} (\mbf{r}_i) \expects{\hat{f}^{(i)}_x}$, and the projection noise contribution to the resolution of the measurement will be given by $\Delta F^{00}_z$, defined in Eq. (\ref{Eq::variance}). The projection-noise limited resolution of this rotation is therefore $\Delta \phi^{00} = \Delta F_z^{00}/\expects{\hat{F}_x^{00}}$. Furthermore, given a SCS initially polarized along x, the mean spin of interest is $\expects{\hat{F}_x^{00}}_{\text{SCS}}= N^{(1)}_\text{eff} f$ where the effective atom number contributing to this signal is given in \erf{Eq::EffectiveAtomNumbers}, while the projection noise in the spin coherent state is $( \Delta F_z^{00})^2_{\text{SCS}}=N^{(2)}_\text{eff} f/2$. The projection noise limited resolution for a SCS preparation is thus, $( \Delta \phi^{00}_\text{SCS})^2=N_\text{eff}^{(2)}/[2f (N^{(1)}_\text{eff} )^2 ]$, and will depend on the shape of the atomic cloud and beam geometry. Putting this together, we define the squeezing parameter for the measured spin wave to be \begin{equation} \label{Eq::SqueezingParam} \zeta \equiv \left(\frac{\Delta \phi^{00}}{\Delta \phi^{00}_{\text{SCS}}}\right)^2 = 2f \frac{ \big(N^{(1)}_\text{eff} \big)^2 }{N_\text{eff}^{(2)}} \frac{\big(\Delta F_z^{00}\big)^2}{\expect{\hat{F}_x^{00}}^2}. \end{equation} This parameter quantifies the degree of ``quantum backaction," on a spin coherent state, accounting for the change in projection noise due to QND measurement as well as the damage done to both the mean spin polarization and variance due to optical pumping. In a real-world metrological application such as an optically probed atomic magnetometer~\cite{BudRom07, SewMit12}, spin rotations are measured by passing the probe through the atom sample and measuring the resulting Faraday rotation in a polarimeter. In addition to spin projection noise, the measurement resolution is then subject also to ``technical noise," including probe shot noise, detector electronic noise, and atom number fluctuations. Under those circumstances, optimizing the squeezing parameter as defined in Eq. (\ref{Eq::SqueezingParam}) is distinct from optimizing the magnetometer sensitivity. \subsection{The dynamical evolution of squeezing} To determine the squeezing as function of time, we employ the SME in \erf{Eq::HomodyneSME} to track $( \Delta{F}_z^{00})^2$ and $\expects{\hat{F}_x^{00}}$. For ensembles with large numbers of atoms, we can work in the central-limit approximation where fluctuations in the spin waves are treated as Gaussian random variables~\cite{PolzikRMP, VasSor12}. Following \cite{JacSte06}, the SME then couples solely means and covariances. The moments of the fundamental spin wave that characterize the spin squeezing parameter then evolve according to \begin{subequations} \label{Eq::FullSpinWaveEOM} \begin{align} d\big( \Delta F_z^{00} \big)^2 =& -\kappa \big[ \left(\Delta F_z^{00} \right)^2 \big]^2 dt +\big( \Delta F_z^{00} \big)^2\Big|_{\rm diff}dt , \label{Eq::Var00Evolution}\\ d \expect{\hat{F}^{00}_x} = & \sqrt{ \frac{\kappa}{4} } \big \langle\mathcal{H}_{00}\big[\hat{F}^{00}_x\big]\big \rangle \, dW \label{Eq::Mean00Evolution} \\ & +\sum_{p,l}\frac{\kappa}{4} \big \langle \mathcal{L}_{pl}\big[ \hat{F}^{00}_x \big] \big \rangle dt +\expect{\hat{F}^{00}_x} \Big|_{\rm diff} dt. \nonumber \end{align} \end{subequations} Because we assume the fundamental mode is measured with unit efficiency, diffuse scattering by local spontaneous emission is the only process contributing to the decoherence of the variance $( \Delta F_z^{00} )^2 $. Collective radiation into other transverse modes commutes with $\hat{F}_z^{00}$ and does not contribute to any decay or noise injection into the fundamental variance. In contrast, the mean spin $\expects{\hat{F}^{00}_x}$ decoheres due to both diffuse scattering and collective scattering into other unmeasured paraxial modes. It also evolves stochastically due to the continuous measurement of $\hat{F}_z^{00}$. However, the contributions to the dynamics from both collective scattering and continuous measurement are small in comparison to diffuse scattering and can be neglected when the radiation pattern of the cloud is well matched to that of the probe. We consider the moment evolution, \erf{Eq::FullSpinWaveEOM}, with the initial condition that the ensemble is in a SCS polarized along $x$. The initial mean spin and variance are $\expects{ \hat{F}_{x}^{00} (t_0) }= N_{\rm eff}^{(1)} f$ and $( \Delta F_z^{00} (t_0) )^2 = N_{\rm eff}^{(2)}f/2$ respectively. Along with the cross-sectional area of the probe laser, $N_{\rm eff}^{(2)}$ specifies the effective optical density, OD$_{\rm eff}$ defined in \erf{Eq::ODeff}. The OD$_{\rm eff}$ is the critical geometric parameter for determining how the atomic density distribution influences collective scattering into the probe mode and ultimately leads to spin squeezing. Both of these effective atom numbers are determined solely by the cloud shape and beam geometry, and can be found from the semiclassical model in Sec. \ref{Sec::SemiclassicalTheory}. For times short compared to the photon scattering rate, where decoherence is negligible, the mean spin is essentially constant and the spin variance is affected only by measurement backaction. The solution to \erf{Eq::FullSpinWaveEOM} takes the familiar form \cite{PolzikRMP} \begin{align} \big( \Delta F_z^{00} (t) \big)^2 & = \frac{\big( \Delta F_z^{00} (t_0) \big)^2 }{1+ \big( \Delta F_z^{00} (t_0) \big)^2 \,\kappa t} = \frac{ \big( \Delta F_z^{00} (t_0) \big)^2 }{1+\xi} \\ & \Rightarrow \zeta = \frac{1}{1+\xi}, \label{Eq::ShortTimeSolution} \end{align} where $\xi$ is the integrated coupling strength in Eq. (\ref{Eq::CouplingStrength}). In \erf{Eq::ShortTimeSolution}, the squeezing parameter decreases as OD$_\text{eff}^{-1}$ for $\xi\gg1$. For longer times, decoherence due to diffuse photon scattering must be included. The mean spin will depolarize according to Eq. (\ref{Eq::1stOrderEvol}), \begin{equation}\label{Eq::MeanSpinArb} \frac{d}{dt} \expect{\hat{F}^{00}_x} = \sum_i \gamma_s(\mbf{r}_i)\beta_{00}(\mbf{r}_i) \big\langle \mathcal{D}_i \big[\hat{f}_x^{(i)}\big] \big\rangle. \end{equation} The variance involves both single atom and pairwise atomic correlations, \begin{align} \label{Eq::VarianceDecompositionApprox} \frac{d}{dt}\big( \Delta F_z^{00} \big)^2 = & \sum_i \big[ \beta_{00}(\mbf{r}_i) \big]^2 \frac{d}{dt}\big\langle(\Delta \hat{f}_z^{(i)})^2\big\rangle \\ & + \sum_{i\neq j} \beta_{00}(\mbf{r}_i) \beta_{00}(\mbf{r}_j) \frac{d}{dt}\big\langle \Delta \hat{f}_z^{(i)}\Delta \hat{f}_z^{(j)} \big\rangle, \nonumber \end{align} where the first term is the spin projection noise of the uncorrelated spins and the second term contains the correlations that generate spin squeezing. Following Eqs. (\ref{Eq::1stOrderEvol}) and (\ref{Eq::2ObservableEOM}), these correlation functions decay due to diffuse scattering according to \begin{subequations} \label{Eq::Correlations} \begin{align} & \frac{d}{dt}\sum_{i} \big\langle (\Delta \hat{f}_z^{(i)})^2 \big\rangle \big|_{\rm diff} = \nonumber \\ & \quad \sum_{i} \gamma_s(\mbf{r}_i) \Big\{ \big\langle\mathcal{D}_i \big[ \hat{f}_z^{(i)2} \big] \big\rangle -2 \big\langle \mathcal{D}_i \big[ \hat{f}_z^{(i)} \big] \big\rangle \big\langle \hat{f}_z^{(i)} \big\rangle \Big\} \label{eq::NoiseTerm2}, \\ & \frac{d}{dt}\sum_{i\neq j} \big\langle \Delta \hat{f}_z^{(i)}\Delta \hat{f}_z^{(j)} \big\rangle \big|_{\rm diff} = \nonumber \\ & \! \sum_{i\neq j} \! \Big\{ \! \gamma_s(\mbf{r}_i) \big\langle \Delta\mathcal{D}_i \big[ \hat{f}_z^{(i)} \big]\Delta \hat{f}_z^{(j)}\big\rangle \! +\! \gamma_s(\mbf{r}_j) \big\langle \Delta \hat{f}_z^{(i)}\Delta \mathcal{D}_j \big[\hat{f}_z^{(j)}\big] \big\rangle \! \Big\} \label{eq::correlationDecay2}. \end{align} \end{subequations} The local decoherence acts via the map $\mathcal{D}_i$ in \erf{Eq::DiffuseME}. \subsubsection{Spin-1/2 ensembles} We first restrict our attention to ensembles of spin $f=1/2$ atoms to focus on spatial effects without the complications that arise for ensembles with larger-spin. Using the fact that the local scattering rate is proportional to the probe intensity, $\gamma_s(\mathbf{r}) = \gamma_{0}\beta_{00}(\mathbf{r})$, the mean spin evolution of \erf{Eq::Mean00Evolution} is \begin{equation} \label{Eq::Mean00GammaBeta} \frac{d}{dt} \expect{\hat{F}^{00}_x} = -\frac{\gamma_{0}}{3} \sum_i \big[ \beta_{00}(\mbf{r}_i) \big]^2 \big\langle \hat{f}_x^{(i)} \big\rangle. \end{equation} The local decoherence does not respect the orthogonality of the transverse paraxial modes and we will see that the diffuse scattering acts to couple the fundamental spin wave to higher order spin waves. Because the transverse modes are orthogonal in a plane at a fixed $z$, we can derive a set of coupled equations by decomposing products of the spatial weighting coefficients, \erf{Eq::Beta}, in the basis of mode functions as follows: \begin{equation} \big[ \beta_{00}(\mbf{r}_\perp, z) \big]^2 =|u_{00}(\mathbf{r}_{\perp}, z)|^4 = \sum_{p,l} c^{00}_{pl}(z) \beta_{pl}(\mathbf{r}_\perp, z) , \end{equation} with $z$-dependent projection coefficients, \begin{equation} c^{00}_{pl}(z) \equiv \frac{1}{A} \int d^2 \mathbf{r}_\perp \big[ u_{00}(\mathbf{r}_\perp, z) \big]^2 u^*_{00}(\mathbf{r}_\perp, z) u_{pl}(\mathbf{r}_\perp, z). \end{equation} Performing the sum over atoms within each coarse-grained slice, it follows that \erf{Eq::Mean00GammaBeta} can be expressed as \begin{equation} \label{Eq::MeanZsliced} \frac{d}{dt} \big\langle \hat{F}^{00}_{x} \big\rangle = - \frac{ \gamma_{0}}{3} \sum_{k, p,l} c^{00}_{pl} (z_k) \big\langle \hat{F}_x^{pl}(z_k) \big\rangle, \end{equation} where we have defined spin waves in coarse-grained slices of thickness $\delta z$ around $z_k$, \begin{equation} \label{Eq::CoarseGrainedSpinWave} \hat{F}^{pl}_z (z_k) \equiv \sum_{i_k} \beta_{pl}(\mbf{r}_{\perp {i_k}}, z_{i_k}) \hat{f}_z^{(i_k)}, \end{equation} with the index $i_k$ labeling atoms in the slice $z_k$. The total spin wave for a given transverse mode is $\hat{F}^{pl}_z = \sum_k \hat{F}^{pl}_z (z_k)$. The mean spin in the fundamental mode $\expects{\hat{F}_x^{00}}$ is thus coupled to other spin waves $\expect{\hat{F}_x^{pl}}$ within each slice $z_k$. Generally, the expected value of the $pl$-spin wave in the slice $z_k$ evolves according to \begin{align} \label{Eq::MeanZsliced} \frac{d}{dt} \expect{\hat{F}^{pl}_{x}(z_k)} = - \frac{ \gamma_{0}}{3} \sum_{p',l'} c^{pl}_{p'l'} (z_k) \expect{\hat{F}_x^{p'l'}(z_k)}, \end{align} with projection coefficients $c^{pl}_{p'l'} (z_k)$ given in \erf{Eq::ProjCoeff}. Details of this derivation are found in Appendix \ref{Appendix::SpinCovariances}. The initial conditions, \erf{Eq::meanSlice}, account for the matching between the probe mode and cloud geometry. By projecting onto the spin waves, we obtain a hierarchy of coupled equations. Numerically, we truncate once the desired convergence is achieved. The effect of diffuse scattering on the evolution of the collective spin variance follows in an analogous manner. For spin-1/2, $\expects{\Delta \hat{f}^2_z }=1/4$ for all atoms. The map for local decoherence, $\Delta \mathcal{D}_i\big[\hat{f}^{(i)}_z\big] = -2 \Delta \hat{f}^{(i)}_z/9$, corresponds to decay of spin-spin correlations with no feeding of coherences. The evolution of the fundamental spin wave variance, \erf{Eq::Var00Evolution}, simplifies to \begin{widetext} \begin{align} \label{Eq::SpinHalfVariance} & \frac{d}{dt} \big( \Delta F_z^{00} \big)^2 = -\kappa \Big[ \big( \Delta F_z^{00} \big)^2 \Big]^2 -\frac{2\gamma_{0}}{9} \sum_{i,j} \big[ \beta_{00}(\mathbf{r}_i) +\beta_{00}(\mathbf{r}_j) \big] \beta_{00}(\mbf{r}_i)\beta_{00}(\mbf{r}_j) \expect{\Delta \hat{f}_z^{(i)}\Delta \hat{f}_z^{(j)}} + \frac{\gamma_{0}}{9} \sum_i \big[ \beta_{00}(\mbf{r}_i) \big]^3, \end{align} \end{widetext} where again we have used \erf{Eq::LocalScatRate}. The first term describes squeezing of the variance due to measurement backaction, the second represents decay of correlations due to diffuse scattering, and the third is the noise injected into the variance from spin flips (optical pumping). Following the same procedure as above, the decay terms are projected onto higher order spin waves, \begin{align} \label{Eq::ProjectedSpin1/2VarianceEOM} & \frac{d}{dt} \big( \Delta F_z^{00} \big)^2 = -\kappa \Big[ \big( \Delta F_z^{00} \big)^2 \Big]^2 \\ & -\frac{4 \gamma_0}{9} \sum_{p,l} \sum_{k,k'} c^{00}_{pl}(z_k) \big\langle \Delta \hat{F}_z^{00}(z_k) \Delta \hat{F}_z^{pl}(z_{k'})\big\rangle + \frac{ \gamma_{0}}{9}N^{(3)}_\text{eff}. \nonumber \end{align} Here, $N^{(3)}_\text{eff}$ is the effective atom number governing the injection of noise through optical pumping, defined in \erf{Eq::EffectiveAtomNumbers}. Equation (\ref{Eq::ProjectedSpin1/2VarianceEOM}) is a covariance description of the dynamics, similar to that commonly employed for spin squeezing \cite{MadMol04}, but which also accounts for local decoherence from first principles. To solve for the fundamental variance, we must track the evolution of the covariances between coarse-grained slices and between transverse modes, $\langle \Delta \hat{F}_z^{pl}(z_k) \Delta \hat{F}_z^{p'l'}(z_{k'})\rangle = \expects{\hat{F}_z^{pl}(z_k)\, \hat{F}_z^{p'l'}(z_{k'}) } -\expects{\hat{F}_z^{pl}(z_k)}\expects{ \hat{F}_z^{p'l'}(z_{k'}) }$. Equations of motion for these covariances follow readily from the SME. A detailed derivation is given in Appendix \ref{Appendix::SpinCovariances}. \subsubsection{Spin-f alkali atom ensembles} \label{Sec::SpinfEnsembles} The constituent atoms in many spin squeezing experiments are alkali metal atoms whose ground state structure is more complex than spin-1/2. For example, in $^{133}$Cs, the ground electronic subspace is defined by two hyperfine manifolds with total spin angular momentum $f = \{3,4\}$. Owing to the large ground-state hyperfine splitting (9.2 GHz in $^{133}$Cs), a single hyperfine manifold $f$ is addressed by the coherent interaction with the probe laser. Though ensembles of higher spin atoms can be squeezed by the same QND measurement process, spin size affects both the coherent squeezing dynamics and decoherence. Recall that the strength of the Faraday interaction is quantified by the coupling strength $\xi$, Eq. (\ref{Eq::CouplingStrength}). Because $\xi\propto 1/f^2$, the atom-light coupling decreases with increasing spin size. This decreased coupling strength is partially offset by an increased robustness to the effects of optical pumping. When $f>1/2$, optical pumping events can be broadly divided into two categories: (i) ``loss" that occurs when an atom is pumped from the $f$ manifold into the other ground hyperfine manifold and (ii) ``spin flips" that leave the atom in the $f$ manifold. Because atoms lost to the other ground manifold are no longer resonant with the probe, loss events decrease the mean spin $\langle\hat{F}_x^{00}\rangle$, though they contribute no excess noise to $(\Delta F_z^{00})^2$. ``Spin flips" are responsible for both a decrease in $\langle\hat{F}_x^{00}\rangle$ and a noise injection into $(\Delta F_z^{00})^2$. For the SCS preparation, the deleterious effects of spin flips are mitigated by ``transfers of coherence" between pairs of magnetic sublevels that reduce the rate of decay of correlations~\cite{CohenTannoudji}. While the interplay between these effects is complex, the rate of spin flips remains a good indicator of an ensemble's robustness to optical pumping. For an ensemble of spin-$f$ alkalis prepared in a SCS, the spin flip rate is $\gamma_s(\mathbf{r})/(12f)$, thus decreasing for larger spin size. Due to these decoherence processes, the dynamics of the squeezing parameter is substantially more complicated for larger spin atoms. For spin-$f$, we obtain the evolution of the mean value of a spin wave in slice $z_k$ by projecting onto the different spin waves in a manner analogous to \erf{Eq::MeanZsliced}, \begin{align} \label{Eq::MeanSpins} \frac{d}{dt} & \expect{\hat{F}^{pl}_{x}(z_k)} = - \frac{2 \gamma_{0}}{9} \sum_{p',l'} c^{pl}_{p'l'} (z_k) \expect{\hat{F}_x^{p'l'}(z_k)} \\ & \quad + \frac{g_F^2 \gamma_0}{9} \sum_{p',l'} \sum_{i_k} c^{pl}_{p'l'} (z_k) \beta_{p'l'} (\mathbf{r}_{\perp i_k},z_k )\mathcal{C}_{i_k}[\hat{f}_x^{(i_k)}]. \nonumber \end{align} Here, we have defined a local superoperator that arises solely from the ``feeding" terms in the master equation: \begin{align} \mathcal{C}_i[\hat{X}] \equiv \hat{f}_z^{(i)}\hat{X}\hat{f}_z^{(i)} + \tfrac{1}{2} \big( \hat{f}_x^{(i)}\hat{X}\hat{f}_x^{(i)} + \hat{f}_y^{(i)} \hat{X}\hat{f}_y^{(i)} \big). \end{align} Similarly, we find equations of motion for the fundamental spin wave variance, \begin{align} \label{Eq::GenCovarianceEvolution} & \frac{d}{dt} \big( \Delta F_z^{00} \big)^2 =- \kappa \big[\left(\Delta F_z^{00} \right)^2\big]^2 \\ & -\frac{4\gamma_{0}}{9} \sum_{p,l} \sum_{k',k}c^{00}_{pl}(z_k)\big\langle \Delta \hat{F}_{z}^{00}(z_{k'}) \Delta \hat{F}_{z}^{pl}(z_k) \big\rangle \nonumber \\ & + \frac{g_f^2 \gamma_{0}}{9} \sum_{p,l}\sum_{k',k,i_k} c^{00}_{pl}(z_k)\beta_{pl}(\mathbf{r}_{i_k}) \Big\{ \big\langle \Delta \hat{F}_{z}^{00}(z_{k'}) \Delta \mathcal{C}_{i_k}[ \hat{f}_{z}^{(i_k)} ] \big\rangle \nonumber \\ &\quad \quad \quad +\big\langle \Delta \mathcal{C}_{i_k}[ \hat{f}_{z}^{(i_k)} ] \Delta \hat{F}_{z}^{00}(z_{k'}) \big\rangle \Big\} \nonumber \\ &+\gamma_{0}\sum_{k,i_k} \big[ \beta_{00}(\mathbf{r}_{i_k}) \big]^3 \bigg\{ \frac{2}{9} \expect{(\hat{f}_z^{(i_k)})^2} \nonumber \\ & \quad \quad \quad + \frac{g_f^2}{9} \left( \big\langle \mathcal{C}_{i_k}[\hat{f}_z^{(i_k)2}]\big\rangle-\big\langle \big\{\hat{f}_z^{(i_k)},\:\mathcal{C}_{i_k}[\hat{f}_z^{(i_k)} ] \big\}_+\big\rangle \right) \bigg\} , \nonumber \end{align} where $\{ \hat{X}, \hat{Y}\}_+$ denotes the anti-commutator. As for the case of spin-1/2, we have an infinite hierarchy of equations that couple spin wave operators in the different $z_k$-slices. In general, the feeding terms in \erf{Eq::GenCovarianceEvolution} couple to covariances outside the set $\expects{\Delta\hat{F}_z^{pl}(z_k) \Delta\hat{F}_z^{p'l'}(z_{k'})}$. This expands considerably the number of equations that must be solved to reach convergence. Solving these equations, furthermore, requires different methods than the spin-1/2 case. A detailed treatment of the spin-$f$ case will be provided in future work. \subsection{Results} Using our formalism we can calculate the dynamics of QND measurement and the peak achievable squeezing in the presence of decoherence. We now consider the fundamental effects of geometry and the optimization of experimentally relevant quantities to maximize spin squeezing. Most of our results are shown for the simplest case of spin-1/2 atoms in order to focus on the effects of spatial modes and spin waves. We also consider some preliminary calculations for spin-$f>1/2$; more complete studies will be presented elsewhere. \subsubsection{Geometric effects of local decoherence for a fixed rate of squeezing}\label{Sec::GeomEffects} The geometry of the atom-laser system plays two distinct roles in determining the amount of achievable squeezing. First, OD$_{\rm eff} \propto N_{\rm eff}^{(2)}$, Eq. (\ref{Eq::ODeff}), is a purely geometrical quantity, derivable from the semiclassical model (see also \cite{MulPol05}). OD$_{\rm eff}$ sets the measurement strength, $\xi$, that characterizes the amount of light that is collectively scattered into the spatial mode of the probe. Second, because of the inhomogeneous intensity profile of the laser mode, the rate of diffuse photon scattering that causes the local decoherence, and ultimately caps the amount of squeezing that is generated, varies across the cloud. Further complications arise from the fact that optical pumping both injects noise into the spin wave variance and causes a decay of the mean spin. To gain physical insight, in this section we fix the OD$_\text{eff}$ as we vary the geometry in order to isolate the effects of local decoherence as they relate specifically to the squeezing parameter, \erf{Eq::SqueezingParam}. For simulations, we choose the ensemble to be a cylindrically symmetric Gaussian cloud with average density \begin{align} \label{Eq::AtomicDistribution} \eta(\mathbf{r}) = \eta_0 \exp \left( - 2\frac{\rho^2}{\sigma_\perp^2} - 2\frac{z^2}{\sigma_z^2} \right), \end{align} where $\sigma_\perp^2$ and $\sigma_z^2$ are the transverse and longitudinal $1/e^2$ variances, $\eta_0$ is the peak density, and $\int d^3 \mathbf{r} \eta(\mathbf{r}) = N$, the total atom number. To characterize the geometry of the atomic distribution we use the \emph{aspect ratio}, defined as AR $\equiv \sigma_z/\sigma_\perp$. A longitudinally-extended, pencil-shaped cloud commonly employed in cold, dipole-trapped atomic ensemble experiments has an AR $\gg 1$; a pancake-shaped cloud that is much wider than it is long has an AR $\ll 1$. Note, we vary $\eta_0$ as a function of the cloud geometry such that OD$_\text{eff} \propto N_\text{eff}^{(2)}/A$ remains constant. To find the peak squeezing, we perform numerical simulations by integrating the evolution of the collective mean spin and variance, Eqs. (\ref{Eq::Mean00GammaBeta}) and (\ref{Eq::SpinHalfVariance}), and then calculating the spin squeezing as a function of time. Figure \ref{Fig::SpinHalfCompareSqueezing} shows the resulting spin squeezing for different cloud geometries for a fixed beam waist, $w_0 = 20$ $\mu$m. The effective optical density is held constant, OD$_\text{eff} = 50$, which guarantees identical squeezing in the absence of decoherence for any geometry. Fig. \ref{Fig::SpinHalfCompareSqueezing}(a) shows the peak squeezing as a function of the AR. An increase in peak squeezing accompanies an increasing aspect ratio, indicating that decoherence is less detrimental to longitudinally extended clouds. The dynamics of the squeezing parameter are plotted in Fig. \ref{Fig::SpinHalfCompareSqueezing}(b) for the opposing cases of a pancake-shaped cloud with AR = 0.1 and a pencil-shaped cloud with AR = 316. For comparison, the short-time solution \erf{Eq::ShortTimeSolution} is shown, which describes the squeezing for either cloud in absence of decoherence. To understand these results we separately examine the dynamical evolution of the projection noise variance and the mean spin in Figs. \ref{Fig::SpinHalfCompareSqueezing}(c-d), both of which contribute to the squeezing parameter. The effects of decoherence lead to different steady state values of the fundamental spin wave variance in Fig. \ref{Fig::SpinHalfCompareSqueezing}(c) because the noise injection due to optical-pumping-induced spin flips, set by $N_{\rm eff}^{(3)}$, is slightly smaller for the pencil than for the pancake (see subplot in Fig. \ref{Fig::SpinHalfCompareSqueezing}(a)). More importantly, the decay rate of the mean spin is a strong function of the AR, as seen in \ref{Fig::SpinHalfCompareSqueezing}(d). For a fixed OD$_\text{eff}$, under consideration here, different cloud geometries correspond to different $N_{\rm eff}^{(1)}$, which determines the mean spin of the ensemble addressed by the beam. The pencil geometry address a larger $N_{\rm eff}^{(1)}$ when compared to the pancake geometry, as seen in the subplot of \ref{Fig::SpinHalfCompareSqueezing}(a). In addition, for the pencil geometry $N_{\rm eff}^{(1)}$ also decays more favorably. This occurs because for a fixed OD$_\text{eff}$, in the pencil geometry a large fraction of the atoms are spread far from the beam waist where rates of optical pumping are lower. For the pancake geometry, to achieve the same OD$_\text{eff}$, more of the atoms the we address are concentrated in the high intensity region and more quickly depolarize. \begin{figure}[b] \includegraphics[width=1\hsize]{figure_3} \caption{Squeezing dynamics for a fixed OD$_\text{eff} = 50$ and different atomic cloud geometries. The laser probe is a TEM$_{00}$ mode with beam waist $w_0 = 20$ $\mu$m. a) Peak squeezing, denoted as the inverse of the squeezing parameter, $\zeta^{-1}$ in dB, as a function of aspect ratio of the cloud. The inset shows effective atom numbers as a function of aspect ratio; $N_\text{eff}^{(2)}$ is constant by design. b) Comparison of squeezing dynamics for clouds with AR= 0.1 (green line) and AR $= 316$ (red line). The behavior in the absence of decoherence, \erf{Eq::ShortTimeSolution} (black dashed line), is plotted at the same OD$_\text{eff}$, showing agreement for short times. c) Dynamics of the spin wave variance for the two clouds, normalized by dividing each by its initial variance, $N^{(2)}_{\rm eff}/4$. d) Dynamics of the mean spin for the two clouds, normalized by dividing each by its initial mean spin, $N^{(1)}_{\rm eff}/2$. For fixed OD$_\text{eff}$, the superior squeezing of the pencil-shaped cloud over the pancake-shaped cloud is attributed to slower decay of the mean spin.} \label{Fig::SpinHalfCompareSqueezing} \end{figure} \begin{figure*} \includegraphics[width=1\hsize]{figure_4} \caption{Squeezing for different cloud geometries with Gaussian atomic density distribution, \erf{Eq::AtomicDistribution}, and a fixed total atom number $N = 9.84 \times 10^6$. a) Contours of peak squeezing, $\zeta^{-1}$ in dB, as a function of cloud aspect ratio and laser probe beam waist. b) Contours of the coupling strength, OD$_{\rm eff}$. The difference between the optimal coupling strength and the resulting squeezing depends on the balance between coherent interactions and decoherence, characterized by different effective atom numbers, $N_{\rm eff}^{(1)}$, $N_{\rm eff}^{(2)}$, and $N_{\rm eff}^{(3)}$, shown in (c). } \label{Fig::SpinHalfScaling} \end{figure*} \subsubsection{Optimizing geometry for fixed atom number} We gain further insight into the nature of the atom-light interface by keeping the atom number $N$ fixed and optimizing the cloud dimensions for peak squeezing. We fix the peak density at $\eta_0 = 5 \times 10^{11}$ cm$^{-3}$, typical of dipole-trapped atoms, and keep the total atom number constant, $N = 9.8 \times 10^6$, for a fixed cloud volume. In Fig. \ref{Fig::SpinHalfScaling}(a), we plot contours of peak squeezing as a function of aspect ratio and beam waist. The optimal peak squeezing, $\zeta_{\rm opt}^{-1} = 10.0$ dB, is found for AR $= 256$ at a beam waist of $w^{\rm opt}_0=31$ $\mu$m. At the optimal geometry, the cloud length extends over several Rayleigh ranges, $\sigma_z/z^{\rm opt}_R = 2.42$, and the transverse width of the cloud is slightly larger than the beam waist, $\sigma_\perp/w^{\rm opt}_0 = 1.09$. To further understand the region of peak squeezing, in Fig. \ref{Fig::SpinHalfScaling}(b), we plot contours of OD$_{\rm eff}$. Comparison of Figs. \ref{Fig::SpinHalfScaling}(a-b) shows that the optimal peak squeezing occurs in a parameter region where OD$_{\rm eff}$ is high, as expected. However, the optimal peak squeezing arises from a balance between high OD$_{\rm eff}$ with low noise injection into the spin wave variance and low decay of the mean spin. Figure \ref{Fig::SpinHalfScaling}(c) shows the fraction of total atoms contributing to the mean spin, $N_{\rm eff}^{(1)}/N$, to the effective optical density, $N_{\rm eff}^{(2)}/ N$, and to the noise injection $N_{\rm eff}^{(3)}/ N$. As the cloud becomes too long and narrow, there does not exist a beam waist that can address a sufficiently large number of atoms while keeping a high OD$_{\rm eff}$. Said another way, when the cloud becomes too long, the diffraction of scattered light is too large to effectively mode match with the probe field, as seen in Fig. \ref{Fig::ModeMatching}(c). Similarly, too small a waist leaves many atoms outside the Rayleigh range and too large a waist increases the beam area, thus decreasing OD$_{\rm eff}$, both manifestations of poor mode matching of the probe and the scattered field from the atom cloud. \begin{figure}[b] \includegraphics[width=1\hsize]{figure_5} \caption{ Squeezing for a fixed peak density $\eta_0 = 5 \times 10^{11}$ cm$^{-3}$ and variable atom number that fills a dipole trap for cold atoms. In a), b), and c) the transverse cloud width is fixed at $\sigma_\perp = 100$ $\mu$m and cloud length is taken to be variable. a) Contours of peak squeezing, $\zeta^{-1}$ in dB. b) Contours of OD$_{\rm eff}$. c) Optimal beam waist for maximizing OD$_{\rm eff}$ (red) and for maximizing peak squeezing (blue). For a given atomic geometry, the beam waist that optimizes the OD$_{\rm eff}$ is not the same as that which optimizes peak squeezing. d) Peak squeezing as a function of cloud size for the optimal beam waist at each point.} \label{Fig::Fixed_SigmaT} \end{figure} \subsubsection{Optimizing the beam waist for a fixed atomic cloud geometry} With a better understanding of how cloud geometry influences decoherence, we study the optimization of squeezing in a situation typical of experiments with dipole-trapped cold atoms, where both the trap dimensions and beam waist $w_0$ can be tuned, the peak atomic density $\eta_0$ is fixed, and the total atom number is variable depending on the trap volume. For each cloud geometry there exists a beam waist that maximizes OD$_\text{eff}$. This is seen in Fig. \ref{Fig::Fixed_SigmaT}(b) where contours of OD$_{\rm eff}$ are shown for a cloud with a fixed transverse width of $\sigma_\perp = 100$ $\mu$m as the cloud length $\sigma_z$ and beam waist $w_0$ are varied. Contours for peak squeezing are shown in Fig. \ref{Fig::Fixed_SigmaT}(a). Comparison with \ref{Fig::Fixed_SigmaT}(b) demonstrates that for a given cloud geometry, the peak squeezing is achieved with a smaller beam waist than that which optimizes OD$_\text{eff}$. This is seen most clearly in Fig. \ref{Fig::Fixed_SigmaT}(c), where we compare the optimal beam waist for maximizing OD$_\text{eff}$ to the beam waist that maximizes peak squeezing. Optimal squeezing occurs at smaller beam waists where the region of the beam with greatest intensity, the Rayleigh range, is smaller. Because the scattering rate $\gamma_s(\mathbf{r})$ is proportional to the local intensity, atoms outside the Rayleigh range experience a decreased rate of optical pumping. Although a smaller Rayleigh range implies a decreased OD$_\text{eff}$ and $N_{\rm eff}^{(1)}$ as well, the reduction of the decoherence rate dominates in this regime. This is a direct analogy to Sec. \ref{Sec::GeomEffects}, in which pencil-shaped clouds with higher mean spins were more robust to decay due to a large number of atoms farther away from the beam waist. Finally, in Fig. \ref{Fig::Fixed_SigmaT}(d) we plot contours of peak squeezing for different geometries at the optimal beam waist for each point. \subsubsection{Relation to the symmetric one-dimensional model} \label{SubSec::1Dmodel} Spin squeezing by QND measurement is traditionally modeled using a one-dimensional description of the atom-light interface where the ensemble is symmetrically coupled to plane waves with no spatial variations \cite{PolzikRMP}. When accounting only for squeezing due to collective scattering and QND measurement, the full three-dimensional system can be effectively described by such a model, with the symmetric OD replaced by OD$_\text{eff}$. When decoherence from local diffuse scattering is included, however, such models become insufficient. In addition, a symmetric description does not account for the difference between the effective atom number contributing to the spin wave variance, $N_{\rm eff}^{(2)}$, that contributing to the mean spin, $N_{\rm eff}^{(1)}$, and that contributing to noise injection by spin flips, $N_{\rm eff}^{(3)}$. To better understand the limit in which we recover the simple symmetric description, consider a symmetric 1D model where an ensemble of spin-1/2 atoms is coupled to a uniform plane wave and scatters collectively into this mode and locally into diffuse modes. In this case a single atom number suffices; every atom contributes equally to the optical density, to the mean spin, and to the injection of noise, $ N_{\rm eff}^{(1)} = N_{\rm eff}^{(2)} = N_{\rm eff}^{(3)}= N$. The equation of motion for the spin wave variance \erf{Eq::SpinHalfVariance} becomes \cite{MadMol04, DeuJes09, TraDeu10, NorDeu12} \begin{align} \label{Eq::VarSpin1/2Symmetric} \frac{d}{dt} \Delta F^2_z & = - \kappa \big( \Delta F^2_z \big)^2 - \frac{4}{9}\gamma_0\, \Delta F^2_z + \frac{1 }{9}\gamma_0 \, N, \end{align} where $\gamma_0$ is the scattering rate and $\kappa = (\sigma_0/A)(4\gamma_0/9)$ is the measurement strength corresponding to the rate of scattering into the probe mode per atom, Eq. (\ref{Eq::MeasurementStrengthperAtom}). This equation can be solved analytically, yielding\begin{align} \Delta F^2_z(t) = \frac{N}{4} \, \frac{\sqrt{ \mbox{OD}+1} + \tanh \left[ \frac{(1+\text{OD})}{ \sqrt{\text{OD}}} \frac{2}{9} \gamma_0 t \right] }{\sqrt{\mbox{OD}+1} + \big(\frac{ \text{OD} }{2}+1\big)\tanh\left[\frac{(1+\text{OD})}{ \sqrt{\text{OD}}} \frac{2}{9} \gamma_0 t \right]}. \end{align} In the limit of short times, $ \gamma_0 t \ll 1$, we recover the expression for QND squeezing in the absence of decoherence, \erf{Eq::ShortTimeSolution}, $ \Delta F^2_z(t) \approx (N/4)[1+\xi(t)]^{-1}$, where $\xi (t) = \mbox{OD} \, \gamma_0 t/9$. In the opposite limit of long times, $\gamma_0 t \rightarrow \infty$, and large optical density, OD $\gg 1$, we find the expected scaling $ \Delta F^2_z(t\rightarrow \infty) \propto \mbox{OD}^{-1/2}$\cite{PolzikRMP}. We can compare the symmetric 1D model to a limiting case of the full three-dimensional model developed here. When the transverse extent of the cloud is much smaller than the beam waist and the longitudinal extent is well within the Rayleigh range, then spatial variations of the field across the cloud are minimal and $\beta_{00}(\mathbf{r}_i) \rightarrow 1$. Although this limiting case replicates the squeezing expected from the symmetric 1D model, it is in fact far from a single-mode description. As discussed in Sec. \ref{Sec::QuantumTheory}, this geometry radiates paraxial light into \emph{many} of transverse modes defined relative to the beam, and the associated spin waves couple together through diffuse scattering, \erf{Eq::ProjectedSpin1/2VarianceEOM}. After numerically solving these coupled equations according to the procedure outlined above, we recover the same results Eq. (\ref{Eq::VarSpin1/2Symmetric}), as is manifest from Eq. (\ref{Eq::SpinHalfVariance}) in the limit that $\beta_{00}(\mbf{r}_i) \rightarrow 1$ over the extent of the atom cloud. \begin{figure} \includegraphics[width=.9\hsize]{figure_6} \caption{ Comparison between the symmetric 1D model and the three-dimensional spin wave model. a) Peak squeezing, $\zeta^{-1}$ in dB, for a spherical cloud with OD$_{\rm eff} = 50$ as the beam waist is increased. Inset shows the convergence of $N_{\rm eff}^{(1)}$ (blue), $N_{\rm eff}^{(2)}$ (green), and $N_{\rm eff}^{(3)}$ (red) as $w_0$ increases. b) Comparison of squeezing dynamics for the extremal waists from a): the smallest $w_0 = 10$ $\mu$m (red line) and the largest $w_0 = 10^4$ $\mu$m (green line). For comparison, the symmetric 1D case using \erf{Eq::VarSpin1/2Symmetric} is plotted with decoherence (black dotted line) and without (black dashed line).} \label{Fig::PlaneWaveComp} \end{figure} We investigate this limit numerically in Fig. \ref{Fig::PlaneWaveComp} for a spherical cloud ($\sigma_\perp, \sigma_z = 100$ $\mu$m) probed by beams of increasing waist $w_0$. In each case OD$_{\rm eff} = 50$, such that in the absence of decoherence the different geometries would achieve identical squeezing. In Fig. \ref{Fig::PlaneWaveComp}(a) we see that as the beam waist is increased, the peak squeezing approaches that of the symmetric 1D model. The inset shows the convergence of the effective atom numbers as the beam waist increases. Figure \ref{Fig::PlaneWaveComp}(b) shows the dynamics of the squeezing parameter $\zeta^{-1}(t)$ for the spherical cloud at both extremes in Fig. \ref{Fig::PlaneWaveComp}(a). For comparison, the squeezing parameter for the symmetric 1D model is plotted both with and without decoherence. The difference between the models is substantial -- the optimal peak squeezing for the symmetric 1D model and full model are $\zeta^{-1}_{\rm peak} = \{ 3.52 \text{ dB}, 4.99 \text{ dB} \}$, respectively. This difference can be understood in terms of the effective atom numbers. The advantage for spin squeezing in the three-dimensional model comes from the fact that $N_{\rm eff}^{(1)} \geq N_{\rm eff}^{(2)} \geq N_{\rm eff}^{(3)}$ due to different dependence on the spatial weightings $\beta_{00}(\mathbf{r})$, while for the symmetric 1D case they are equal. For the three-dimensional model, not only can the effective number of atoms contributing to the noise injection be smaller than that contributing to the OD$_{\rm eff}$, but the effective number of atoms contributing to the mean spin, and thus the signal, is larger than both. Inspecting Fig. \ref{Fig::SpinHalfCompareSqueezing}(d) we see an additional advantage for the three-dimensional model -- when geometry is properly chosen, the mean spin decays at a much reduced rate. \subsubsection{Spin $f>1/2$ atoms} We present here some initial results that illustrate the differences between spin-1/2 and larger spin ensembles. Consider an ensemble of $^{133}$Cs atoms prepared in the $f=4$ ground hyperfine manifold. Figure \ref{Fig::Spin4Scaling}(a) shows contours of peak squeezing as a function of aspect ratio using the atomic density given by \erf{Eq::AtomicDistribution}. Note that the peak squeezing is substantially smaller than the peak squeezing for spin-1/2 (compare to Fig. \ref{Fig::SpinHalfScaling}(a)) and the optimal aspect ratio and beam waist are different for spin-4 alkali atoms than for spin-1/2 atoms. This can be attributed to a reduction of the coupling strength $\xi$ that is not compensated by an equal reduction in the decoherence rate. In principle, the coupling strength can be increased by internal spin control of the $f=4$ hyperfine spin \cite{NorDeu12}. Figure \ref{Fig::Spin4Scaling}(b) compares the squeezing dynamics for spin-4 and spin-1/2 ensembles at the same geometry and illustrates not only the disparity in peak squeezing but also in the time at which it occurs. These effects arise from the subtle interplay between the rates of depolarization and injected noise due to spin flips when applied to the spatial modes of larger spin ensembles. \begin{figure}[b] \includegraphics[width=1\hsize]{figure_7} \caption{ Squeezing of a spin-4 ensemble. a) Contours of peak squeezing for an ensemble of spin-4 atoms for the same fixed atom number as in Fig. \ref{Fig::SpinHalfScaling} as function of beam waist and cloud aspect ratio. The optimal squeezing, $\zeta^{-1} = 7.8$ dB, is achieved for a geometry AR = 300, $w_0 = 28$ $\mu$m. b) Dynamics of the inverse of the squeezing parameter for the optimal geometry in a) for spin-4 (red line) and for spin-1/2 (green line) } \label{Fig::Spin4Scaling} \end{figure} \section{Summary and Outlook} \label{Sec::Conclusion} The entangling power of the quantum interface between photons and an ensemble of cold atoms is at the heart of a variety of quantum information processing tasks. When considering extended atomic clouds in dipole traps, one must consider a full three-dimensional description of the electromagnetic modes and atomic density distribution in order to optimize this entangling power. Inhomogeneous coupling between atoms and photons is essential to maximize the strength of the quantum interface, but this comes with substantial complexity in the theoretical description. The model presented in this work addressed these issues and examined regimes of optimal performance. We have studied the strength of the atom-photon interface in a traveling wave configuration in the context of spin squeezing via QND measurement. We developed a description in terms of quantized paraxial modes of the field in order to model the inhomogeneous atom-light coupling across the atomic ensemble, which leads to two distinct effects. First, the collective coupling describes a generalization of the Faraday interaction that entangles the quantized Stokes vector of the laser field with a spin wave defined by the weighted ensemble of atoms that indistinguishably radiates into the mode of the probe. The spin wave that is squeezed is defined by the probe mode we measure in a balanced polarimeter. Second, diffusely scattered photons lead to optical pumping and decoherence across the ensemble at a rate proportional to the local probe intensity. The delicate balance of these two effects favors certain geometries for spin squeezing. We have numerically investigated the ultimate limits to spin squeezing based on a stochastic master equation, including the effects of QND measurement backaction and decoherence by photon scattering into unmeasured modes. Unlike the usual one-dimensional description in which the amount of squeezing is set by a single parameter, the optical density, we find that due to inhomogeneous coupling, multiple parameters are required. Of particular importance in a metrological context are the mean collective spin and the projection noise variance, determined by effective atom numbers $N^{(1)}_\text{eff}$ and $N^{(2)}_\text{eff}$ respectively. Optimal geometries maximize the effective optical density, OD$_\text{eff}$, proportional to $N^{(2)}_\text{eff}$, while minimizing the depolarization of $N^{(1)}_\text{eff}$ and injected noise into the spin wave by optical pumping. We found that optimal mode matching occurs for geometries where a large number of atoms are addressed by a beam with a small transverse area, yielding a high OD$_{\rm eff}$, but also where the depolarization rate due to optical pumping is relatively small. This geometry corresponds to a longitudinally extended, pencil-shaped cloud, with a probe beam chosen to optimize the tradeoffs between OD$_\text{eff}$ and decoherence. Such a geometry is far from the regime describing squeezing of a symmetric atomic spin ensemble, as is typically assumed. One recovers the symmetric description only for ensembles confined with extents much smaller than the beam waist and Rayleigh range, which yield much smaller OD$_\text{eff}$. While the three-dimensional model developed in this work was specifically tailored to study the problem of spin squeezing by QND measurement, it can be extended to other protocols involving the quantum interface between photons and free-space atomic ensembles. Mode-matching and spatial effects are important for other spin squeezing protocols including the double-pass counter-twisting interaction \cite{TakTak05,TraDeu10} or the recently proposed planar squeezing protocol \cite{PueMit13}. Understanding spatial effects in order to identify regimes of strong coupling is also essential for quantum memories and repeaters in free-space atomic ensembles. In addition, recent work has considered ensembles of higher-spin alkali atoms, in which control over the rich internal hyperfine structure can enhance the entangling strength of the atom-light interface \cite{NorDeu12}. Quantifying the gains achievable though such control techniques requires a realistic description of the inhomogeneous interaction between light and atoms. Finally, a multimode description of the entangling Hamiltonian offers the possibility to exploit spatial modes and their associated spin waves as a resource. The creation of entanglement between spin waves could lead to novel states with potential application in continuous variable quantum computation and communication \cite{CV2007}. \emph{Acknowledgments}. We thank Robert Cook for helpful discussions and insights. This work was supported by NSF Grants PHY-0969391, 0969997, 1212445, 1306171, 1307520, and AFOSR Grant No. Y600242.